Feyerabend’s Epistemological Anarchism: How Science Works And Its Importance For Science Education 3030368580, 9783030368586, 9783030368593

This book argues that the traditional image of Feyerabend is erroneous and that, contrary to common belief, he was a gre

511 76 2MB

English Pages 233 Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Feyerabend’s Epistemological Anarchism: How Science Works And Its Importance For Science Education
 3030368580,  9783030368586,  9783030368593

Table of contents :
Endorsements......Page 6
Preface......Page 8
Acknowledgments......Page 11
Contents......Page 12
Chapter 1: Introduction: Exploring Epistemological Anarchism......Page 16
1.1 Origins of Epistemological Anarchism......Page 18
1.1.1 Science Education and Feyerabend......Page 20
1.2 Chapter Outlines......Page 22
2.1 Introduction......Page 38
2.2 Feyerabend’s Epistemological Anarchism......Page 39
2.3 Feyerabend Versus Popper and Lakatos......Page 41
2.4 Was Lakatos an Epistemological Anarchist?......Page 42
2.5 Feyerabend and Scientific Expertise......Page 43
2.6 Feyerabend Versus Galileo and Copernicus......Page 45
2.7 Feyerabend and Recent Philosophy of Science......Page 47
2.8 Feyerabend and Perspectivism......Page 50
2.9 Feyerabend and Feminist Epistemology......Page 51
2.10 Feyerabend and the Practice of Science......Page 52
3.1 Method......Page 54
3.1.2 Classification of Articles......Page 55
3.2 Results and Discussion......Page 56
3.2.1 Acid-Base Equilibria......Page 57
3.2.2 Anarchistic Methodology......Page 60
3.2.3 Constructivism......Page 61
3.2.4 Critical Thinking......Page 63
3.2.5 Diversity of Methods......Page 64
3.2.7 Evolution, Knowledge and Belief (to Give Meaning to Life)......Page 65
3.2.9 Incommensurability......Page 70
3.2.11 Newtonian Method......Page 72
3.2.12 Normal Science, Dogmatism and Science Education......Page 74
3.2.13 Polanyi’s Tacit Knowledge......Page 76
3.2.14 Science and Religion......Page 77
3.2.15 Scientific Expertise and Galileo......Page 78
3.2.16 Scientific Method......Page 80
3.2.17 Situated Learning......Page 83
4.1 Method......Page 85
4.2.1 Alternative Literary Forms......Page 86
4.2.3 Evaluation......Page 87
4.2.5 Nature of Science......Page 88
4.2.6 Proliferation of Theories......Page 89
4.2.7 Scientific Method......Page 90
4.2.8 Teacher Demonstrations......Page 91
4.2.9 Worldviews......Page 92
5.1 Method......Page 94
5.2.1 African and Modern Medicine......Page 95
5.2.2 Alternative Approaches to Growth of Knowledge......Page 96
5.2.3 Constructive Alternativism......Page 98
5.2.4 Diversity of Rival Theories......Page 99
5.2.5 Genius in Science......Page 100
5.2.6 History of Science......Page 101
5.2.7 Objectivity Versus Subjectivity......Page 102
5.2.8 Presuppositions of Science Teachers......Page 103
5.2.9 Rationalism......Page 107
5.2.10 Scientific Method......Page 108
6.1 Method......Page 110
6.2.1 Historical-Investigative Approach to Science......Page 111
6.2.2 Kuhn and Normal Science......Page 112
6.2.3 Nature of Science......Page 113
6.2.4 Postmodernism......Page 115
6.2.5 School Science Curriculum......Page 117
6.2.6 Science as Cultural Tyranny......Page 119
Chapter 7: Feyerabend’s Counterinduction and Science Textbooks......Page 121
7.1 Brownian Motion......Page 122
7.2 Kinetic Theory of Gases......Page 123
7.3 Michelson-Morley Experiment......Page 126
7.4 The Oil-Drop Experiment......Page 128
7.5 Alpha Particle Scattering Experiment......Page 135
7.6 Bohr’s Incorporation of “quantum of action” to Classical Electrodynamics......Page 137
7.7 Photoelectric Effect......Page 140
7.8 Wave-Particle Duality......Page 143
7.9 Mendeleev’s Periodic Table of Chemical Elements......Page 146
7.10 Lewis’s Postulation of the Covalent Bond......Page 156
7.11 Discovery of the Planet Neptune......Page 159
7.12 Discovery of the Elementary Particle Neutrino......Page 161
7.13 Discovery of the Tau Lepton......Page 162
8.1 Feyerabend’s Hyperbolic Flourishes......Page 167
8.2 Feyerabend’s Epistemological Anarchism......Page 168
8.2.1 Counterinduction (Accepting Unsupported Hypotheses)......Page 169
8.2.2 Current View of a Science May Soon Be Voted Out of Office......Page 172
8.2.3 Does Science Always Provide the One “Correct” Model (Theory)......Page 174
8.2.5 History of a Science Becomes an Inseparable Part of the Science Itself......Page 175
8.2.6 Inferring Objectivity from Empirical Approaches......Page 177
8.2.7 Methodological Pluralism: Diversity of Methods......Page 178
8.2.8 Nature of Science......Page 179
8.2.11 Scientific Method: Stockpiling and Ordering of Observations......Page 180
8.2.13 The New Grew Out of the Old......Page 181
8.2.14 Unnatural Nature of Science......Page 182
8.2.15 Was Feyerabend a Postmodern or Perspectival Realist?......Page 183
8.3 Educational Implications......Page 184
Appendix 1: Articles from the Journal Science & Education (Springer) Evaluated in This Study (n = 78)......Page 186
Appendix 2: Distribution of Articles (Science & Education) According to Author’s Area of Research, Context of the Study and Level (Classification), n = 78......Page 190
Appendix 3: Articles from the Journal of Research in Science Teaching (Wiley Blackwell) Evaluated in This Study, n = 21......Page 194
Appendix 4: Distribution of Articles (Journal of Research in Science Teaching) According to Author’s Area of Research, Context of the Study and Level (Classification), n = 21......Page 195
Appendix 5: Articles from the Journal Interchange Evaluated in This Study (n = 15)......Page 196
Appendix 6: Distribution of Articles (Interchange) According to Author’s Area of Research, Context of the Study and level (Classification), n = 15......Page 197
Appendix 7: Articles from the International Handbook of Research in History, Philosophy and Science Teaching (Springer), n = 6......Page 198
Appendix 9: List of General Chemistry Textbooks Evaluated in Different Studies of This Book (n = 128)......Page 199
Appendix 10: List of General Physics Textbooks Evaluated in Different Studies of This Book (n = 103)......Page 205
References......Page 210
Name Index......Page 225
Subject Index......Page 230

