Vous êtes sur la page 1sur 37

Science & Education 1, 11-47, 1992.

11
9 1992 Kluwer Academic Publishers. Printed in the Netherlands.

History, Philosophy, and Science Teaching:


The Present Rapprochement 1

M I C H A E L R. MATTHEWS

Education Department, Auckland University, Auckland, New Zealand

ABSTRACT. This paper traces the use of, and arguments for, the history and philosophy of
science in school science courses. Specific attention is paid to the British National Curriculum
proposals and to the recommendations of the US Project 2061 curriculum guidelines. Some
objections to the inclusion of historical material in science courses are outlined and answered.
Mention is made of the Piagetian thesis that individual psychological development mirrors
the development of concepts in the history of science. This introduces the topic of idealisation
in science. Some significant instances are itemised where science education has, at its con-
siderable cost, ignored work in philosophy of science. Arguments for the inclusion of the
history and philosophy of science in science teacher education programmes are given. The
paper finishes with a list of topical issues in present science education where collaboration
between science teachers, historians, philosophers, and sociologists would be of considerable
benefit.

INTRODUCTION

In 1986 a paper was published titled 'Science Education and Philosophy of


Science: Twenty-five Years of Mutually Exclusive Development' (Duschl
i986). This was an account of the largely separate development of science
education from the disciplines of history and philosophy of science. In the
last five years there has been a significant rapprochement between these
fields. Both the theory, and importantly the practice, of science education
is becoming more informed by the history and philosophy of science.
Given the well-documented crisis of contemporary science education -
evidenced in the flight from the science classroom of both teachers and
students, and in the appaUingly high figures on science illiteracy (Matthews
1988) - these initiatives are timely. The American National Science Foun-
dation charged that 'the nation's undergraduate programs in science, math-
ematics and technology have declined in quality and scope to such an
extent that they are no longer meeting national needs. A unique American
resource has been eroded' (Heilbron 1987, p. 556). History, philosophy,
and sociology of science does not have all the answers to this crisis but it
has some of the answers: it can humanise the sciences and make them
more connected with personal, ethical, cultural, and political concerns; it
can make classrooms more challenging and thoughtful and thus enhance
critical thinking skills; it can contribute to the fuller understanding of
scientific subject matter - it can contribute a little to overcoming the 'sea
of meaninglessness' which one commentator said has engulfed science
classrooms, where formula and equations are recited but few people know
12 MICHAEL R. M A T T H E W S

what they mean; it can improve teacher training by assisting the develop-
ment of a richer and more authentic epistemology of science, that is a
greater understanding of the structure of science, and its place in the
intellectual scheme of things. This last matter is the beginning of the sort
of discipline understanding that Schulman (1987) and before him, but
largely neglected, Scheffler (1970) have been urging teacher education
programmes to promote.
There are a number of elements in the rapprochement. By far the most
significant is the inclusion of history and philosophy of science components
in various national school curricula. This is occurring in the national
science curriculum for England and Wales, in the American Project 2061
recommendations for high-school science, in the Danish national school
curriculum, and in the Dutch PLON curricular materials. It is not the
inclusion of HPS as another item of subject matter, but rather the more
general incorporation of HPS themes in the approach to the subject mat-
ter, and teaching, of the curricula. Science curricula have usually included
a section titled 'The Nature of Science': more attention is now being
paid to these sections, and it is increasingly recognised that the history,
philosophy and sociology of science contribute to a better, fuller, and
richer understanding of the questions posed in these sections. One major
opening for historical and philosophical contributions to science education
is in the widespread Science, Technology and Society (STS) programmes
in both schools and universities. These developments have significant
implications for teacher training.
There are other elements and indicators of a rapprochement. One has
been the staging of the first international conference on 'HPS and Science
Teaching' at Florida State University in November 1989. 3 A second is the
series of conferences sponsored by the European Physical Society on
'History of Physics and Physics Teaching' in Pavia (1983), Munich (1986),
Paris (1988), and Cambridge (1990). 4 A third has been the conference
held by the British Society for the History of Science on 'History of Science
and Science Teaching' held at Oxford University in 1987 (Shortland &
Warick 1989). These activities have generated about 300 scholarly papers
on the subject, as well as a great deal of historically and philosophically
informed teaching material. Additionally the American National Science
Foundation has now commenced two programmes to promote HPS in
school science teaching. Some US science teacher education programmes
have made HPS compulsory. One state (Florida) has made completion of
a HPS course necessary for being licensed as a science teacher.
The advocates of HPS in science teaching and in teacher training are
in part arguing for a 'contextualist' approach to science education. That
is a science education that teaches science by teaching it in its social,
historical, philosophical, ethical, and technological contexts. In part this
is a reworking of the old argument that science education should be an
education about science as well as in science. To use the terminology of
the British National Curriculum, school science students should learn
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 13

something about the 'Nature of Science' as well as the content of current


science.
The arguments or rapprochement are in many ways repeating the calls
that commenced with Mach 5 in the late nineteenth-century and were
repeated by Duhem in the early twentieth-century. Such calls have been
made again by people like Nunn, Conant, Holton, Robinson, Schwab,
Martin, and Wagenstein. They can be seen in numerous British and Amer-
ican reports.

CURRICULA REFORMS

The new British National Curriculum in Science, and the American AAAS
2061 Project are worth discussing at the outset of any review of HPS
and science education because they so clearly show the classroom and
programme implications of the rapprochment. The former is a course
already in place; the latter is a long-planned, and comprehensive set of
proposals for a new American school science curriculum.
The British National Curriculum Council introduction to the HPS lobe
of the course (consisting of about 5% of the total programme), says that:

pupils should develop their knowledge and understanding of the ways in which scientific
ideas change through time and how the nature of these ideas and the uses to which they
are put are affected by the social, moral, spiritual and cultural contexts in which they are
developed (NCC 1988, p. 113).

As an example of the type of skills and understanding the NCC are


trying to promote in the new curriculum, its statements about abilities 4-
16 year-olds should have are illustrative. They should be able to:
9 distinguish between claims and arguments based on scientific data and
evidence and those which are not;
9 consider how the development of a particular scientific idea or theory
relates to its historical and cultural, including the spiritual and moral,
context;
9 study examples of scientific controversies and the ways in which scientific
ideas have changed (NCC 1988, p. 113).
The American Association for the Advancement of Science (AAAS)
established in 1985 an extensive national study, Project 2061, to recom-
mend an overhaul of school science. In 1989, after four years of deliber-
ation and consultation, the project published its recommendations in a
report titled Science for All Americans (AAAS 1989). Project 2061 was
independent of the British NCC deliberations, yet there is considerable
agreement about the need for school science to be more contextual, more
historical, and more philosophical or reflective.
Science for All Americans contains 12 chapters giving the recommen-
dations for school science of the National Council On Science and Tech-
nology Education. Chapter One is on 'The Nature of Science'. It includes
14 MICHAEL R. MATTHEWS

discussions of objectivity, the mutability of science, the possible ways to


demarcate science from pseudo-science, evidence and its relation to theory
justification, scientific method, explanation and prediction, ethics, social
policy, and the social organisation of science. The intention is that these
themes be developed and discussed within science courses and that pupils
completing school science know something of them; the intention is not
that the topics be added to science courses and HPS be substituted for
science content knowledge.
The introduction to Chapter Ten on 'Historical Perspectives' says:
'There are two principal reasons for including some knowledge of history
among the recommendations. One reason is that generalizations about
how the scientific enterprise operates would be empty without concrete
examples.' It goes on to say that: 'A second reason is that some episodes
in the history of the scientific endeavour are of surpassing significance to
our cultural heritage. Such episodes certainly include Galileo's role in
changing our perception of our place in the universe'.
The report devotes one-and-a-half pages to the Galileo episode - 'Dis-
placing the Earth from the Center of the Universe'. It is an informed and
sensitive treatment of astronomical evidence, the role of sense perception,
mathematical models, realism and instrumentalism, metaphysics, technol-
ogy, rhetoric and theology. It deals with other historical episodes in a
similar style.
In both the British and US curricula proposals it is not envisaged that
a rhetoric of conclusions about science be replaced by a rhetoric of con-
clusions about HPS. No one expects children to solve the realism/instru-
mentalism debate, nor should they learn catechism-like that there were
fifteen reasons why Galileo was right and the cardinals wrong. Rather
they are expected to appreciate something of the intellectual issues that
are at stake in these matters; to appreciate that there are questions to ask
and to begin to think not just about answers, but what would count as
answers, and what kinds of evidence would support our answers.
To convert curricular plans into classroom realities will require new
assessment guidelines and practice, new teaching materials, and above all
the inclusion of appropriate HPS courses in teacher training. They will
also require, as is explicitly stated in the Project 2061 proposals, the scaling
down of the content of curricula that are 'overstuffed and undernourished'
(AAAS 1989, p. 14). A study of the New York State Biology syllabus shows
the dimension of this problem: teachers are required to introduce 1,440
new scientific terms and concepts in one year (Swift 1988). There is now
greater appreciation of what Mach a century ago advocated: the teaching
of less, and the understanding of more. Mach put the matter this way: 'I
believe the amount of matter necessary for a useful education . . . . is very
s m a l l . . . I know nothing more terrible than the poor creatures who have
learned too m u c h . . . What they have acquired is a spider's web of
thoughts too weak to furnish sure supports, but complicated enough to
produce confusion' (Mach 1943, p. 366).
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 15

HISTORY IN THE SCIENCE CURRICULUM

In Britain there has been a long, if thin and uneven, tradition of incorpor-
ating the history of science in science education. This has been well
documented by Edgar Jenkins (1989, 1990) and W. J. Sherratt (1982,
1983). According to Jenkins the first clear statement can be found as far
back as 1855 in the Duke of Argyll's Presidential Address to the British
Association for the Advancement of Science, where he said 'What we
want in the teaching of the young, is, not so much the mere results, as
the methods and, above all, the history of science' (Jenkins 1989, p. 19).
The BAAS repeated these calls at its 1917 conference where it remarked
that the history of science supplied a solvent of that artificial barrier
between literary studies and science which the school timetable sets up
(Jenkins 1989, p. 19). Mach, and those influenced by him, also argued
that in order to understand a theoretical concept it was necessary to
understand its historical development; understanding was necessarily his-
torical. Mach in his 1883 classic said that:
The historical investigation of the development of a science is most needful, lest the
principles treasured up in it become a system of half-understood precepts, or worse, a
system of prejudices. Historical investigation not only promotes the understanding of that
which now is, but also brings new possibilities before us. (Mach 1883/1960, p. 316) 6

Nunn and other historically-minded educationalists argued this case in


the inter-war years. A number of examining boards ran separate courses
in the history of science, but by the 1980's the number of candidates
presenting was meagre. Prior to the National Curriculum the history of
science found some place in the Nuffield programmes, and in the recently
introduced SISCON and SATIS courses.
Through the 1970's and 80's, the British Association for Science Edu-
cation in a number of its reports (Alternatives for Science Education 1979,
Education through Science 1981), urged the incorporation of more histori-
cal and philosophical material into the science curriculum. One problem
that it recognised was that teachers were not adequately prepared to teach
this contextual science. In its 1963 report (The Training of Graduate
Teachers), it said of graduate teachers that 'Many behave and think scien-
tifically as a result of their training but they lack an understanding of the
basic nature and aims of science' (p. 13). This point was remade in a 1981
review of the place of philosophy of science in British science-teacher
education (Manuel 1981).
In the United States after the second world war, the history of science
figured prominently in college science courses for non-science majors. The
dominant influence here was the president of Harvard University, James
B. Conant, whose case-study approach was widely adopted. He had de-
veloped this whilst in charge of undergraduate general education at Har-
vard, and popularised it in a series of government reports and paper-back
best-sellers most notably Understanding Science: An Historical Approach
16 MICHAEL R. MATTHEWS

(1947). His two volume Harvard Case Histories in Experimental Science


(1957) became a text in many science courses.
Conant's influence cannot be overestimated. Kuhn says that it was
Conant 'who first introduced me to the history of science and thus initiated
the transformation in my conception of the nature of scientific advance'
(Kuhn 1970, p. xi; the relationship of Kuhn to Conant is traced in Merton
1977, pp. 81-89). The subsequent post-Kuhnian transformations in the
history and philosophy of science are well known. Gerald Holton makes
a similiar admission, and like Kuhn, says that the experience of being a
scientist who suddenly had to teach a Harvard General Education course
built around the history of science was a Pauline experience. Leo Klopfer
adopted the Harvard Case Studies for use in schools, and they were trialled
with considerable success (Klopfer and Cooley 1963). Gerald Holton was
subsequently instrumental in developing in the early 1960's, with Stephen
Brush, Fletcher Watson, James Rutherford, and others, the Harvard Pro-
ject Physics course for secondary schools. The Conant-Harvard connec-
tion is carried through to the 1980's with James Rutherford's appointment
as director of the Project 2061 programme of the AAAS.
I. Bernard Cohen, the prominent Harvard historian of science, also
argued for the introduction of historical material into college science
programmes. He organised a symposium on the topic at the 1950 annual
conference of the American Association of Physics Teachers. He contrib-
uted the lead paper, 'A Sense of History in Science' (1950) which took a
number of standard treatments of historical episodes from science texts
and pointed out the many inaccuracies in these accounts. This distortion
of the historical record extended up to twentieth century physics. He cited
the common, yet mistaken, accounts of J. J. Thompson's experiments to
determine the ratio of e/m. Cohen gives advice on acquiring better knowl-
edge of the history of science, urges teachers to try their hand at writing
history, and maintains that a historical sense will make lectures 'richer
and more profound and a greater interest t o . . . students'.
However, as is welldocumented (Duschl 1985), the major science cur-
ricular reforms of the 1960's proceeded largely without the participation
of either historians or philosophers of science (many of them proceeded,
one might add, without school teachers). There were two notable excep-
tions, one was the previously mentioned Harvard Project Physics Course,
the other, versions of the American Biological Science Curriculum Study
(BSCS).
The Harvard Project Physics course, which at its peak accounted for
about 15% of high school physics students in the US, has been the most
widely used school science curriculum based upon historical principles and
concerned with the cultural and philosophical dimension of science. Its
success in retaining students, involving women in science courses, de-
veloping critical reasoning skills, and raising scores on attainment tests,
has provided evidence for the contemporary advocates of HPS. This evi-
dence is reviewed in Aikenhead (1974), Holton (1978), Russell (1981),
Brush (1989), and the symposium on Project Physics in The Physics
H I S T O R Y , P H I L O S O P H Y , AND S C I E N C E T E A C H I N G 17