Citation preview

Contemporary Trends and Issues in Science Education 50

Mansoor Niaz

Feyerabend’s Epistemological Anarchism How Science Works and its Importance for Science Education

Contemporary Trends and Issues in Science Education Volume 50

Series Editors Dana L. Zeidler, University of South Florida, Tampa, USA Editorial Board Michael P. Clough, Iowa State University, Ames, IA, USA Fouad Abd-El-Khalick, The University of North Carolina, Chapel Hill, NC, USA Marissa Rollnick, University of the Witwatersrand, Johannesburg, South Africa Troy D. Sadler, University of Missouri, Columbia, MO, USA Svein Sjøeberg, University of Oslo, Oslo, Norway David Treagust, Curtin University of Technology, Perth, Australia Larry D. Yore, University of Victoria, British Columbia, Canada

The book series Contemporary Trends and Issues in Science Education provides a forum for innovative trends and issues connected to science education. Scholarship that focuses on advancing new visions, understanding, and is at the forefront of the field is found in this series. Accordingly, authoritative works based on empirical research and writings from disciplines external to science education, including historical, philosophical, psychological and sociological traditions, are represented here.

More information about this series at http://www.springer.com/series/6512

Mansoor Niaz

Feyerabend’s Epistemological Anarchism How Science Works and its Importance for Science Education

Mansoor Niaz Department of Chemistry Universidad de Oriente Cumaná, Sucre, Venezuela

ISSN 1878-0482     ISSN 1878-0784 (electronic) Contemporary Trends and Issues in Science Education ISBN 978-3-030-36858-6    ISBN 978-3-030-36859-3 (eBook) https://doi.org/10.1007/978-3-030-36859-3 © Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG. The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

Endorsements

This is a very important book. Paul Feyerabend and Imre Lakatos contend that science advances through a multiplicity of methods. It is important not to disparage any methodology. Feyerabend infamously stated “anything goes.” Feyerabend has often been misunderstood because of his provocative statements, and some have characterized him as being antiscience. To the contrary, Feyerabend as Niaz shows through this remarkable book penetrates to the essence of science. Feyerabend has pointed out that producing genuinely new knowledge requires considerations of two incommensurable theories (principle of counterinduction). In his seminal book Against Method analyzing Galileo, Feyerabend shows that if you only consider one theory, you will not come to another theory even if your theory is not entirely supported by experimental evidence. Niaz carefully analyzes references to Feyerabend in the literature and displays the importance of Feyerabend’s philosophy in analyzing historical episodes. Calvin Kalman, Concordia University, Canada Paul Feyerabend’s “epistemological anarchism” remains a controversial perspective within the philosophy of science but offers a valuable alternative to empiricism and realism and an antidote to scientism. In this book, Mansoor Niaz explores the antecedents, contexts, and features of Feyerabend’s work, offers a more nuanced understanding, and then reviews and considers its reception in the science education and philosophy of science literature. He also considers its influence in science textbooks and on the teaching of science and explores ways in which epistemological anarchism can complement and enrich our approaches to the learning and teaching of science. This is a valuable contribution to scholarship about Feyerabend, with the potential to inform further research as well as science education practice. I thoroughly enjoyed the book and commend it to anyone interested in science education and its intellectual underpinnings in the philosophy of science. David Geelan, Griffith University, Australia

v

To Magda and Sabuhi For their love, patience, and support

Preface

Almost 30 years ago when I started reading the literature on history and philosophy of science (HPS) (Popper, Polanyi, Hanson, Kuhn, Lakatos, Laudan, and Feyerabend), the empirical nature of science was considered as an unquestioned axiom of both science and science education. Although Feyerabend was included in the reading list, most colleagues considered that there was not much to learn due to his postmodern and relativist stance on HPS. Empirical science was like a jewel in the crown of the science curriculum. I vividly recall a symposium at a major science education conference in the 1990s that discussed the role of the epistemic subject in Piaget’s genetic epistemology. Based on Galilean idealization, a major cognitive psychologist argued that Piagetian stage theory cannot be refuted by accumulating empirical data. On the contrary, a renowned science educator reasoned that based on empirical evidence, Piagetian stage theory could not be sustained. Some of the papers were published in the Journal of Research in Science Teaching and Science & Education. The debate, at times heated, lasted for almost a year with no clear mandate (in my opinion) for the science education community. Even today, most educators and even researchers in science education accept some form of the recipe-­ like scientific method. History of science shows that Newtonian theory could not be refuted for almost 200 years, despite empirical evidence to the contrary. In the first decade of the twentieth century, the landscape started to change as Einstein’s special theory of relativity (not the empirical evidence) challenged Newtonian theory. However, at about the same time, there was also another player on the field whose role has generally been ignored. Physical chemist-philosopher of science Pierre Duhem questioned the empirical (inductive) nature of Newtonian theory by showing that Newton’s claim, that all his hypotheses were based on experimental evidence, was false and at best a dream. Both Feyerabend and Lakatos considered the contribution of Duhem as essential for understanding progress in science. With this background, it is easier to understand Lakatos’ statement, “Inductivism claims that a proposition is scientific if provable from facts; what we shall now set out to do is to show that no proposition whatsoever can be proven from facts.” Even some science educators working within HPS may not agree with this statement. However, if we can go along Lakatos, it ix