Teacher 5 2, (1967). The lessons from the failures of Project Physics are
also instructive, in particular the failure to adequately acquaint teachers
with HPS so that they could cope with the curriculum in the open-ended
and critical manner expected. This has already been identified as a serious
problem for implementation of the 'Nature of Science' component of the
British National Curriculum (NCC 1988, p. 21).
The BSCS was informed by the ideas of the biologist-philosopher-
educationalist J. J. Schwab 7 who promoted the pedagogical creed of 'sci-
ence as enquiry'. Schwab wrote the Teachers' Handbook for the BSCS
curriculum, and in it he advocated the historical approach, saying that
'the essence of teaching of science as enquiry would be to show some of
the conclusions of science in the framework of the way they arise and are
tested . . . . [it] would also include a fair treatment of the doubts and
incompleteness of science" (1963, p. 41). History is also advocated because
it 'concerns man and events rather than conceptions in themselves. There
is a human side to enquiry' (1963, p. 42).
In the early 1960's the International Commission on Physics Education
raised the usefulness of the history of physics for physics education. By
1970 a symposium on the subject was convened at MIT under the direction
of Stephen Brush and Allen King. The proceedings were published (Brush
& King 1972) and contained a significant challenge, by Klein, to the whole
enterprise of using history of physics in physics teaching. This will be
discussed in the next section.
In the 1970's the American Physical Society established a section on
the History, of Physics, at the same time the History of Science Society
established an Education Committee. Both have continued to be active in
educational matters. Stephen Brush has been influential in both, producing
numerous historical studies for use in classrooms?
The history of chemistry in the US has always been more marginal to
pedagogy than the history of physics or biology. There have been many
calls for the inclusion of history of chemistry into chemical education;
these have been documented by Kaufmann (1989).
In other countries the history of science has had a similarly chequered
career in science pedagogy. Brief accounts of the European experiences
can be found in papers contained in Thomsen (1986). Teichmann at the
Deutsches Museum in Munich has replicated historical experiments and
provided notes and instruction for teachers; a similiar effort has been
made by Bevilacqua and associates at the University of Pavia. Krasilchik
(1990) discusses an interesting approach to the subject in Brazil; Tamir
(1989) discusses the situation in Israel.

HISTORY UNDER ATFACK

The contextualist tradition asserts that the history of science contributes


to science teaching because: (1) it motivates and engages pupils; (2) it
humanises the subject matter; (3) it promotes the better comprehension
18 MICHAEL R. MATTHEWS

of scientific concepts by tracing their development and refinement; (4)


there is intrinsic worth in understanding certain pivotal episodes in the
history of science - the Scientific Revolution, Darwinism etc; (5) it demon-
strates that science is mutable and changeable, and that consequently
current scientific understanding is liable to be transformed, which (6)
thus combats scientistic ideology; and finally, (7) history allows a richer
understanding of scientific method and displays the patterns of change in
accepted methodology.
In 1970, at the above-mentioned MIT conference, these justifications
for history came under attack from two fronts: firstly, it was said that the
only history possible in science courses was pseudo-history; secondly, it
was said that exposure to history of science weakened the scientific convic-
tions required for successful completion of an apprenticeship in science.
The first case was argued by Martin Klein (1972); the second was in part
prompted by the analysis offered in Thomas Kuhn's classic The Structure
of Scientific Revolutions (first edition 1962, second edition 1970).
Klein's argument was basically that teachers of science (more parti-
cularly physics) select and use historical materials to further scientific or
pedagogical purposes:

We are, in other words, planning to select, organize, and present these historical materials
on decidedly nonhistorical, perhaps even antihistorical grounds. This is a very hazardous
enterprise, if we are as concerned about the integrity and quality of the history we teach
as we are about the physics (Klein 1972, p. 12).

He goes on to say:

One reason it is difficultto make the history of physics serve the needs of physics teaching
is an essential difference in the outlooks of physicist and historian. . . . . It is so hard to
imagine combining the rich complexity of fact, which the historian strives for, with the
sharply defined simple insight that the physicist seeks. (Klein 1972, p. 16)

His conclusion is that if good science teaching is historically informed,


then it will only be able to use bad history. He prefers no history to bad
history.
Whitaker (1979) pushed these claims further in a paper titled 'History
and Quasi-history in Physics Education'. His concern was to identify the
prevalent fabrication of history to suit not just pedagogical ends, but the
ends of scientific ideology, or the view of science held by the writer. These
cases abound in textbooks. One that has been much discussed is the
widespread account of how Einstein's theory of special relativity was
motivated by the null result of the Michelson-Morley experiment. This is
a Popperian-inspired myth. Another myth is the prevalent view that Ein-
stein's postulation of the photon followed experiments on the photo-
electric effect, rather than preceded the experiments. Among countless
other examples is one presented on the first few pages of the PSSC Physics
text describing Galileo's discovery of the law of isochrony of the pendulum
by timing the swings of the chandelier in the church at Pisa. This dubious
H I S T O R Y , P H I L O S O P H Y , AND S C I E N C E T E A C H I N G 19

account is used by PSSC as the very model of scientific methodology. The


problems with it are discussed in Matthews (1987).
Whitaker says of quasi-history that it is the 'result of the large numbers
of books by authors who have felt the need to enliven their account of
[these episodes] with a little historical background, but have in fact re-
written the history so that it fits in step by step with the physics' (Whitaker
1979, p. 109).
Quasi-history is not just Klein's pseudo-history, or simplified history,
where mistakes of omission are likely to occur, or where the story might
fall short of the lofty standard of 'the truth, the whole truth, and nothing
but the truth'. In quasi-history we have manufactured history masquerad-
ing as genuine history. This is akin to Lakatos' 'rational reconstructions'
of history (1978) where history is written to support a particular version
of scientific methodology and where historical figures are painted in the
hues of the current methodological orthodoxy.
Quasi-history is a complex matter. We know that objectivity in history
is, at one level, impossible: history does not just present itself to the eye
of the beholder; it has to be manufactured. Materials and sources have
to be selected; questions have to be framed; decisions about the relevant
contributions of internal and external factors in scientific change have to
be made. All of these matters are going to be influenced by the social,
national, psychological, and religious views of the historian. More impor-
tantly they are going to be influenced by the theory of science, or the
philosophy of science, held by the historian. Just as a scientist's theory
affects how they see, select, and work upon their material, so also will a
historian's theory affect how they see, select, and work upon their ma-
terial. As has been said by many people, if philosophy of science without
history of science is empty, then history of science without philosophy of
science is blind. 9
The history of interpretation of Galileo's achievements and methodol-
ogy, and the changing translations of his works, illustrates the problem
of theory affecting how historical events and documents are seen. To
nineteenth-century philosophers and scientists Galileo was an inductivist
and empiricist. William Whewell says of him that he 'had perhaps a
preponderating inclination towards facts, and did not feel, so much as
some other persons of his time, the need of reducing them to ideas'
(1840/1947, p. 220). David Brewster in 1830 saw him as a Baconian figure
and held that 'the principles of inductive philosophy' could be found in
his work (Finocchiaro 1980, p. 152). When positivism was in its ascendancy
Galileo was portrayed in texts as a positivist. Mach says of him that
~Galileo did not supply us with a theory of the falling of bodies, but
investigated, wholly without preformed opinions, the actual facts of falling'
(Mach 1883/1960, p. 167). In the 1930's a thunder-clap broke over empiri-
cist readings of Galileo when Alexandre Koyr6 announced that Galileo
was really a Platonist (1939, 1943). The history of Platonic interpretation
has been reviewed in McTighe (1967). William Shea's (1972) rationalist
20 MICHAEL R. M A T T H E W S

account of Galileo has much in common with Koyr6's. But empiricism


and rationalism by no means exhaust the interpretative field. William
Wallace has written a series of studies placing Galileo in the late-Scholastic
Aristotelian tradition (1981, 1984); a view suggested in Randall's study of
the School of Padua (1940). The picture that emerges from Stillman
Drake's many articles and books (1978, 1980) is that of Galileo as the
patient experimentalist. Paul Feyerabend's recent anarchist, or dadaist,
interpretation of Galileo is well known; it is the linch-pin of his own case
against the primacy of any single scientific method (1975).
One reviewer of the contrasting work of Koyr6 and Drake has written
that:
Koyr6's Galileo seemed to inhabit a largely philosophical world - of Platonism, Coper-
nicanism, rationalism, and thought experiments. Drake's Galileo on the other hand, is
more active and less contemplative.., a keen observer, experimenter, and inven-
t o r . . . These considerable differences in the conclusions of Drake and Koyr6 result largely
from differences is their styles of approach. (MacLachlan 1990, p. 124)

That an interpretation of a text and event reflects the opinions of the


age or interpreter was something highlighted by late nineteenth-century
biblical scholars, the critical school of Renan and others. Albert Schweitz-
er's monumental The Quest for the Historical Jesus (1910) after reviewing
the history of images and theologies of Jesus concluded that 'each age and
each individual constructed lives of Jesus in their own image' - a con-
clusion illustrated sixty years later when the American evangelist Billy
Graham claimed that 'Jesus was the greatest general manager that the
world has known'.
The translation of Galileo's work had also, understandably, been affec-
ted by the presuppositions of the translator. I. Bernard Cohen has drawn
attention to the quite gratuitious insertion of 'by experiment' into the text
of the Discourse on Two New Sciences in Crew and de Salvio's translation
where Galileo mentions that he has discovered some hitherto unknown
properties of motion (1977). Peter Machamer (1978) notes that Drake's
translation of the same work renders Galileo's 'ratio' as 'rule', when
the original text implies 'formal cause'. As Drake believed that Galileo
abandoned the notion of final cause, it is not suprising that the term is
not so translated.
The problem is of course deeper than just interpretation affecting per-
ception. Bacon long ago, with his discussion of the Idols of the Mind,
recognised how personal and cultural assumptions, including language,
affect vision and understanding. Bacon gave empiricist counsel, saying
that we ought to minimise the extent of these biases, and see the world
as it actually is. Such advice is now recognised as being a little too simple:
not just literature, history and politics, but natural science also has its
hermeneutical problem, a consideration argued by Toulmin (1983) and
Markus (1987). The fairly orthodox historian of science, A. C. Crombie,
has recognised the necessity for hermeneutical interpretation in the histori-
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 21

ography of science in his own study of 'Philosophical Presuppositions and


Shifting Interpretations of Galileo' (1981).
The second charge brought against genuine history of science in science
courses was that it can sap the neophyte scientific spirit. This charge was
developed by Thomas Kuhn among others. In a 1959 essay on science
education and its psychological and intellectual effects he said:

The single most striking feature of this education is that, to an extent totally unknown in
other creative fields, it is conducted entirely through t e x t b o o k s . . , nor are science students
encouraged to read the historical classics of their fields - works in which they might
discover other ways of regarding the problems discussed in their t e x t b o o k s . . , it remains
a dogmatic initiation in a pre-established tradition. (Kuhn 1959/1977, p. 228-229)

This initiation Kuhn says is necessary because 'no part of science has
progressed very far or very rapidly before this convergent educa-
t i o n . . , became possible' (p. 237). Kuhn advanced these ideas about the
virtue of a conformist science education in his most influential The Struc-
ture of Scientific Revolutions where he says that in a science classroom the
history of science should be distorted, and earlier scientists should be
portrayed as working upon the same set of problems that modern scientists
work upon (1970, p. 138). This distortion aims to make the apprentice
scientist feel part of a successful truth-seeking tradition: 'Textbooks thus
begin by truncating the scientist's sense of his discipline's history and then
proceed to supply a substitute for what they have eliminated' (p. 137).
Stephen Brush developed the Kuhn charge further in his 'Should the
History of Science be Rated X?' (1974). Here it was suggested that history
of science could be a bad influence on students because it undercuts the
certainties of the scientistic dogma so useful for maintaining the enthusi-
asm of the apprentice. He perhaps tongue-in-cheek suggests that history
should be confined to mature scientific audiences.
Kuhn's points can be traced back to reactions to Poincarr's instrumental-
ist account of science at the turn of the century. Heilbron recounts how
the President of the British Association for the advancement of Science
in 1901 said of Poincarr's theory of science that 'If the confidence that his
methods are weapons with which he can fight his way to the truth were
taken from the scientific explorer, the paralysis of those engaged in a
hopeless task would fall upon him' (Heilbron 1983, p. 178).
The Klein-Kuhn charges are serious, but their main points can be
accommodated without jettisoning history from science courses. In peda-
gogy, as in most things, the subject matter frequently needs to be simpli-
fied. This is as true of history of science as it is of economics, or science
itself. The fact that history of science is simplified is no conclusive strike
against it. The pedagogical task is to produce a simplified history that
illuminates the subject matter, yet is not a caricature of the historical
process. The simplification will be relevant to the age group being taught,
and the overall curriculum being presented. The history and the science
can become more complex as the educational situation demands. The
22 M I C H A E L R. M A T T H E W S

problem of gross distortion is best dealt with by the better presentation


of HPS in pre-service and in-service training: most distortions are made
in good-faith. The hermeneutical problem of interpretation in the history
of science, far from being an embarrassment or impediment to the use of
history, can be the occasion to introduce students to the significant ques-
tions of how we read texts and interpret events, to the complex problems
of meaning: students know from their everyday life that people see things
differently, the history of science is a natural vehicle for illustrating how this
fact impinges on science itself.
The proof of a pudding is in its eating. Historical studies have shown
their worth for science teachers: Arons (1988), Pumpfrey (1989), Bevilac-
qua (1990) and the Proceedings of the 1983, 1986, and 1988 European
Physical Society conferences all contain the sort of applied, and pedagogi-
cally useful, history of science that Heilbron has urged historians to collab-
orate with educators in producing. There is no evidence that such ap-
proaches diminish scientific understanding; they may diminish a certain
scientistic conviction but that is maybe no bad thing. The success of
Harvard Project Physics is an outstanding counter example to the Kuhn-
Brush "worries: there it is demonstrated that a good science education,
without indoctrination, is in fact possible (Siegel 1979).