x

Preface

would be easier to understand Feyerabend. According to Feyerabend, science as we know it cannot be based entirely on theories that are consistent with all the facts and hence the need to admit counterinduction and unsupported hypotheses. How science works was a major concern of Feyerabend, and following Duhem and Lakatos before, he suggested: traditional science teaching emphasizes objective facts – these are, however, not immutable truths but rather working hypotheses. This chain of thought, Duhem ➔ Lakatos ➔ Feyerabend, took several decades to develop until it crystalized into epistemological anarchism. No wonder, some philosophers of science consider Lakatos to be a “closet anarchist.” How science works was of prime importance to Feyerabend, and this precisely led him to claim that in the practice of science, all rules are broken and hence the infamous “anything goes.” Interestingly, practicing chemist Roald Hoffmann (Nobel Laureate) considers Feyerabend to be a “malevolent genius” who admired science. Hoffmann goes beyond by suggesting that practicing science leads to “violating categories” that approximate to Feyerabend’s epistemological anarchism. This is based on the transgression of method which is embedded in the idea that even a well-confirmed theory could have an alternative, and this leads to understanding how science really works. In both philosophy of science and science education, Feyerabend is generally considered to be against rationalism, antiscience, and for having espoused anything goes and postmodernism. Actually, he questioned those forms of rationalism that were rigid and pompous. In his philosophy of science, anything goes means do not put a limit on your imagination. Although 50 years ago his philosophy of science could be considered as anarchistic and postmodern, at present, Feyerabend’s vision of science coincides with various aspects of philosophy of science in the twenty-first century. Similarly, recent research in philosophy of science has recognized that his oeuvre was more modern (in the Enlightenment tradition) than postmodern. Based on this research, I have argued in this book that the traditional image of Feyerabend is erroneous and that on the contrary he was a great admirer of science and presented a vision of science that represented how science really works. Furthermore, it has been shown that Feyerabend could even be considered as a perspectival realist. In writing this book, I did not have any particular course in mind. This has the advantage that the book could be adopted for various types of courses, such as teaching the nature of science, introduction to the history and philosophy of science, understanding the dynamics of scientific progress, and the role of counterinduction in the history of science. The intended audience for this book is secondary- and university-level teachers, science teacher educators, researchers in science education, and graduate students. Chapter 1 explores the origin of Feyerabend’s epistemological anarchism. The relationship between epistemological anarchism and how science works is established in Chap. 2. Understanding epistemological anarchism in research reported in the journal Science & Education is the subject of Chap. 3. Next, Chap. 4 deals with understanding the epistemological anarchism in the Journal of Research in Science Teaching. Understanding epistemological anarchism in the journal Interchange is the subject of Chap. 5. Chapter 6 deals with understanding the epistemological

Preface

xi

anarchism in a reference work (International Handbook of Research in History, Philosophy and Science Teaching). The relationship between Feyerabend’s counterinduction and science textbooks is explored in Chap. 7. Finally, Chap. 8 deals with the status of Feyerabend’s philosophy of science in the context of the challenges of the twenty-first century. The following are some of the salient features of this book that can help the readers to follow the line of argument developed in the different chapters: 1. How did Feyerabend’s philosophy of epistemological anarchism originate, its influence on other philosophers of science, and its recent appraisal in the philosophy of science literature. 2. A detailed account and evaluation, over a period of over 30 years, of how the science education community conceptualizes Feyerabend’s epistemological anarchism. 3. Science is the gaining of reliable knowledge that is contingent – dependent on the theories of the time. 4. Scientists can accept a hypothesis despite empirical evidence to the contrary. 5. There are no general rules (recipe-like) governing the scientific method. 6. Progress in science is too complex and rich to be based on a set of well-defined a priori rules. 7. Transgression of method is embedded in the idea that even a well-confirmed theory could have an alternative, and this leads to counterinduction (accepting unsupported hypotheses). 8. Questioning methodological rules and emphasizing the practice of science lead to pluralism that helps to check dogmatism. 9. It is desirable to introduce and elaborate hypotheses which are inconsistent with highly confirmed theories. 10. The single best approach to solving a problem is counterproductive as it blinds us to alternative possibilities. Cumaná, Estado Sucre, Venezuela

Mansoor Niaz

Acknowledgments

My institution, Universidad de Oriente (Venezuela), has provided support for research activities over the last many years. Despite my initial skepticism toward Feyerabend, reading Roald Hoffmann on the Philosophy, Art, and Science of Chemistry (2012) was a very revealing experience. In order to understand progress in science, Hoffmann introduces the idea of “transgression of categorization” and claims that it approximates to Feyerabend’s epistemological anarchism. This was quite intriguing to me as I was then working on Daston and Galison’s thesis of objectivity and its implications for science education. Hoffmann’s thesis of “transgression of categorization” is included in a section entitled “Violating Categories.” In order to understand the underlying idea, I sent the following question to Hoffmann: “Would you agree that this transgression (irrational reasoning) in some sense approximates to what Daston and Galison (2007) have referred to as violating the rules dictated by objectivity?” Hoffmann responded in the affirmative. This response opened a new perspective for me in which I could correlate Feyerabend’s epistemological anarchism, Hoffman’s transgression and Daston and Galison’s changing nature of objectivity. In other words, violating rules and categories can lead to proliferation of theories and hence the evolving nature of objectivity. In a sense, the present book attempts to understand the interrelationships and educational implications of epistemological anarchism, transgression, and objectivity. I am grateful to Calvin Kalman (Concordia University, Canada) for having critically read the preliminary versions of Chaps. 1, 2, 3, 4, 7, and 8. His remarks and criticisms with respect to various subjects (especially counterinduction) were most helpful. My thanks to David Geelan (Griffith University, Australia) for having critically read the preliminary versions of Chaps. 2 and 3. His comments and criticisms helped to improve the style of the manuscript and various substantive issues. Matteo Motterlini (Universitá Vita Salute del San Raffaele, Milan) provided me a copy of his recent work on Feyerabend that helped me to better understand scientific expertise.