HISTORY OF SCIENCE AND THE PSYCHOLOGY OF LEARNING

A significant part of the recent literature on HPS and Science Teaching


has been concerned with the conjunction of the history of science and the
psychology of learning. More specifically, in what ways do individual
cognitive development and the process of historical conceptual develop-
ment shed light upon each other? This issue has had a long history; it has
received a new impetus from the mushrooming cognitive theories of sci-
ence in which the concepts and the methods of cognitive science have
been used to study the processes and history of science (Giere 1987, Jung
1986, Nessersian 1989).
The issues were first explored in Hegel's The Phenomenology of Mind
(1806) where in the early pages the idea of a dialectic of theories of
knowledge is connected to a dialectic of historical forms of consciousness.
That is, as epistemology (an objective feature) develops, the experience
(a subjective feature) upon which knowledge is based changes. Mach and
Duhem repeat this idea at the end of the 19th century. This century the
view was given its most influential exposition in the writings of Jean Piaget,
indeed it underlays his whole account of cognitive development.
In a much-quoted passage in Genetic Epistemology (1970) Piaget says:

The fundamental hypothesis of genetic epistemology is that there is a parallelism between


the progress made in the logical and rational organisation of knowledge (history of science)
and the corresponding formative psychological processes (p. 13).
H I S T O R Y , P H I L O S O P H Y , AND SCIENCE T E A C H I N G 23

Piaget's latest and most extensive argument for the thesis is found in his
Psychogenesis and the History of Science (1989). However the nature of
the 'parallelism' is somewhat unclear in Piaget, and more so in those he
inspired: positions vary from analogy, to weak isomorphism (Mischel 1971,
p. 326), to strong isomorphism (Murray 1979, p. ix)J ~
Thomas Kuhn popularised the 'cognitive ontogeny recapitulates scien-
tific phylogeny' thesis among historians and philosophers of science (Kuhn
1977, p. 21). Conversely the historian of science, Alexander Koyrr, ob-
served that it was Aristotle's physics that taught him to understand Piaget's
children. The philosopher Philip Kitcher has recently (1988) affirmed that
developmental psychologists can gain insights into the linguistic advances
of young children by studying the shifts that have occurred in the history
of science; and that historians and philosophers of science can learn from
experimental results and analyses of child psychologists. Kitchner (1985)
provides a comprehensive bibliography of philosophical literature on Pia-
get.
Nussbaum (1983) provided an early review of literature in science edu-
cation on individual cognition and cultural cognition, or theory develop-
ment, titled 'Classroom Conceptual Change: the Lesson to be Learned
from the History of Science.' Carey has correctly warned that success in
comprehending the complexity of conceptual change in science students
will 'require the collaboration of cognitive scientists and science educators,
who together must be aware of the understanding of science provided by
both historians and philosophers of science' (1986, p. 1125). An extensive
recent review has been by Duschl, Hamilton, and Grandy (1990).
Piaget's work lead to one obvious area of investigation: do the intuitive,
immediate, 'concrete', conceptions of children mirror the early stages in
the development of scientific understanding in different domains? At one
level of simplification the answer is yes. Children do seem to have pre-
instruction understanding, or naive beliefs, that parallel early scientific,
or pre-scientific notions. This has been much-demonstrated for the field
of mechanics. McCloskey (1983), DiSessa (1982), Clement (1983), Cham-
pagne (1980), Whitaker (1983), McDermott (1984), and Robin and Ohlsson
(1989) are just a few to suggest that naive conceptions of force and motion
mirror the fundamentals of Aristotelian dynamics. Bartov has shown that
intuitive conceptions of biological processes are highly teleological (Bartov
1978). Others point to a Lamarckian-like account of inheritance in children
(Brumby 1979). Mas, Perez and Harris (1987) conducted a study on
adolescent beliefs in chemistry among students who had studied up to five
years of chemistry at school. A significant proportion continued to hold the
Aristotelian-like belief that gases had no mass despite repeated teaching of
the atomic hypothesis of gases.
This last matter, the persistance of intuitive or naive beliefs in the face
of science instruction to the contrary - McCloskey's (1983) demonstration
that 80% of college physics students continued to believe in impetus is
representative of many such studies - has generated productive exchanges
24 MICHAEL R. MATTHEWS

between teachers, psychologists, philosophers, and historians of the type


called for by Carey.

IDEALIZATION IN SCIENCE

Clearly there is something of fundamental importance for the under-


standing of cognition in this resistance of belief to instruction. One philo-
sophical issue is the nature of classical (Newtonian) physics and its relation
to common sense and observation. Too often the epistemological break
from common sense and the everyday world about us involved in Newton-
ian science has been unrecognised, or discounted, in science instruction,
and thus the apparent failure of instruction to instruct has become an
enigma. This under-estimation of the epistemological break from the
everyday world required by classical mechanics is particularly prevalent
among those teachers (the majority) holding empiricist theories of science.
It in some ways parallels the failure of the early empiricist 'champions' of
Newton to understand correctly the true nature of the Scientific Revolu-
tion going on about them: a revolution more dependent upon idealisation,
mathematical analysis, and theoretically informed experiment than it was
upon the patient observations so beloved by empiricist philosophers, and
the writers of introductory chapters in science texts (Mittelstrass 1972).
D u h e m at the turn of the century warned against grounding science
instruction in common sense. He observed:
Now is it clear merely in the light of common sense that a body in the absence of any
force acting on it moves perpetually in a straight line with constant speed? Or that a body
subject to a constant weight constantly accelerates the velocityof its fall? On the contrary
such opinions are remarkably far from common-sense knowledge; in order to give birth
to them, it has taken the accumulated efforts of all the geniuses who for two thousand
years have dealt with dynamics (Duhem 1954. p. 263).

Seventy years later the pedagogical issues raised by this disjunction of


science and common sense are again occupying the attention of science
educators informed by some knowledge of the history and philosophy of
science, Champagne and others expressed the matter in the following
terms: 'the arduousness of learning mechanics is expressed in the effort
required as students shift their thinking from one paradigm to another.
Paradigm shifts are not accomplished easily, neither in the scientific en-
terprise nor in the minds of students' (Champagne et al., 1980, p. 1077).
But even where the paradigm shift is recognised, the effect on students
of learning and accepting a new paradigm so discordant with common
sense and observation can still be underestimated. The 'paradigm shift'
from medieval to classical mechanics, from everyday to scientific under-
standing is not just a matter of getting students to see things differently.
This is still an Aristotelian/empiricist way of stating the matter: seeing is
just not as important as this tradition maintains.
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 25

The enthusiasm of many philosophers, and science educators, for


Kuhn's duck/rabbit problem, Hanson's old/young lady problem, hidden
and ambiguous figures, gestalt switches, and other associated ways of
representing the importance of theory to observation or of the mind to
perception - is understandable, but it has limited application in under-
standing the scientific revolution. Observation simply did not play the
major part that these characterisations suggest. Talk of theory dependence
of observation is interesting, and can serve pedagogical purposes, but it
does not shed much light upon the scientific revolution, or modern science.
It is not seeing things differently, but constructing idealised objects, and
representing and manipulating these mathematically that is the differentia
of the 'new science of a very old subject', to use the words Galileo used
in introducing his Two New Sciences. Galileo did not 'see' balls on inclined
planes as colourless circles on tangents, he saw them just as anyone
else did, but he described them differently and used his mathematical
descriptions in a new theoretical apparatus; similarly Newton did not 'see'
pendulums as point masses on the end of weightless strings, he also saw
them just as everyone else did, but he described them differently and
again used these new descriptions as ingredients in a new theoretical
apparatus. There are observational skills to be developed in science edu-
cation but they do not have the significance often given to them (Norris
1985).
One very simple example illustrates the magnitude of the task. Every
physics student learns by rote the law of the isochrony of the pendulum
- that the period of each swing is the same as any other swing, and that
this is independent of the amplitude and mass of the pendulum. In the
PSSC Physics text the student does not get beyond page four before
he/she learns this law and its supposed discovery by means of Galileo
looking at a swinging chandlier and timing his pulse beat. Yet it is clear
to anyone that any pendulum, of any length, and any mass, if set swinging
will soon come to a halt. This is a conclusive demonstration that pendu-
lums are not isochronic: the period of the last swing is not the same as
the first, a truly isochronic pendulum would ceaselessly swing. How do
we reconcile the law with the observation. The usual way is to say 'forget
what you see, learn the law.' A more sophisticated way is to say 'science
is not really about those kinds of pendulums you are using, it is about
ideal ones' where there is no friction, no air pressure, where the string is
weightless and so on.' This satisfies the HPS informed teacher, but does
it satisfy the student?
Schecker (1988, reproduced in this issue) has addressed some of these
questions in an interesting way. He recognises that the 'main progress in
Galilean/Newtonian physics is the release of thinking from the bounds of
direct, sensual experience . . . . The directly perceptible and measurable
phenomena are imperfect representations of the true 'order' [which is]
only accessible by idealization' (p. 217). He asked 254 high-school students
to comment upon the following statement:
26 M I C H A E L R. M A T T H E W S

In physics lessons there are often assumptions or experiments of thought, which obviously
cannot be realized in actual experiments, like completely excluding air resistance and
other frictional effects or assuming an infinitely lasting linear motion.

The students were asked whether the method was useful or not useful.
11% said it was useless, 'why should I consider something that does not
exist?'; a large group, up to 50%, said it was useful, but only for physics
because physics did not deal with reality, 'I don't need to refer everything
to reality. I am simply interested in physics.'; only about 25% had any
comprehension of the method of idealization in science.
The history of the study of pendulum motion (Matthews 1987, 1990)
can shed much philosophical light upon these pedagogical matters, parti-
cularly the paradoxical finding that students do not believe that physics
deals with reality. Galileo regarded his discovery of isochrony as pivotal
to his whole new physics. Yet his 'discoveries' were vigorously denied by
Guilobaldo del Monte who was Galileo's own patron, and a person de-
scribed by Stillman Drake as the 'greatest mechanician of the 16th cen-
tury.' The argument between them encapsulates the epistemological divide
between the old Aristotelian, empirical science and the new idealistic,
mathematical, experimentalist science characteristic of the Scientific Revo-
lution. Del Monte insisted to Galileo that the pendulums he tested were
not isochronic: cork and brass ones did not have the same period, long
and short ones did not have the same period, all pendulums came to a
halt within two or three dozen oscillations. Galileo replied that these
results applied only to actual pendulums, and that if ideal ones were
studied (where friction, air resistance, and the weight of the string were
eliminated), then they would indeed be found to be isochronic. Galileo
arrived at his law of isochrony by mathematical (specifically geometrical)
calculation. Del Monte asserted that the mathematics was fine, but it was
not physics; physics was to be about the real world, not an ideal one. Not just
del Monte, but Huygens and a host of others disbelieved Galileo's claims
about the isochronic motion of the pendulum; indeed Huygens said that
Galileo must have invented the experiments rather than actually per-
formed them. It is little wonder that when teachers rely on a simple,
unguided discovery method, students will come up with del Monte's result,
not Galileo's. A little history of science can prepare teachers for this
result. A little philosophy of science can assist teachers to interpret the
results to students.
There is a difference between the real objects of the world, and the
theoretical objects of science. To confuse the first with the second is to
confuse Aristotelian with Newtonian science. Di Sessa remarked on the
failure of standard discovery learning that 'it seems that very few subjects,
if any, had learned much characteristically Newtonian from dealing with
the everyday w o r l d . . , thought experiments might be more useful than
"playing around"' (1982, p. 62). To expect students to learn anything
Newtonian by playing around with objects is to underestimate the epis-
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 27

temological revolution inaugurated by Galileo and Newton; and also to


underestimate the pedagogical problems in comprehending the classical
scientific worldview. Playing around with, or prolonged looking at, real
material objects, will not generate the point masses, inertial bodies, force
definitions, geometry and calculus that are all essential parts of the theoret-
ical objects of mechanics. The former enter into science when they are
depicted in terms of the latter. Once so dressed, and in the system, they
can be operated on by the conceptual apparatus of science.
Some related issues are discussed in Steinberg, Brown, and Clement's
detailed historical/psychological paper (1990) on 'The Conceptual Diffi-
culties Impeding Isaac Newton and Contemporary Physics Students'. The
relatively neglected topic of thought experiments in science education is
discussed in Helm, Gilbert and Watts (1985), Stinner (1990), Winchester
(1990), and Matthews (1989a).
The philosophical issues associated with idealisation and abstraction in
science has of course implications for the massive misconceptions (Helm
1980) or alternative conceptions (Driver and Easley 1978) or naive concep-
tions literature in science education. To go back to del Monte: he had no
misconceptions about pendulum motion, he only had misconceptions
about the new science's view of pendulum motion. It is commonly
said that Aristotelians had misconceptions about the real world;
it would be more accurate to say that they had misconceptions
about the contrived and constructed world of Galileo and Newton, that
is the theoretical objects of the new science, not the material objects
around them. The philosophical issue also has implications for the Piage-
tian cognitive conflict account of conceptual change: clearly experience is
not going to provide the cognitive conflict that is supposedly the motor of
conceptual change. To repeat what has been said: experience is very
Aristotelian. So much so that one prominent historian of the Scientific
Revolution remarked that 'observation and experience.., had a very
small part in the edification of modern science; one could even say that
they constituted the chief obstacles that it encountered on its way' (Koyr6
1968, p. 90). The recognition of scientific idealisation gives rise to many
questions in this area: what is a misconception? is the real world the
touchstone against which we judge our conceptions or is the touchstone
just another conceptualisation of the world? The issue of idealization is
belatedly being attended to in the HPS and Science Teaching literature.
Floden et al. (1987) have discussed pedagogical problems occasioned by
the rupture of science education from everyday experience. Garrison and
Bentley (1990) develop the theme in a discussion of the influential Posner
et al. (1982) theory of conceptual change. Ginev provides a very sophisti-
cated account of science curricula built upon the recognition that 'The
very process of idealization is considered as the epistemological differentia
specifica of science' (1990, p. 65).
History and philosophy can make the idealizations of science more
human, understandable, and explain them as artifacts in their own right
28 M I C H A E L R. M A T T H E W S

to be appreciated. This is important for students being introduced to the


'world of science'. A failure to appreciate just what idealization is, and
what it is not, has been the basis of a lot of anti-scientific criticism. It was
of course Newtonian idealization that the Romantic reaction was directed
against. For them (Keats, Goethe, etc.) the rich world of lived experience
was not captured by the colourless point masses of Newton. In the 20th
century Marcuse, Husserl, Tillich and others have repeated versions of this
charge. The prominent German science-educator, Martin Wagenschein
(1962), has written on this dichotomy at the heart of an education in
science. The point which has to be stressed here is that the theoretical
objects of science are not meant to be accounts of the material objects in
the world, at least not in the sense of pictures of them.
A historically and philosophically literate science teacher can assist
students to grasp just how science captures, and does not capture, the
real, subjective, lived world. More commonly the student is left with
the unhappy choice between disowning their own world as a fantasy, or
disowning the world of science as a fantasy. Once more following Mach,
the lived phenomenal world is vital to science instruction, that is where
curiosity and wonder begin, but it is not to be confused with an inertial
world, or a world of ideal gases. Some of these matters are discussed in
Eger (1972) and Passmore (1978).

PHILOSOPHY OF SCIENCE AND SCIENCE TEACHING

Ten years ago Ennis opened a comprehensive review of the US literature


on philosophy of science and science teaching (the only article on science
education to appear in a philosophy of science journal in decades), with
the melancholic observation that: 'With some exceptions philosophers of
science have not shown much explicit interest in the problems of science
education' (Ennis 1979, p. 138). He neglected England and the significant
contribution of Whitehead - for this see Birch (1988). If Ennis could then
count the number of such philosophers on one hand - Dewey, Scheffler,
Martin, Margenau, Nagel - they can now, encouragingly, be counted on
at least two hands. Some names that could be added are Siegel, Harre,
Buchdahl, Ruse, and Pitt. Among science educators, Richard Duschl in
the US, Derek Hodson in the British community, and Walter Jung in the
European community, have recently advocated greater cooperation with
philosophers of science.
The philosopher who contributed most to opening the dialogue between
HPS and science education was Michael Martin in his popular book Con-
cepts of Science Education (1972). Both Ennis and Martin were analytic
philosophers. In five chapters dealing with Inquiry, Explanation, Defi-
nition, Observation, and Goals, Martin provides ample evidence of the
usefulness of philosophy for the improvement of instruction, texts, and
statements of aims and objectives of science courses.
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 29

Two examples give a sense of Martin's contribution. He notes that


explanation is central to science instruction, and that the concept recurs
repeatedly in texts, specifically mentioning the BSSC Biological Science:
An Inquiry into Life, and the ESCP Investigating the Earth, yet in neither
text is any attention paid to discussing what a scientific explanation, and how
good ones might be differentiated from bad ones. He notes that some
elementary elaboration of the covering law model, and the statistical-
probabilistic models, would do no harm, and most probably do consider-
able good. In the last chapter he reproduces a statement of objectives
for science education from the then important US Educational Policies
Commission's Education and the Spirit of Science (1966). Under its third
goal of science teaching it says:
The longingto knowis the motivationfor learning; data and generalizationare the forms
whichknowledgetakes. Generalizations are induced from discrete bits of'information
gathered throughobservationconductedas accuratelyas the circumstancespermit.

Martin notes, in contradiction to this, that some scientific theories are not
generalizations, and that most scientific hypotheses are not generated by
induction, and that further, observation requires theory.
This last matter, the lack of awareness of the theory dependence of
observation and its implications, well illustrates the lamentable separation
of professional philosophers of science from science educators, and from
government advisory bodies. The Spirit of Science report was published
in 1966. Inductivism was at that time in full retreat in philosophy of
science: Hanson's Patterns of Scientific Discovery had appeared in 1958,
Popper's Logic of Scientific Discovery had been translated in 1959, Toul-
min's Foresight and Understanding appeared in 1961, Kuhn's Structure of
Scientific Revolutions appeared in 1962. There was ample material close
at hand, to say nothing of material across the Atlantic (Bachelard), and
material further back in time (Collingwood, Fleck) that could have infor-
med the deliberations of such a high ranking US government agency
issuing a major report on science education. Instead the material was
neglected, and dubious slogans were promulgated without any hint that
they were debatable, much less that they were likely to be false.
It was not just Government reports that suffered from an ignorance of
developments in philosophy of science. Robert Gagne, one of the most
influential learning theorists of 1960-80 and a prominent figure behind
inquiry-learning curricula in science, in 1963 lent his considerable prestige
to a debatable view of scientific inquiry as inquiry where:
a set of activities whichbegins with a careful set of systematicobservations, proceeds to
design the measurements required, deafly distinguishesbetween what is observed and
what is inferred.., and drawsreasonableconclusions.(Gagne 1963, from Hodson,forth-
coming).

Numerous NSF curricula projects of the early 1960's, and the Nuffield
scheme of the 1960's in England suffered similarly from ignorance of
30 M I C H A E L R. M A T T H E W S

developments in philosophy of science, promulgating a discovery approach


to science replicating the supposed inductive methods of science. Bruner's
slogan - ~produce scientists by having students be scientists' - was admir-
able if teachers and curricular framers had a reasonable idea of what it
was to be a scientist. Most accepted the inductivist mythology of the
textbooks. Stevens (1978), Forge (1979), and Duschl (1985) have written
on the philosophical assumptions of these NSF and Nuffield curricula.
If policy documents, leading educational theorists, curricula, and
textbooks embody and uncritically promulgate certain philosophical posi-
tions, it is not surprising that classroom science teachers do likewise. A
teacher's theory about the nature of science, his or her epistemology, can
be conveyed explicitly or implicitly. This epistemology effects the class-
room behaviour of teachers (Robinson 1969). How this epistemology is
formed, what effects it has on teacher practice, and how it contributes to
students' images of science, has been the subject of many recent studies:
Abell (1989), Rowell & Cawthron (1982), Jacoby & Spargo (1989), Leder-
man & Zeidler (1987), and Koulaidis & Ogborn (1989). This research
becomes the more important as schemes such as the British National
Curricula, the 2061 Project and others come into effect in which topics
such as 'The Nature of Science' are part of the curricula.
A teacher's epistemology is informally formed; it is picked up in the
process of the textbook education described by Kuhn. It consists of current
prejudices which have been little bothered by historical information, or
by philosophical analysis. In only two of 55 institutions providing science
teacher training in Australia is a course in HPS compulsory. Of the fifteen
leading centres of science teacher training in the US only half require a
course in philosophy of science (Loving 1992); the proportion in the
remaining hundreds of centres is likely to be far less. The situation in the
UK is no more encouraging.

HPS AND TEACHER EDUCATION

Many have argued that HPS should be part of the education of science
teachers - the British Thomson Report in 1918 had said 'some knowledge
of the history and philosophy of science should form part of the intellectual
equipment of every science teacher in a secondary school' (p. 3). One
argument for HPS is that it produces better (more coherent, stimulating,
critical, humane, etc.) teaching. This utilitarian argument is not the only
one: a case can be made for teachers having critical knowledge (meaning
historical and philosophical knowledge) of their subject matter quite inde-
pendently of whether this knowledge is directly used in pedagogy - there
is more to a teacher than meets the classroom.
Michael Polanyi made the obvious point that HPS should be as much
a part of science education as literary and musical criticism is part of
literary and musical education (Harre 1983, p. 141). It would be odd to
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 31

think of a good literature teacher who has no knowledge of the elements


of literary criticism: of the tradition of debate over what identifies good
literature, how literature is related to social interests, the history of literary
forms etc. So also it should be equally odd to think of a good science
teacher who has no reasonably sophisticated knowledge of the terms of
their own discipline - 'cause', 'law', 'explanation', 'model', 'theory', 'fact';
no knowledge of the often conflicting objectives of their own discipline -
to describe, to control, to understand; or no knowledge of the cultural
and historical dimension of their own discipline. Israel Scheffler in a
largely neglected paper of 1970 argued just this point. This is part of the
difference between being educated in science and being simply trained in
science: teachers ought to be educated. HPS clearly contributes to this
richer understanding of science.
This is a point of connection to the voluminous literature on scientific
literacy: if to be literate is to have some depth of understanding of the
words and concepts of a discourse, then the history and philosophy of
science clearly contributes to a deeper and more critical scientific literacy
(Miller 1983).
To advocate the importance of the history and philosophy of science
for science teachers is not novel. The opening pages of a 1929 text for
science teachers describes a successful science teacher as one who:
knows his own s u b j e c t . . , is widely read in other branches of science.., knows how to
t e a c h . . , is able to express himself lucidly.., is skilful in manipulation.., is resourceful
both at the demonstration table and in the l a b o r a t o r y . . , is a logician.., is something of
a philosopher.., is so far an historian that he can sit down with a crowd of boys and talk
to them about the personal equations, the lives, and the work of such geniuses as Galileo,
Newton, Faraday and Darwin (quoted in Sherratt 1983, p. 418).

This ideal has current relevance. As has been mentioned, the new
curricula being developed and implemented in Britain, the US, Denmark,
and Canada will require just such qualities in a teacher if they are to be
successfully taught. Episodes in the history of science, and questions about
the nature (philosophy) of science are part of these curricula. Furthermore
in Britain, the US, Australia and elsewhere there are efforts to identify
outstanding science teachers and to assess teachers. This requires delinea-
ting the qualities of a good science teacher; and increasingly some com-
petence or familiarity with HPS topics is being required.
In the US, the Stanford-based, Carnegie-funded, National Teacher As-
sessment Project, directed by Lee Shulman, is the foremost teacher as-
sessment programme. It is intellectualist in its criteria of teacher com-
petence and rejects the behaviourist, managerial, measures of teacher
competence so long enshrined in evaluation practice. Shulman asks about
the 'missing paradigm' - the command of subject matter - and the ability
to make it intelligible to students, abilities requiring the wider view pro-
vided by HPS. In one of his influential publications, Shulman has said:
To think properly about content knowledge requires going beyond knowledge of the
32 MICHAEL R. MATTHEWS

facts or concepts of a domain. It requires understanding the structures of the subject


matter... Teachers must not only be capable of defining for students the accepted truths
in a domain. They must also be able to explain why a particular proposition is deemed
warranted, why it is worth knowing, and how it relates to other propositions, both within
the discipline and without, both in theory and in practice. (Shulman 1986, p. 9).