xiii

Contents

1 Introduction: Exploring Epistemological Anarchism������������������������������   1 1.1 Origins of Epistemological Anarchism����������������������������������������������   3 1.1.1 Science Education and Feyerabend����������������������������������������   5 1.2 Chapter Outlines ��������������������������������������������������������������������������������   7 2 Epistemological Anarchism and How Science Works������������������������������  23 2.1 Introduction����������������������������������������������������������������������������������������  23 2.2 Feyerabend’s Epistemological Anarchism������������������������������������������  24 2.3 Feyerabend Versus Popper and Lakatos����������������������������������������������  26 2.4 Was Lakatos an Epistemological Anarchist?��������������������������������������  27 2.5 Feyerabend and Scientific Expertise ��������������������������������������������������  28 2.6 Feyerabend Versus Galileo and Copernicus����������������������������������������  30 2.7 Feyerabend and Recent Philosophy of Science����������������������������������  32 2.8 Feyerabend and Perspectivism������������������������������������������������������������  35 2.9 Feyerabend and Feminist Epistemology��������������������������������������������  36 2.10 Feyerabend and the Practice of Science����������������������������������������������  37 3 Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal Science & Education (Springer)������������������������������������������������������������������������������������������������������  39 3.1 Method������������������������������������������������������������������������������������������������  39 3.1.1 Grounded Theory��������������������������������������������������������������������  40 3.1.2 Classification of Articles��������������������������������������������������������  40 3.2 Results and Discussion ����������������������������������������������������������������������  41 3.2.1 Acid-Base Equilibria��������������������������������������������������������������  42 3.2.2 Anarchistic Methodology�������������������������������������������������������  45 3.2.3 Constructivism������������������������������������������������������������������������  46 3.2.4 Critical Thinking��������������������������������������������������������������������  48 3.2.5 Diversity of Methods��������������������������������������������������������������  49 3.2.6 Enlightenment ������������������������������������������������������������������������  50 3.2.7 Evolution, Knowledge and Belief (to Give Meaning to Life)������������������������������������������������������������������������������������  50 xv

xvi

Contents

3.2.8 Gender������������������������������������������������������������������������������������  55 3.2.9 Incommensurability����������������������������������������������������������������  55 3.2.10 Nature of Science��������������������������������������������������������������������  57 3.2.11 Newtonian Method������������������������������������������������������������������  57 3.2.12 Normal Science, Dogmatism and Science Education������������  59 3.2.13 Polanyi’s Tacit Knowledge�����������������������������������������������������  61 3.2.14 Science and Religion��������������������������������������������������������������  62 3.2.15 Scientific Expertise and Galileo����������������������������������������������  63 3.2.16 Scientific Method��������������������������������������������������������������������  65 3.2.17 Situated Learning��������������������������������������������������������������������  68 4 Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal of Research in Science Teaching (Wiley-­Blackwell)������������������������������������������������������������������������  71 4.1 Method������������������������������������������������������������������������������������������������  71 4.2 Results and Discussion ����������������������������������������������������������������������  72 4.2.1 Alternative Literary Forms������������������������������������������������������  72 4.2.2 Creativity��������������������������������������������������������������������������������  73 4.2.3 Evaluation ������������������������������������������������������������������������������  73 4.2.4 Feminism��������������������������������������������������������������������������������  74 4.2.5 Nature of Science��������������������������������������������������������������������  74 4.2.6 Proliferation of Theories ��������������������������������������������������������  75 4.2.7 Scientific Method��������������������������������������������������������������������  76 4.2.8 Teacher Demonstrations����������������������������������������������������������  77 4.2.9 Worldviews ����������������������������������������������������������������������������  78 5 Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal Interchange (Springer)������������������  81 5.1 Method������������������������������������������������������������������������������������������������  81 5.2 Results and Discussion ����������������������������������������������������������������������  82 5.2.1 African and Modern Medicine������������������������������������������������  82 5.2.2 Alternative Approaches to Growth of Knowledge������������������  83 5.2.3 Constructive Alternativism�����������������������������������������������������  85 5.2.4 Diversity of Rival Theories ����������������������������������������������������  86 5.2.5 Genius in Science��������������������������������������������������������������������  87 5.2.6 History of Science������������������������������������������������������������������  88 5.2.7 Objectivity Versus Subjectivity����������������������������������������������  89 5.2.8 Presuppositions of Science Teachers��������������������������������������  90 5.2.9 Rationalism ����������������������������������������������������������������������������  94 5.2.10 Scientific Method��������������������������������������������������������������������  95 6 Understanding Epistemological Anarchism (Feyerabend) in Research Reported in Reference Work��������������������������������������������������  97 6.1 Method������������������������������������������������������������������������������������������������  97 6.2 Results and Discussion ����������������������������������������������������������������������  98 6.2.1 Historical-Investigative Approach to Science ������������������������  98