To explain why a particular proposition is deemed warranted - the law


of inertia, the principle of conservation of energy, the theory of evolution,
continental drift theory, accounts of atomic structure etc. - requires know-
ing something about how evidence relates to theory appraisal, this is the
standard business of epistemology. Shulman's ideas are reflected in the
National Board for Professional Teaching Standards assessment guidelines
- What Teachers Should K n o w and Be A b l e to Do (1989).
An evaluation package for biology teachers that has been developed as
part of the Carnegie project tries to assess teachers grasp of the nature of
science, its processes, and determinants. In their words 'Do teachers hold
a rich conception of the scientific enterprise as an interaction of the facts,
laws and theories of a domain, mastering the skills to construct such
knowledge, and recognising that this knowledge is influenced by and has
influence on human society' (Collins 1989, p. 64).
As the history and philosophy of science becomes a more recognised
component of teacher education it is timely to ask what kind of HPS
courses are appropriate. Recent literature contains accounts of a number
of such courses, and reflections upon their adequacy. A consensus is that
for such courses to be seen as relevant to a prospective teacher they
should be applied, or practical courses. Sending education students to a
Philosophy Department to do HPS is not the most satisfactory way of
proceeding. To use the words of Bevilacqua, who has been promoting
HPS among teachers in Italy, such courses become yet another 'brick in
the wall', or another chore to be completed before getting on to teaching.
HPS courses should begin with problems that teachers can see as pertinent
to their teaching or professional development. Such courses are described
in Johnson and Stewart (1990), Eger (1987), Bybee (1990), Bakker and
Clark (1989), and Ruse (1990).
My own course (Matthews 1990b) which has run with some success for
a number of years, is based upon selections of the writings of Galileo,
Boyle, Newton, Huygens, Darwin and others. I have found, not surpris-
ingly, that teachers appreciate the opportunity to read something of their
work. From hundreds of biology graduates I have found only a handful
who have read any writing of Darwin; from hundreds of physics graduates
I have encountered none who have read anything from Galileo or Newton.
As one teacher stated 'teachers are hungry for this knowledge'. The
philosophical issues - realism, instrumentalism, authority, reductionism,
causality, explanation, idealization etc. are dealt with and developed as
they arise out of the text. Most of the texts used have been published in
Matthews (1989b).
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 33

SOME TOPICAL QUESTIONS

Ennis in 1979 lists six areas of concern to science teachers that would
benefit by philosophical attention. These are: scientific method, criteria
for critical thinking about empirical statements, the structure of scientific
disciplines, explanation, value judgements by scientists, and test develop-
ment. Ten years later there is point in producing another such list. The
topics I would propose are: Feminism, Constructivism, Ethics, Metaphys-
ics, Idealization, and Rationality. In one form or another these issues and
their implications have surfaced in science education debate. This is not
to say that more prosaic matters ought not to be discussed - research in
Australia showed that about one third of the first year chemistry class in
one university did not know that 'affirming the consequent' was invalid
reasoning: they believed that if A implies B, and B is the case, then A is
then true. This suggests that simple logic and reasoning skills are not out
of place, even if more complex matters cannot be dealt with.
(1) Feminism has provided strong challenges to the assumptions of both
science teaching and of the philosophy of science. It is notorious that
women do not continue on in science studies. There is a mountain of
literature on the subject - see the 1987 special issue of The International
Journal of Science Education. Much of it is empirical, dealing with impedi-
ments to women's progress and interest in science. In addition there is a
philosophical claim demanding a response. Bleier (1984), Harding (1986),
Keller (1985), and Martin (1989) have all argued the case for a masculine
bias in the very epistemology of western science. Science educators have
begun to pay attention to this critique. Martin has said that 'To bring the
gender-based critique into science education is to incorporate teaching
about science into the teaching of science' (Martin, 1989, p. 251). It is
known that ideologies of class, race, and religion have effected the conduct
o f s c i e n c e - Lysencko's genetics, Nazi haematologists finding a Jewish
blood type, and aspects of medieval science provide examples of each. It
is a priori thus possible that sexist ideology can affect science, including
its epistemology. Harding provides a vision of a new kind of knowledge
promoted by feminism, it:

seeks a unity of knowledge combining moral and political with empirical understanding.
And it seeks to unify knowledge of and by the heart with that which is gained by and
about the brains and hand. It sees inquiry as comprising not just the mechanical observation
of nature and others but the intervention of political and moral illumination 'without which
the secrets of nature cannot be uncovered'. (Harding 1986, p. 241)

This is an epistemological claim and needs to be defended or criticised


accordingly, this requires attention to the history and philosophy of
science.
(2) Constructivism is arguably the dominant epistemology among sci-
ence educators. It is articulated in the works of von Glasersfeld (1989),
Novak (1987), Driver (1985) and many others. One recent commentary
34 MICHAEL R. MATTHEWS

has identified the following schools or variants of constructivism: contex-


tual, dialectical, empirical, information-processing, methodological, mod-
erate, Piagetian, postepistemological, pragmatic, radical, realist, social
and socio-historical. To which can be added, humanistic (Cheung & Taylor
1991), Constructivism is in the Piagetian, cognitive, learning theory tra-
dition, it is opposed to behaviourism of a Skinnerian or Gagneian type,
so long dominant in science education. As with most of this tradition, its
theory of mind is fundamentally Kantian. Constructivism's ontology varies
from radical idealism (particularly in some of von Glasersfeld's writings),
through to Popperian three-world theory. Its pedagogical practice is anti-
didactic, and student-centred with an emphasis on student engagement in
problem identification, hypothesis development, testing, and argument.
There are a host of philosophical issues in constructivist theory that
deserve elaboration: What account of the social dimension of knowledge
is given? What are the criteria for the adequacy of student conceptions:
are they judged against norms of the scientific community, or against other
student accounts, or against the individuals' prior conceptions? Is there a
confusion of successful pedagogical practice with epistemological claims?
Driver in one publication recognises an essential tension in constructivist
practice: that between getting students to construct and make meaningful
their own accounts of something and getting them to participate in a
scientific community that has its own theory and understandings (Driver
& Oldham 1986). The former makes no epistemological judgement, it only
refers to psychological mechanisms; the latter does make epistemological
judgements. Teachers who are concerned with students entering and mas-
tering this sphere of public knowledge will standardly need to indicate
that explanations that students might devise and feel content with are in
fact inadequate. Strike (1987) and Gruender (1989) have raised some of
these questions; Suchting (1992) provides a penetrating critique.
One deep-seated problem with much constructivist writing is that it
maintains a fundamentally empiricist conception of knowledge, despite
their stated antipathy to such views. This is suggested when constructivists
say that we are only aware of our sense impressions, we cannot have
access to the world as it is, knowledge is the correspondence of mental
image or ideas to reality, it is the individual subject who is the locus of
knowledge claims. All of these are standard empiricist claims. But scien-
tific knowledge requires idealisations and these cannot by definition mirror
or correspond to the world. So either correspondence has to be dropped
as a criterion of knowledge, or no branch of modern science can be
accepted as knowledge. The former seems more sensible. But if this
is done then a large part of the constructivist argument for relativism
disappears.
(3) Ethical questions increasingly arise in almost all areas of the science
curriculum. The greenhouse effect, pollution, extinction of species, genetic
engineering, military technology and the employment of scientists in the
defense industries, the cost and direction of scientific research, nuclear
H I S T O R Y , P H I L O S O P H Y , AND SCIENCE T E A C H I N G 35

energy and nuclear war, and so on - are all matters that are raised by
students, and appear in new science curricula. The once straight-forward
and unreflective use of animals for science experiments and laboratory
dissections is now questioned (see Cordero, this issue) and in many places
severely limited. An aim of the New Zealand science syllabus is 'the care
of animals' and recognition of their rights. At the same time in philosophy
these questions are being dealt with in applied ethics, and environmental
ethics courses. Hitherto, partly under the influence of belief in value-free
science, these questions have been ignored in science education. They can
no longer be ignored.
John Ziman and numerous others have correctly pointed out that ortho-
dox science education has for long promulgated naive materialism, primi-
tive positivism, and complacent technocracy (Ziman 1980). These posi-
tions (ideologies) are now in intellectual retreat. The pedagogical
consequences of the 'Death of Value-free Science' must now be faced.
The PLON project in The Netherlands, the SISCON project in the UK,
various Canadian projects (Aikenhead 1980), STS courses and the Project
2061 proposals in the US are the most obvious curricular responses. A
considerable literature on ethics in science education has been generated
(Mendelsohn 1976, Gosling and Musschenga 1985). Teachers can benefit
from considering some of the arguments advanced by philosophers who
have considered similar issues.
The recent exchange between Eger, Hesse, Shimony, and others (Zygon
23 (3), 1988) on 'rationality in science and ethics' (reproduced in Matthews
1991) shows what such collaboration can achieve. Eger (1989) has also
addressed the question of the 'interests' of science, taking up issues that
the work of Habermas and the Frankfurt School pose for understanding
the social role of science and the fundamental structures of the discipline.
(4) Metaphysical issues naturally emerge from the subject matter of
science: Einstein spoke of the scientist being a philosopher in working-
man's clothes. Ethical issues in science raise questions about our respons-
ibility to nature which in turn is prompted by a new conception of nature
itself. The mechanical, Laplacian, world-view of Newtonian science is
being challenged by a new ontology of nature. Gotschl (1990) says that
'The extensive call for responsibility results in an ethical and anthropologi-
cal revolution beside the scientific and technical revolution.' Teachers
ought to have some grasp of the issues: there are a host of competing
ontologies in the market-place waiting to take the place of the mechanical
world-view, some make considerably more sense than others. An HPS
informed science teacher can contribute to the evaluation of them.
Historical studies in science vividly portray the independence of science
and metaphysics. The Galilean/Aristotelian controversy over final caus-
ation, the Galilean/Kepler controversy over the lunar theory of tides, the
Newtonian/Cartesian argument over action at a distance, the Newtonian/
Berkelian argument over the existence of absolute space and time, the
Newtonian/Fresnal argument over the particulate theory of light, the
36 MICHAEL R. M A T T H E W S

Darwinian/Paley argument over design and natural selection, the Mach/


Bohr argument over atomic theory, the Einstein/Copenhagen dispute over
the deterministic interpretation of quantum theory etc. - all bring to the
fore metaphysical issues. Metaphysics is pervasive in science. Peirce had
said in his 'Notes on Scientific Philosophy' that 'Find a scientific man who
proposes to get along without any metaphysics.., and you have found
one whose doctrines are thoroughly vitiated by the crude and uncritized
metaphysics with which they are packed.' These interconnections are dis-
cussed in Wartofsky (1979), Gjertsen (1989), and Matthews (1989b).
Woolnough (1989) has discussed an important topic in the history of
science that is usually ignored in the science syllabus: the role of religious
belief in the motivation and conceptualizations of great scientists. Students
learn often enough that Newton discovered three laws and the formulae
for them; they learn less often that Newton said when he wrote his
Principia, the foundation of all modern science, 'I had an eye upon such
principles as might work with considering men for the belief of a Deity;
and nothing can rejoice me more than to find it useful for that purpose'
(Thayer 1953, p. 46). They also learn often enough that Boyle discovered
an important law and what its formulae is; they learn less often that he
left a provision in his will for a set of public lectures 'for proving the
Christian religion against notorious infidels' and that he believed his own
mechanical philosophy admirably suited for proving the existence of a
Designer of the universe. Despite the fact that the history of western
science is largely a history of the endeavours of people who understood
themselves as doing work which proclaimed the majesty of God, nothing
much is heard of this in the typical science classroom treatment of these
figures and their discoveries. There are engaging psychological, cultural,
and philosophical stories that can profitably be told.
If science has developed as a dialogue with metaphysics (to say nothing
of interjections from the political, economic and social realms), then to
teach science as a soliloquy in which science just talks to itself and grows
entirely by self-criticism is to impoverish the subject matter.
Substantive content in science will also give rise to metaphysical ques-
tions. The biologist Charles Birch states this at its most general level:
'Any teacher of science eventually conveys his or her answer to the ques-
tion - what is nature made o f ? . . . The dominant answer today is given
in terms of the mechanistic model of nature' (Birch 1988, p. 33). Birch
argues against the mechanistic, Cartesian world-view in favour of a White-
headian process metaphysics. Following Whitehead he says that 'the role
of education is to deal with facts and ideas in such a way that we are left
with a generalised understanding': metaphysics will figure in this under-
standing. Thomas Settle tackles, in an admirably non-technical way, the
question addressed by Birch: 'Can Physicalism be Avoided in Teaching
School Science?' (1990). It is noteworthy that these and other papers
(Stinner 1990) are resurrecting the philosophical and pedagogical views of
H I S T O R Y , P H I L O S O P H Y , AND SCIENCE T E A C H I N G 37

Alfred North Whitehead first published sixty years ago in his The Aims
of Education (1923).
The considerable literature generated by the Philosophy for Children
Movement suggests that children are both capable of, and interested
in, pursuing elementary philosophical questions (Lipman & Sharp 1978,
Dawson-Galle 1990). The science classroom provides ample opportunities
to do this.
(5) Idealization is the sine qua non of modern mathematical science,
yet it is very little understood by teachers, it rarely occurs in textbook
accounts of the scientific method, and is often ignored by philosophers
who conduct discussion of induction, falsification, and the testing of theory
completely oblivious to the fact that it is idealized laws and theories that
are being discussed, and simple logic is inappropriate to their evaluation.
Also much science education literature on concept acquisition proceeds
in an Aristotelian manner, in which idealizations are treated as empirical
generalisations. Clearly the acquisitions of the concept of point mass,
frictionless surface, inertial frame, elastic collision, rigid body, etc. does
not occur in the Aristotelian manner, they do not arise from looking at
bodies and inducing common features. Galilean/Newtonian idealization
was a monumental conceptual achievement, arguably something that sepa-
rates human thought from all animal cognition. This achievement must
be imparted to students - they will not acquire the idealisations by looking
at nature. Polish philosophers of science have dealt extensively with
the logic and philosophy of idealization in the sciences (Nowak 1980,
Krajewski 1982).
Duhem's felicitious expression 'the logic of a subject is not necessarily
the logic of its presentation' needs to be remembered. History alerts
teachers to the need for a phenomenal, approach to idealisations - stud-
ents need to know what the idealisations relate to.
(6) Rationality is a topic that conjoins HPS and science education. In
HPS the rationality of scientific theory change, that is of the history of
science, has been a hotly contested issue. Rationalists were awakened
from their slumbers by the blast of Kuhn's 1962 Structure in which he said
scientific transformations depend not so much upon rational persuasion as
they do upon mob psychology and the mortality of the aged. Just as
philosophers were coming to terms with this, and Feyerabend's extension
of the thesis (see the defenses of rationalism in Shapere 1984 and Siegel
1987), the Edinburgh school of sociologists of science - David Bloor,
Barry Barnes, Steven Shapin, and Michael Mulkay - further criticised
rationalism in their strong programme which tries to account for all scien-
tific change in externalist terms (for some of the key expositions and
criticisms see Brown 1984). These are major claims that impinge on the
very rationale of science education and deserve the attention of teachers.
Among educators the topic of critical thinking - what is it? how can it
be promoted? is it transferable across disciplines? - has likewise been
38 M I C H A E L R. M A T T H E W S

a hotly contested area. It is clear that a discussion of critical thinking


independently of scientific thinking is greatly truncated. Siegel (1989)
defends rationality and reason giving as the hall-marks of science edu-
cation against dogmatists on the one hand, and Feyerabendian ir-
rationalists on the other. Eger (1990) takes up the question of how such
a conception allows for the role of commitment, or faith, that has been
so important to the development of science.

CONCEUSION

I have given an account of what I see as the emerging confluence of


themes in science, philosophy, history, and science education that paint a
richer, and more variegated, picture of science than has been generally
portrayed in school texts and classrooms. New curricula are attempting to
bring something of this richer picture into the classroom. The success of
this will depend upon first introducing appropriate, science-teacher related
courses in the history and philosophy of science into pre-service and in-
service teacher education. Science is one of the great achievements of
human culture. Science teaching, to use the words of a 1918 British
Association for the Advancement of Science report, should convey 'more
of the spirit and less of the valley of dry bones' of this achievement. If
this is done, then some beginnings can be made to overcoming the present
intellectual and social crisis of science education.

NOTES

1. This paper is an augmented version of a paper which originally appeared in Studies in


Science Education 18, 1990. It is offered as one view of the present state of studies on
the relationship of history, philosophy and science teaching. This view is of course
partial; others will stress different matters. Ranging as it does over issues in the history
of science, the philosophy of science, the history of science education, and current
developments in science education - it is likely that justice has not been done to any
particular issue. Recognising this, there is still some point at the outset of this new
journal in trying to give some sense of the literature and questions that are occupying
each of the diverse science communities the journal is serving.
It is hoped to soon publish a comparable paper dealing with mathematics education.
The research was made possible because of assistance from the University of New
South Wales. It has been enriched because of the cooperation of numerous scholars
around the world whose manuscripts I have read for special issues of journals I have
edited in recent years, and whose advice and conversation I have benefited from.
2. The British National Curriculum is documented in NCC (1988). It is discussed in Ak-
eroyd (1989), Solomon (1990), and Ray (1990). The Danish curriculum in 'The History
and Technology of Science' is discussed in Nielsen & Nielsen (1988), and Nielsen and
Thomsen (1990). In the Netherlands there has been a 'Physics in Society' course since
1981, see Eijkelhof & Swager (1983); and since 1972 various materials generated by the
PLON project (Project Curriculum Development in Physics, PO Box 80.008, 3508 TA
Utrecht) have incorporated a HPS dimension. The Project 2061 proposals are contained
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 39

in AAAS (1989), and are discussed in Stein (1989). A discussion of STS programmes
and a guide to the literature can be found in McFadden (1989).
3. The conference was based upon six journals containing 55 papers: Educational Philos-
ophy and Theory 20(2), 1988; Interchange 20(2), 1989; Synthese 80(1), 1989; Studies in
Philosophy and Education 10(1), 1990; International Journal of Science Education 12(3),
1990; and Science Education 75(1), 1991. A two volume set of Proceedings was published
(Herget 1989, 1990). A selection of the papers, along with some others, have been
published as a book (Matthews 1991).
4. For Proceedings of the conferences see Bevilacqua & Kennedy (1983), Thomsen (1986),
and Blondel & Brouzeng (1988).
5. Ernst Mach's 19th century writings on science education are as wide-ranging and stimulat-
ing as they are ignored. An introduction to his views can be found in Matthews 1989a.
6. Albert Einstein provides a dramatic confirmation of Mach's views on the utility of
historical investigation. In his autobiographical essay he comments how in the late
nineteenth-century physicists never tired of trying to base Maxwell's theory of electro-
magnetism upon mechanical principles. He says that 'it was Ernst Mach who, in his
History of Mechanics, shook this dogmatic faith; this book exercised a profound influence
upon me in this regard while I was a student' (Schlipp 1951, p. 21). This was the catalyst
that allowed Einstein to 'enter upon a critique of mechanics as the foundation of physics'.
7. Schwab was long associated with the University of Chicago and was imbued with
its 'great books' tradition. He had independently of Kuhn and contemporaneously with
him, enunciated a distinction between 'fluid' and 'stable' periods of scientific inquiry
which parallels Kuhn's better known distinction of 'revolutionary' and 'normal' science.
He sustained a deep involvement with the theory and practice of education comparable
to that of another Chicagoan philosopher-educator, John Dewey. A selection of his
articles is contained in Ford and Pugno (1964). A list of his publications occurs in (Dublin
1989).
8. Brush can be contacted at The Institute for Physical Science and Technology, University
of Maryland, College Park, MD 20747, USA. See his 1988a,b for classroom materials,
and his 1989 for a review of the field.
9. A review of the problems and literature in historiography of science can be found in
Kragh (1987). On connections between the philosophy of science and the history of
science see, to begin with, Giere (1973), McMullin (1975), and Wartofsky (1976).
10. Just whether the recapitulation law is meant to apply to biological or to conceptual
matters has been debated. In places Piaget denies a biological recapitulation, saying 'Let
us guard against returning to the simplistic idea of a necessary parallelism between the
development of the race and that of the individual, a parallelism which biologists have
shown to be equivocal and conjectural (in Kitchener 1986, 6). He is more committed to
a conceptual form of parallelism, and maintains that similar mechanisms are involved
in the transformation of scientific theory as there are in individual conceptual change:
decentration, assimilation-accommodation, equilibration, constructivism etc. Some of
this literature is reviewed in Siegel (1982).

REFERENCES

Aikenhead, G. S.: 1980, Science in Social Issues: Implications for Teaching, Science Council
of Canada, Ottawa.
Akeroyd, F. M.: 1989, 'Philosophy of Science in a National Curriculum', in D. E. Herget
(ed.) The History and Philosophy of Science in Science Teaching, Florida State University,
pp. 15-22.
American Association for the Advancement of Science: 1989, Science for All Americans,
AAAS, Washington.
40 M I C H A E L R. M A T T H E W S

Arons, A. B.: 1988, 'Historical and Philosophical Perspectives Attainable in Introductory


Physics Courses', Educational Philosophy and Theory 20(2), 13-23. Reprinted in M. R.
Matthews (ed.) History, Philosophy and Science Teaching: Selected Readings, OISE Press,
Toronto and Teachers College Press, New York, 1991.
Association for Science Education: 1963, The Training of Graduate Science Teachers, ASE,
Hatfield, Herts.
Association for Science Education: 1979, Alternatives for Science Education. A Consultative
Document, ASE, Hatfield, Herts.
Association for Science Education: 1980, What is Science?, ASE, Hatfield, Herts.
Association for Science Education: 1981, Education through Science, ASE, Hatfield, Herts.
Baker, G. R. & Clark, L.: 1989, 'The Concept of Explanation: Teaching the Philosophy of
Science to Science Majors', in D. E. Herget (ed.) The History and Philosophy of Science
in Science Teaching, Florida State University, pp. 23-29.
Bevilacqua, F.: 1990, 'Can the History of Physics Improve Physics Teaching?', in D. E.
Herget (ed.) The History and Philosophy of Science in Science Teaching, vol II, Florida
State University.
Bevilacqua, F. and Kennedy, P. J. (eds.): 1983, Proceedings of the International Conference
on Using History of Physics in lnnovatory Physics Education, University of Pavia, Pavia.
Birch, C.: 1988, 'Whitehead and Science Education', Educational Philosophy and Theory
20(2), 33-41.
Bleier, R.: 1984, Science and Gender, Pergamon Press, New York.
Brock, W. H.: 1987, 'History of Science in British Schools: Past, Present & Future', in M.
Shortland & A. Warwick (eds.) Teaching the History of Science, Basil Blackwell, Oxford,
pp. 30-41.
Brown, J. R. (ed.): 1984, Scientific Rationality: the Sociological Turn, Reidel, Dordrecht.
Brumby, M.: 1979, 'Problems in Learning the Concept of Natural Selection', Journal of
Biological Education 13, 119-122.
Brush, S. G.: 1974, 'Should the History of Science be Rated X?' Science 18, 1164-1172.
Brush, S. G.: 1988a, The History of Modern Science: A Guide to the Second Scientific
Revolution, 1800--1950, Iowa State University Press, Ames 10.
Brush, S. G. (ed.): 1988b, History of Physics: Selected Reprints, American Association of
Physics Teachers, College Park MD.
Brush, S. G.: 1989, 'History of Science & Science Education' Interchange 20(2), 60-70.
Brush, S. G. & King, A. L. Y. (eds.): 1972, History in the Teaching of Physics, University
Press of New England, Hanover, NH.
Buchdahl, G.: 1983, 'Styles of Scientific Thinking', in P. Bevilacqua & P. J. Kennedy (eds.)
Using History of Physics in lnnovatory Physics Edcuation, University of Pavia, pp. 106-
127.
Bybee, R. W.: 1990, 'Teaching History and the Nature of Science in Science Courses: A
Rationale', Science Education.
Carey, S.: 1986, 'Cognitive Psychology and Science Education', American Psychologist 41,
1123-1130.
Cawthorn, E. R. & Rowell, J. A.: 1978, 'Epistemology & Science Education', Studies in
Science Education 5, 31-59.
Champagne, A, B., Klopfer, L. E. & Anderson, J.: 1980, 'Factors Influencing Learning of
Classical Mechanics', American Journal of Physics 48, 1074-1079.
Champagne, A. B., Gunstone, R. F., and Klopfer, L. E.: 1983, 'Naive Knowledge and
Science Learning', Research in Science and Technology Education 1(2), 173-183.
Clement, J.: 1983, 'A Conceptual Model Discussed by Gahleo and Intuitively Used by
Physics Students', in D. Genter and A. L. Stevens, (eds.), Mental Models, Erlbaum,
Hillsdale, NJ, pp. 325-339.
Cohen, I. B.: 1950, 'A Sense of History in Science', American Journal of Physics 18, 343-
359.
Cohen, I. B.: 1977, 'History and the Philosopher of Science', in P. Suppes (ed.) The Structure
of Scienttfic Theories, 2nd edition, University of Illinois Press, Urbana, pp. 308-360.
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 41