Contents

xvii

6.2.2 Kuhn and Normal Science������������������������������������������������������  99 6.2.3 Nature of Science������������������������������������������������������������������  100 6.2.4 Postmodernism����������������������������������������������������������������������  102 6.2.5 School Science Curriculum��������������������������������������������������  104 6.2.6 Science as Cultural Tyranny�������������������������������������������������  106 7 Feyerabend’s Counterinduction and Science Textbooks������������������������  109 7.1 Brownian Motion������������������������������������������������������������������������������  110 7.2 Kinetic Theory of Gases ������������������������������������������������������������������  111 7.3 Michelson-Morley Experiment ��������������������������������������������������������  114 7.4 The Oil-Drop Experiment ����������������������������������������������������������������  116 7.5 Alpha Particle Scattering Experiment����������������������������������������������  123 7.6 Bohr’s Incorporation of “quantum of action” to Classical Electrodynamics��������������������������������������������������������������������������������  125 7.7 Photoelectric Effect��������������������������������������������������������������������������  128 7.8 Wave-Particle Duality ����������������������������������������������������������������������  131 7.9 Mendeleev’s Periodic Table of Chemical Elements��������������������������  134 7.10 Lewis’s Postulation of the Covalent Bond����������������������������������������  144 7.11 Discovery of the Planet Neptune������������������������������������������������������  147 7.12 Discovery of the Elementary Particle Neutrino��������������������������������  149 7.13 Discovery of the Tau Lepton������������������������������������������������������������  150 8 Conclusion: Feyerabend and Challenges of the Twenty-First Century ������������������������������������������������������������������������������������������������������  155 8.1 Feyerabend’s Hyperbolic Flourishes������������������������������������������������  155 8.2 Feyerabend’s Epistemological Anarchism����������������������������������������  156 8.2.1 Counterinduction (Accepting Unsupported Hypotheses) ����  157 8.2.2 Current View of a Science May Soon Be Voted Out  of Office��������������������������������������������������������������������������������  160 8.2.3 Does Science Always Provide the One “Correct” Model (Theory)��������������������������������������������������������������������  162 8.2.4 Epistemological Anarchism: Lifting the Lid Off  the Pandora’s Box? ��������������������������������������������������������������  163 8.2.5 History of a Science Becomes an Inseparable Part of the Science Itself��������������������������������������������������������������  163 8.2.6 Inferring Objectivity from Empirical Approaches����������������  165 8.2.7 Methodological Pluralism: Diversity of Methods����������������  166 8.2.8 Nature of Science������������������������������������������������������������������  167 8.2.9 Role of Genius in Science����������������������������������������������������  168 8.2.10 Scientific Expertise Needs a Critical Appraisal��������������������  168 8.2.11 Scientific Method: Stockpiling and Ordering of Observations ��������������������������������������������������������������������  168 8.2.12 Skeletons in the Newtonian Cupboard����������������������������������  169 8.2.13 The New Grew Out of the Old����������������������������������������������  169

xviii

Contents

8.2.14 Unnatural Nature of Science������������������������������������������������  170 8.2.15 Was Feyerabend a Postmodern or Perspectival Realist?����������������������������������������������������������������������������������  171 8.3 Educational Implications������������������������������������������������������������������  172 Appendices��������������������������������������������������������������������������������������������������������  175 References ��������������������������������������������������������������������������������������������������������  199 Name Index������������������������������������������������������������������������������������������������������  215 Subject Index����������������������������������������������������������������������������������������������������  221