Collins, A.: 1989, 'Assessing Biology Teachers: Understanding the Nature of Science and Its
Influence on the Practice of Teaching', in D. E. Herget (ed.) The History and Philosophy of
Science in Science Teaching, Florida State University, pp. 61-70.
Conant, J. B.: 1945, General Education in a Free Society: Report of the Harvard Committee,
Harvard University Press, Cambridge.
Conant, J. B.: 1947, On Understanding Science, Yale University Press, New Haven.
Conant, J. B.: Science and Common Sense, Yale University Press, New Haven.
Conant, J. B. (ed.): 1957, Harvard Case Histories in Experimental Science, 2 vols., Harvard
University Press, Cambridge.
Connelly, F. M.: 1974, 'Significant Connections Between Philosophy of Science and Science
Education', Studies in Philosophy and Education 8, 245-257.
Crombie, A. C.: 1981, 'Philosophical Presuppositions and the Shifting Interpretations of
Galileo', in J. Hintikka et al. (eds.) Theory Change, Ancient Axiomatics, and Galileo's
Methodology, Reidel, Boston.
Dijksterhuis, E. J.: 1986, The Mechanization of the World Picture, Princeton University
Press, Princeton (orig. 1961).
DiSessa, A. A.: 1982, 'Unlearning Aristotelian Physics: A Study of Knowledge-Based Learn-
ing', Cognitive Science 6, 37-75.
Drake, S.: 1978, Galileo At Work, University of Chicago Press, Chicago.
Drake, S.: 1980, Galileo, Oxford University Press, Oxford.
Driver, R. et al. (eds.): 1985, Children's Ideas in Science, Open University Press, Milton
Keynes.
Driver, R. & Oldham, V.: 1986, 'A Constructivist Approach to Curriculum Development
in Science', Studies in Science Education 13, 105-122.
Dublin, M.: 1989, 'J. J. Schwab - A Memoir and a Tribute', Interchange 20(2), 112-118.
Duhem, P.: 1906/1954, The Aim & Structure of Physical Theory, trans. P. P. Wiener,
Princeton University Press, Princeton.
Duschl, R. A.: 1985, 'Science Education & Philosophy of Science, Twenty-five Years of
Mutually Exclusive Development', School Science and Mathematics 87(7), 541-555.
Duschl, R. A.: 1988, 'Abandoning the Scientistic Legacy of Science Education', Science
Education 72(1), 51-62.
Duschl, R. A., Hamilton, R. and Grandy, R. E.: 1990, 'Psychology and Epistemology:
Match or Mismatch when Applied to Science Education?', InternationalJournal of Science
Education 12(3), 230-243.
Eger, M.: 1972, 'Physics & Philosophy: A Problem for Education Today', American Journal
of Physics 40, 404-415.
Eger, M.: 1987, 'Philosophy of Science in Teacher Education', in J. D. Novak (ed.) Miscon-
ceptions & Educational Strategies, Cornell University, vol I, pp. 163-176.
Eger, M.: 1988, 'A Tale of Two Controversies: Dissonance in the Theory and Practice of
Rationality', Zygon 23(3), 291-326. Reprinted in M. R. Matthews (ed.) History, Philos-
ophy and Science Teaching: Selected Readings, OISE Press, Toronto and Teachers College
Press, New York 1991.
Eger, M.: 1989, 'The "Interests" of Science and the Problems of Education', Synthese 80(1),
81-106.
Eger, M.: 1989b, 'Rationality and Objectivity in a Historical Approach: A Response to
Harvey Siegel', D. E, Herget (ed.), The History and Philosophy of Science in Science
Teaching, Florida State University, Tallahassee, pp. 143-153.
Eijkelhof, H. & Swager, J.: 1983, Physics in Society: New Trends in Physics Teaching IV,
UNESCO, Paris.
Ennis, R. H.: 1979, "Research in Philosophy of Science Bearing on Science Education', in
P. D. Asquith & H. E. Kyburg (eds.) Current Research in Philosophy of Science, PSA,
East Lansing, pp. 138-170.
Feyerabend, P. K.: 1975, Against Method, New Left Books, London.
Finocchiaro, M. A.: 1980a, Galileo and the Art of Reasoning, Dordrecht, Reidel.
Finocchiaro, M. A. et al.: 1980b, 'A Symposium on the Use of the History of Science in the
42 M I C H A E L R. M A T T H E W S

Science Curriculum', Journal of College Science Teaching 10(i), 14-33.


Floden, R. E., Buchmann, M., & Schwille, J. R.: 1987, 'Breaking with Everyday Experi-
ence', Teachers College Record 88(4), 485-506.
Forge, J. C.: 1979, 'A Role for the Philosophy of Science in the Teaching of Science',
Journal of Philosophy of Education 13, 109-118.
Furiomas, C. J. et al.: 1987, 'Parallels Between Adolescents' Conceptions of Gases and the
History of Chemistry', Journal of Chemical Education 64(7), 616-618.
Garrison, J. & Bentley, M.: 1989, 'Science Education, Conceptual Change, and Breaking
with Everyday Experience', Studies in Philosophy and Education 10(1), 19-36.
Gagne, R. M.: 1963, 'The Learning Requirements for Enquiry', Journal of Research in
Science Teaching 1(2), 144-153.
Giere, R. N.: 1973, 'History and Philosophy of Science: Intimate Relationship or Marriage
of Convenience', British Journal for the Philosophy of Science 24,282-297.
Giere, R. N.: 1987, 'The Cognitive Study of Science', in N. J. Nersessian (ed.) The Process
of Science, Martinus Nijhoff, Dordrecht, pp. 139-160.
Gill, W. (ed.): 1977, 'Symposium on History, Philosophy & Science Teaching', The Aus-
tralian Science Teachers Journal 23(2), 4-91.
Ginev, D.: 1989, 'Toward a New Image of Science', Studies in Philosophy and Education
10(1), 63-72.
Gjertsen, D.: 1989, Science and Philosophy, Penguin, Harmondsworth.
Glasersfeld, E. yon: 1989, 'Cognition, Construction of Knowledge, and Teaching', Synthese
80(1), 121-140.
Gosling, D. & Musschenga, B. (eds.): 1985, Science Education & Ethical Values, Georgetown
University Press, Washington D.C.
Gotschl, J.: 1990, 'Philosophical and Scientific Conceptions of Nature and the Place of
Responsibility', International Journal of Science Education 12(3), 288-296.
Gruender, C. D.: 1989, 'Some Philosophical Reflections on Constructivism', in D. E. Herget
(ed.) The History and Philosophy of Science in Science Teaching, Florida State University,
pp. 170-176.
Harding, S.: 1986, The Science Question in Feminism, Cornell University Press, Ithaca.
Harre, R.: 1983, 'History & Philosophy of Science in the Pedagogical Process', in R. W.
Home (ed.), Science Under Scrutiny, Reidel, Dordrecht, pp. 139-157.
Heilbron, J. L.: 1983, 'The Virtual Oscillator as a Guide to Physics Students Lost in Plato's
Cave', in F. Bevilacqua and P. J. Kennedy (eds.), Using History of Physics in Innovatory
Physics Education, Pavia, pp. 162-182.
Heilbron, J. L.: 1987, 'Applied History of Science', ISIS 78, 552-563.
Helm, H., Gilbert, J. & Watts, D. M.: 1985, 'Thought Experiments & Physics Education.
Parts I, II', Physics Education 20, 124-131; 211-217.
Helm, H. & Novak, J. D. (eds.): 1983, Proceedings of the International Seminar on Miscon-
ceptions in Science & Mathematics, Education Department, Cornell University, Ithaca.
Herget, D. E. (ed.): 1989, The History and Philosophy of Science in Science Teaching,
Florida State University, Tallahassee, FL.
Herget, D. E. (ed.): 1990, More History and Philosophy of Science in Science Teaching,
Florida State University, Tallahassee, FL.
Herron, M. D.: 1971, 'The Nature of Scientific Inquiry', School Review 79, 170-212.
Hesse, M. B.: 1988, '"Rationality" in Science and Morals', Zygon 23(3), 327-332.
Hodson, D.: 1982, "Science - the Pursuit of Truth? Parts I, II', School Science Review
63(225), 643-652; (226), 23-30.
Hodson, D.: 1986a, 'Philosophy of Science and the Science Curriculum', Journal of Philos-
ophy of Education 20, 241-251.
Hodson, D.: 1986b, "Rethinking the Role & Status of Observation in Science Education',
Journal of Curriculum Studies 18(4), 381-396.
Hodson, D. : 1988a, 'Experiments in Science and Science Teaching', Educational Philosophy
and Theory 20(2), 53-66.
H I S T O R Y , P H I L O S O P H Y , AND SCIENCE T E A C H I N G 43

Hodson, D.: 1988b, 'Toward a Philosophically More Valid Science Curriculum', Science
Education 72, 19-40.
Holton G.: 1952, Introduction to Concepts & Theories in Physical Science, Addison-Wesley,
New York.
Holton, G. et al.: 1967, 'Symposium on the Project Physics Course', The Physics Teacher
5(5), 196-231.
Holton, G.: 1975, 'Science, Science Teaching & Rationality', in S. Hook et al. (eds.) The
Philosophy of the Curriculum, Promethus Books, Buffalo.
Holton, G.: 1978, 'On the Educational Philosophy of the Project Physics Course', in his The
Scientific Imagination: Case Studies, Cambridge University Press, Cambridge.
Holton, G.: 1986, '"A Nation At Risk" Revisited', in his The Advancement of Science &
Its Burdens, Cambridge University Press, Cambridge.
Home, R. W. (ed.): 1983, Science Under Scrutiny: The Place of History and Philos.ophy of
Science, Reidel, Dordrecht.
Jacoby, B. A. & Spargo, P. E.: 1989, 'Ptolemy Revived?', Interchange 20(2), 33-53.
Jenkins, E.: 1990, 'History of Science in Schools: Retrospect and Prospect in the UK',
International Journal of Science Education 12(3), 274-281. Reprinted in M. R. Matthews
(ed.) History, Philosophy and Science Teaching: Selected Readings, OISE Press, Toronto
and Teachers College Press, New York 1991.
Jung, W.: 1986, 'Cognitive Science and the History of Science', in P. V. Thomsen (ed.)
Science Education and the History of Physics, University of Aarhus.
Kauffman, G. B.: 1989, 'History in the Chemistry Curriculum', Interchange 20(2), 81-94.
Reprinted in M. R. Matthews (ed.) History, Philosophy and Science Teaching: Selected
Readings, OISE Press, Toronto and Teachers College Press, New York 1991.
Keller, E. F.: 1985, Reflections on Gender and Science, Yale University Press, New Haven.
Kitcher, P.: 1988, "The Child as Parent of the Scientist', Mind and Language 3(3), 217-
228.
Kitchner, R. F.: 1985, 'A Bibliography of Philosophical Work on Piaget', Synthese 65(1),
139-151.
Klein, M. J.: 1972, 'Use and Abuse of Historical Teaching in Physics', in S. G. Brush & A.
L. King (eds.), History in the Teaching of Physics, University Press of New England,
Hanover.
Klopfer, L. E. & Cooley, W. W.: 1963, 'The History of Science Cases for High Schools in
the Development of Student Understanding of Science and Scientists: A Report on the
HOSC Instruction Project', Journal of Research in Science Teaching 1, 33-47.
Koulaidis, V. & Ogborn, J.: 1989, 'Philosophy of Science: An Empirical Study of Teachers'
Views', International Journal of Science Education 11(2), 173-184.
Koyrr, A.: 1939/1978, Galileo Studies, trans. J. Mepham, Harvester Press, Hassocks, Sussex.
Koyr6, A.: 1943, 'Galileo and Plato', Journal of the History of Ideas 4, 400--428. (Reproduced
in his Metaphysics and Measurement, 1968.)
Koyrr, A.: 1986, Metaphysics and Measurement, Harvard University Press, Cambridge.
Kragh, H.: 1987, An Introduction to the Historiography of Science, Cambridge University
Press, Cambridge.
Krajewski, W. (ed.): 1982, Polish Essays in the Philosophy of the Natural Sciences, Reidel,
Dordrecht.
Krasilchik, M.: 1990, 'The "Scientists": A Brazilian Experiment in Science Education',
International Journal of Science Education 12(3), 282-287.
Kuhn, T. S.: 1959, 'The Essential Tension: Tradition and Innovation in Scientific Research',
The Third University of Utah Research Conference on the Identification of Scientific Talent,
University of Utah Press, Salt Lake City. Reprinted in his The Essential Tension, Univer-
sity of Chicago Press, Chicago, pp. 225-239.
Kuhn, T. S.: 1977, "Concepts of Cause in the Development of Physics' in his The Essential
Tension, University of Chicago Press, Chicago, pp. 21-30.
Lakatos, I.: 1978, 'History 6f Science and Its Rational Reconstructions', in J. Worrall and
44 M I C H A E L R. M A T T H E W S

G. Currie (eds.) The Methodology of Scientific Research Programmes, Cambridge Univer-


sity Press, Cambridge, pp. 102-138.
Lederman, N. G. & Zeidler, D. L.: 1987, 'Science Teachers Conceptions of the Nature of
Science: Do They Really Influence Teaching Behaviour?', Science Education 71(5), 721-
734.
Lipman, M. & Sharp, A. M. (eds.): 1978, Growing Up with Philosophy, Temple University
Press, Philadelphia.
Loving, C. C.: 1992, 'The Scientific Theory Profile: A Philosophy of Science Model for
Science Teachers', Journal of Research in Science Teaching.
Mach, E.: 1883/1960, The Science of Mechanics, Open Court Publishing Company, LaSalle
I1.
Mach, E.: 1895/1943, 'On Instruction in the Classics & the Sciences', in his Popular Scientific
Lectures, Open Court, LaSalle.
Machamer, P.: 1978, 'Galileo and the Causes', in R. E. Butts and J. C. Pitt (eds.) New
Perspectives on Galileo, Dordrecht, Reidel, pp. 161-181.
MacLachlan, J.: 1990, 'Drake Against the Philosophers', in T. H. Levere & W. R. Shea
(eds.) Nature, Experiment, and the Sciences, Kluwer, Dordrecht, pp. 123-144.
Manuel, D. E.: 1981, 'Reflections on the Role of History & Philosophy of Science in School
Science Education', School Science Review 62(221), 769-771.
Markus, G.: 1987, 'Why Is There No Hermeneutics of Natural Sciences: Some Preliminary
Theses', Science in Context 1(1), 5-51.
Martin, J. R.: 1989, 'Ideological Critiques and the Philosophy of Science', Philosophy of
Science 56, 1-22.
Martin, J. R.: 'What Should Science Educators Do About the Gender Bias in Science?', in
M. R. Matthews (ed.) History, Philosophy and Science Teaching: Selected Readings, OISE
Press, Toronto and Teachers College Press, New York 1991, pp. 151-166.
Martin, M.: 1972, Concepts of Science Education, Scott, Foresman & Co., New York
(reprinted, University Press of America, 1985).
Martin, M.: 1986, 'Science Education & Moral Education', Journal of Moral Education
15(2), 99-108. Reprinted in M. R. Matthews (ed.) History, Philosophy and Science
Teaching: Selected Readings, OISE Press, Toronto and Teachers College Press, New York
1991.
Mas, C. J., Perez, J. H., and Harris, H. H.: 1987, 'Parallels between Adolescents' Concep-
tion of Gases and the History of Chemistry', Journal of Chemical Education 64(7), 616-
618.
Matthews, M. R.: 1987, 'Galileo's Pendulum & the Objects of Science', in B. & D. Arnstine
(eds.) Philosophy of Education, pp. 309-319, Philosophy of Education Society.
Matthews, M. R.: 1988, 'A Role for History and Philosophy in Science Teaching', Edu-
cational Philosophy and Theory 20(2), 67-81.
Matthews, M. R.: 1989a, 'Ernst Mach and Thought Experiments in Science Education',
Research in Science Education 18, 251-258.
Matthews, M. R. (ed.): 1989b, The Scientific Background to Modern Philosophy, Hackett
Publishing Company, Indianapolis.
Matthews, M. R.: 1990a, 'History, Philosophy, and Science Teaching: The Case of Pendulum
Motion', Research in Science Educatton 19, 187-198.
Matthews, M. R.: 1990b, 'History, Philosophy, and Science Teaching: What Can Be Done
in an Undergraduate Course?', Studies in Philosophy and Education 10(1), 93-98.
Matthews, M. R. (ed.): 1991, History, Philosophy, and Science Teaching: Selected Readings,
OISE Press, Toronto.
McCloskey, M.: 1983, 'Intuitive Physics', Scientific American 248, 114-122.
McDermott, L. C.: 1984, 'Research on Conceptual Understanding in Mechanics', Physics
Today 37, 24-32.
McMullin, E.: 1975, 'History and Philosophy of Science: a Marriage of Convenience?',
Boston Studies in the Philosophy of Science 32, 515-531.
HISTORY, PHILOSOPHY, AND SCIENCE TEACHING 45