Chapter 1

Introduction: Exploring Epistemological Anarchism

Paul Karl Feyerabend has been considered as the worst enemy of science, an enfant terrible, and is generally considered to be the maximum exponent of epistemological anarchism (Theocaris & Psimopoulos, 1987). However, what it actually entails is difficult to understand and its meaning elusive. For example, in his famous Against Method (Feyerabend, 1975a) he provides counterexamples against standard methodological advice, such as: Do not allow theories in contradiction with observations, only allow theories that can be potentially falsified, generalize inductively from facts, maintain a clear difference and distinction between theories and facts, and metaphysical ideas need not have a central role in scientific theorizing. In other words in actual scientific practice such advice may not be followed, and this was Feyerabend’s major concern. No wonder, the New York Times headlined its obituary of Feyerabend as “Anti-science philosopher” (published March 8, 1994) and then continued to attribute the following: Dr. Feyerabend held that the rationality of science did not really exist and that the special status and prestige of scientists are based on their own claims to objective truth … Dr. Feyerabend was one of the most radical challengers to the long-accepted notion that science is rational and progressive. If there was progress in science, he insisted, it was because scientists broke every principle in the rationalists’ rule book and adopted the principle that “anything goes” (Italics added).

Interestingly, even philosophers of science hold similar views with respect to Feyerabend (e.g., Achinstein, 2004; Hattiangadi, 1977; Preston, 1997). The reference to objective truth is particularly important in this context. Before the publication of Einstein’s theory of relativity starting in 1905 and the development of quantum mechanics, many physicists also considered classical Newtonian mechanics to be objectively true, and this perspective started to change in the early twentieth century (cf. Giere, 2006a, 2006b). Similarly, it is generally considered that Feyerabend espoused guidelines such as: Scientists should not be constrained by the rules of the methodologist, fruitful violations of rules cannot be anticipated and legislated for in advance, and new practices will alter methodologies (Gower, 1997, © Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_1

1

2

1  Introduction: Exploring Epistemological Anarchism

pp. 246–247). The last advice is particularly important as it provides insight with respect to how science is actually done or practiced. Such descriptions of Feyerabend’s contributions (as an anti-science philosopher) to history and philosophy of science (HPS) are also fairly common in both science and science education. In contrast, Kadvany (2001) considers that such descriptions are not only false but also ignore the fact that Feyerabend had a great love and admiration for classical science (p. 151). Furthermore, Feyerabend was completely allied in his criticism of “normal science” (Kuhn, 1970) and its dangers with K. Popper and I. Lakatos. This is important as both Popper and Lakatos are generally considered to espouse rationalism in history and philosophy of science (HPS). Kadvany (2001) presents the following overall perspective: “Indeed, Feyerabend provided something of a critical doctrine of science’s role in society, arguing for a kind of libertarianism of ideas, adapting John Stuart Mill’s ideas on freedom from On Liberty to the proliferation of theories in science and knowledge generally” (p. 151). In his lectures on scientific method delivered in 1973 at the London School of Economics, Lakatos traces the origins of epistemological anarchism to the ancient Greek skeptic Pyrrho (360–270 BC) and considered Paul Feyerabend to be his supporter / follower: Note that epistemological anarchism has nothing to do with Mao’s “let a hundred flowers bloom,” since ‘flowers’ is, of course, a normative term. Rather, flowers and weeds may bloom together—there is no demarcation line between them. This presents a very important problem. Feyerabend has absolutely no intention of imposing a subjective distinction between flowers and weeds on anybody. Any system of beliefs—including Popper’s philosophy of science—is free to grow and influence any other, but none can claim epistemological superiority (Reproduced in Motterlini, 1999, p. 25).