McTighe, T. P.: 1967, 'Galileo's Platonism: A Reconsideration', in E. McMullen (ed.)


Galileo: Man of Science, Basic Books, New York, pp. 365-388.
Mendelsohn, E.: 1976, 'Values and Science: a Critical Reassessment', The Science Teacher
43(1), 20-23.
Merton, R. K.: 1977, 'The Sociology of Science: An Episodic Memoir', in R. K. Merton &
J. Gaston (eds.) The Sociology of Science in Europe, Southern Illinois University Press,
Carbondale, IL., pp. 3-141.
Miller, J. D.: 1983, 'Scientific Literacy: A Conceptual & Empirical Review', Daedalus
112(2), 29-47.
Mischel, T.: 1971, 'Piaget: Cognitive Conflict and the Motivation of Thought', in T. Mischel
(ed.) Cognitive Development and Epistemology, New York, pp. 311-355.
Mittelstrass, J.: 1972, 'The Galilean Revolution: The Historical Fate of a Methodological
Insight', Studies in the History and Philosophy of Science 2, 297-328.
Murray, F. R. (ed.): 1979, The Impact of Piagetian Theory on Education, Philosophy,
Psychiatry, and Psychology, Baltimore.
National Curriculum Council: 1988, Science in the National Curriculum, NCC, York.
Nersessian, N. J.: 1989, 'Conceptual Change in Science and in Science Education', Synthese
80(1), 163-184. Reprinted in M. R. Matthews (ed.) History, Philosophy and Science
Teaching: Selected Readings, OISE Press, Toronto and Teachers College Press, New York
1991.
Nielsen H. & Thomsen, P.: 1990, 'History and Philosophy of Science in the Danish Curricu-
lum', International Journal of Science Education 12(3), 308-316.
Norris, S. P.: 1985, The Philosophical Basis of Observation in Science and Science Edu-
cation', Journal of Research in Science Teaching 22(9), 817-833.
Novak, J. D.: 1987, 'Human Constructivism: Toward a Unity of Psychological & Epistemo-
logical Meaning Making', in J. D. Novak (ed.), 1987, Misconceptions & Educational Stra-
tegies, vol I, pp. 349-360.
Novak, J. D. (ed.): 1987, Proceedings of the Second International Seminar on Misconceptions
& Educational Strategies in Science & Mathematics, Education Department, Cornell Uni-
versity, Ithaca.
Nowak, L.: 1980, The Structure of Idealization, Reidel, Dordrecht.
Nussbaum, J.: 1983, 'Classroom Conceptual Change: The Lesson to be Learned from the
History of Science', in H. Helm & J. D. Novak (eds.) Misconceptions in Science &
Mathematics, Department of Education, Cornell University, pp. 272-281.
Passmore, J.: 1978, Science and its Critics, Duckworth, London.
Piaget, J.: 1970, Genetic Epistemology, Columbia University Press, London.
Piaget, J. & Garcia, R.: 1989, Psychogenesis and History, Columbia University Press, New
York.
Posner, G~ et al.: 1982, 'Accommodation of a Scientific Conception: Towards a Theory of
Conceptual Change', Science Education 66(2), 211-227.
Pumfrey, S.: 1987, 'The Concept of Oxygen: Using History of Science in Science Teaching',
in M. Shortland & A. Warwick (eds.) Teaching the History of Science, Basil Blackwell,
Oxford, pp. 142-155.
Randall, J. H.: 1940, 'The School of Padua and the Emergence of Modern Science', Journal
of the History of Ideas 1, 177-206. Reprinted in his The Career of Philosophy, Volume
One, Columbia University Press, New York, 1961.
Robin, N. and Ohlsson, S.: 1989, 'Impetus Then and Now: A Detailed Comparison between
Jean Buridan and a Single Contemporary Subject', in D. E. Hegert (ed.), The History
and Philosophy of Science Teaching, Florida State University, pp. 292-305.
Rohrlich, F.: 1988, 'Four Philosophical Issues Essential for Good Science Teaching', Edu-
cational Philosophy and Theory 20(2), 1-6.
Rowell, J. A.: 1989, 'Piagetian Epistemology: Equilibration and the Teaching of Science',
Synthese 80(1), 141-162.
46 MICHAEL R. M A T T H E W S

Rowell, J. A. & Cawthron. E. R.: 1982, 'Images of Science: An Empirical Study', European
Journal of Science Education 4(1), 79-94.
Russell, T. L.: 1981, 'What History of Science, How Much and Why?' Science Education
65, 51-64.
Schecker, H.: 1988, "The Paradigmatic Change in Mechanics: Implications of Historical
Processes on Physics Education', in C. Blondel & P. Brouzeng (eds.) Science Education
and the History of Physics, pp. 215-220. Reprinted this volume.
Schemer, I.: 1970, 'Philosophy and the Curriculum', in his Reason and Teaching, London,
Routledge, 1973, pp. 31-44.
Schilpp, P. A. (ed.): 1951, Albert Einstein, second edition, Tudor, New York.
Schwab, J. J.: 1963, Biology Teacher's Handbook, Wiley, New York.
Schwab, J.: 1964, 'Structure of the Disciplines: Meaning & Significances', in G. W. Ford &
L. Pugno (eds.) The Structure of Knowledge & the Curriculum, Rand McNally & Co.,
Chicago.
Siegel, H.: 1979, 'On the Distortion of the History of Science in Science Education', Science
Education 63, 111-118.
Siegel, H.: 1982, 'On the Parallel between Piagetian Cognitive Development & the History
of Science', Philosophy of Social Science 12, 375-386.
Settle, T.: 1990, 'How to Avoid Implying that Physicalism is True: A Problem for Teachers
of Science', International Journal of Science Education 12(3), 258-264.
Shapere, D.: 1984, Reason and the Search for Knowledge, Reidel, Dordrecht.
Shea, W. R.: 1972, Galileo's Intellectual Revolution, Macmillan, London.
Sherratt, W. J.: 1982, 'History of Science in the Science Curriculum: An Historical Perspec-
tive', School Science Review 64, 225-236, 418-424.
Shortland, M. & Warwick, A. (eds.): 1989, Teaching the History of Science, Basil Blackwell,
Oxford.
Shuell, T. 1987, "Cognitive Psychology and Conceptual Change: Implications for Teaching
Science'. Science Education 71,239-250.
Shulman, L. S.: 1986, 'Those Who Understand: Knowledge Growth in Teaching', Edu-
cational Researcher 15(2), 4-14.
Shulman, L. S.: 1987, 'Knowledge and Teaching: Foundations of the New Reform', Harvard
Educational Review 57(1), 1-22.
Siegel, H.: 1987, Relativism Refuted, Reidel, Dordrecht.
Siegel, H.: 1989, 'The Rationality of Science, Critical Thinking, and Science Education',
Synthese 80(1), 9-42. Reprinted in M.R. Matthews (ed.) History, Philosophy and Science
Teaching: Selected Readings, OISE Press, Toronto and Teachers College Press, New York
1991.
Solomon, J.: 1990, 'Teaching About the Nature of Science in the British National Curricu-
lum', Science Education.
Stein, F.: 1989, 'Project 2061: Education for a Changing Future', in D. E. Herget (ed.) The
History and Philosophy of Science in Science Teaching, Florida State University, pp. 339-
343.
Steinberg, M. S., Brown, D. E., & Clement, J.: 'Genius is not Immune to Persistent
M~sconceptions: Conceptual Difficulties Impeding Isaac Newton and Contemporary Phys-
ics Students', International Journal of Science Education 12(3), 256-273.
Stevens, P. : 1978, 'On the Nuffield Philosophy of Science', Journal of Philosophy of Edu-
cation 12, 99-111.
Stinner, A.: 1990, 'Philosophy, Thought Experiments, and Large Context Problems in the
Secondary School Physics Course', International Journal of Science Education 12(3), 244-
257.
Strike, K. A.: 1987, 'Towards a Coherent Constructivism', in J. D. Novak (ed.) Misconcep-
tions & Educational Strategies, Education Department, Cornell University, vol. I, pp. 481-
489.
Suchting, W. A.: 1992, 'Constructivism Deconstructed', Science & Education 1 (3).
H I S T O R Y , P H I L O S O P H Y , AND SCIENCE T E A C H I N G 47

Summers, M. K.: 1982, 'Philosophy of Science in the Science Teacher Education Curriculum'
European Journal of Science Education 4, 19-28.
Swift, J. N.: 1988, 'The Tyranny of Terminology: Biology', The Science Teachers Bulletin
60(2), 24-26.
Tamir, P.: 1989, 'History and Philosophy of Science and Biological Education in Israel',
Interchange 20(2), 95-98.
Thayer, H. S. (ed.): 1953, Newton's Philosophy of Nature, Macmillian, New York.
Thomson, J. J.: 1918, Natural Science in Education, HMSO, London. (Known as the Thom-
son Report.)
Thomsen, P. V. (ed.): 1986, Science Education and the History of Physics, University of
Aarhus.
Toulmin, S. E.: 1983, 'The Construal of Reality: Criticism, in Modern & Post-Modern
Science', in W. J. T. Mitchell (ed.), The Politics of Interpretation, University of Chicago
Press, Chicago, pp. 99-117. (Originally, Critical Inquiry 9(1), 1982, 93-111.)
Viennot, L.: 1979, 'Spontaneous Reasoning in Elementary Dynamics', European Journal of
Science Education 1,205-221.
Wagenschein, M.: 1962, Die Padagogische Dimension der Physik, Westermann,
Braunschweig.
Wallace, W. A.: 1981, 'Galileo and Reasoning ex suppositione', in W. A. Wallace, Prelude
to Galileo, Reidel, Dordrecht, pp. 129-159.
Wallace, W. A.: 1984, Galileo and His Sources, Princeton University Press, Princeton.
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.
Wartofsky, M.: 1968, 'Metaphysics as a Heuristic for Science', in R. S. Cohen & M. W.
Wartofsky (eds.) Boston Studies in the Philosophy of Science 3, 123-172. Republished in
his Models, Reidel, Dordrecht, 1979, pp. 40-89.
Wartofsky, M. W.: 1976, 'The Relation Between Philosophy of Science and History of
Science', in R. S. Cohen, P. K. Feyerabend & M. W. Wartofsky (eds.) Essays in Memory
of Imre Lakatos, Reidel, Dordrecht. (Boston Studies in the Philosophy of Science Vol.
39). Republished in his Models, Reidel, Dordrecht, 1979.
Welch, W.: 1979, 'Twenty Years of Science Education Development: A Look Back', Review
of Research in Education 7,282-306.
Whewell, W.: 1840/1947, Philosophy of the Inductive Sciences, London.
Whitaker, M. A. B.: 1979, 'History and Quasi-history in Physics Education Pts I, II', Physics
Education 14, 108-I 12, 239-242.
Whitaker, R. J.: 1983, 'Aristotle is Not Dead: Student Understanding of Trajectory Motion',
American Journal Physics 51(4), 352-357.
Winchester, I.: 1990, 'Thought Experiments and Conceptual Revision in Science', Studies
in Philosophy and Education 10(1), 73-80.
Woolnough, B. E.: 1989, 'Faith in Science', School Science Review 70(252), 133-137. Re-
printed in M. R. Matthews (ed.) History, Philosophy and Science Teaching: Selected
Readings, OISE Press, Toronto and Teachers College Press, New York 1991.
Ziman, J.: 1980, Teaching and Learning about Science and Society, Cambridge University
Press, Cambridge.

Vous aimerez peut-être aussi