Furthermore, according to Lakatos, epistemological anarchism claims that any belief is as good as any other and that science is a set of beliefs on a par with Azande beliefs (people from central Africa). This led Lakatos to suggest that Feyerabend espouses the policy of “let all weeds grow” and that intellectual influence is directly proportional to the vocal energy, the faith and the propaganda skills of competing groups. In a letter written to Lakatos, dated February 1973, Feyerabend sent his “Theses on Anarchism” and clarified how epistemological anarchism differs from skepticism: While the sceptic either regards every view as equally good, or equally bad, or desists from making such judgments altogether, the epistemological anarchist has no compunction in defending the most trite, or the most outrageous statement. While the political anarchist wants to remove a certain form of life, the epistemological anarchist may want to defend it, for he has no everlasting loyalty to, and no everlasting aversion against, any institution and any ideology (Reproduced in Motterlini, 1999, p. 114. A revised version of this thesis was published as Feyerabend, 1975b).

In the same letter Feyerabend then continued and recognized that some of these aspects of epistemological anarchism may lead some epistemologists to consider them as unscrupulous and opportunist, and finally concluded: “[The epistemological

1.1  Origins of Epistemological Anarchism

3

anarchist] … will try to convince his audience that the only universal rule that can safely be in agreement with the moves the scientist must make to advance his subject is anything goes” (Theses on anarchism, reproduced in Motterlini, 1999, p. 116, italics in the original). Anything goes with respect to scientific practice is perhaps one of the most controversial aspects of Feyerabend’s philosophy of science. In a letter written to Lakatos in March 1973, Feyerabend clarifies that anything goes does not mean skepticism but rather may even include law and order, argument and irrationalism (Reproduced in Motterlini, 1999, p. 324).

1.1  Origins of Epistemological Anarchism In a letter written to I. Lakatos, dated January 20, 1972, Feyerabend refers to the origin of his ideas with respect to “epistemological anarchism”. He recounts that in 1965 while discussing the example of Brownian motion with a colleague it occurred to him that even such a highly confirmed theory could have an alternative. This would require the invention of new methods rather than adapting to “reason”, and besides that he also acknowledged the role played by his Wittgensteinian upbringing. It is plausible to suggest that the idea of “transgression” of method was embedded in this episode. In the same letter, Feyerabend also referred to the scientific method in the following terms: “… the pleasant surprise I got when Sir Karl [Popper], then Prof. P., started his lectures on scientific method (in 1952) by saying: ‘I am Professor of Scientific Method; but there is no scientific method …’ which I liked …” (Reproduced in Motterlini, 1999, p.  272). The idea of transgression is much more clearly expressed in Against Method: “…there is not a single rule, however plausible, and however firmly grounded in epistemology, that is not violated at some time or other … Such violations are not accidental events, they are not results of insufficient knowledge or of inattention which might have been avoided. On the contrary, we see that they are necessary for progress” (Feyerabend, 1975a, p. 23). In another letter written to I. Lakatos dated 25 July 1969, Feyerabend stated, “… the ‘objectivity’ of science is just moonshine?” (Reproduced in Motterlini, 1999, p.  169). Interestingly, Preston (1997) considers that there is a strong connection between the ideas of Feyeraband and the work of Michael Polanyi. In a conversation with Feyerabend at the University of Auckland in 1974, Robert Nola asked him what philosophical stance he would adopt if everyone around him was an epistemological anarchist like himself. Given Feyerabend’s penchant for rhetoric and perhaps half in jest, he responded: “… under such conditions rationality might well have to be supported, supposing it were desirable that science survive” (Reproduced in Nola, 1996, p.  471). In his autobiography, Killing Time (Feyerabend, 1995), he clarified further, “I never ‘denigrated reason’, whatever that is, only some petrified and tyrannical versions of it” (p. 134).

4

1  Introduction: Exploring Epistemological Anarchism

At this stage it is interesting to consider the relationship between the development of Lakatos’s philosophy of science and epistemological anarachism. In a letter written to Lakatos, dated 20th March 1973, Feyerabend stated: Combining the common sense standards of scientists [scientific practice] with the methodology of scientific research programmes [MSRP], Lakatos utilises the intuitive plausibility of the former to support the latter: a splendid Trojan horse that can be used to smuggle real, honest (a word so dear to Lakatos) anarchism into the minds of our most dedicated rationalists (Reproduced in Motterlini, 2002a, 2002b, p. 36, italics added).

MSRP represents Lakatos’s (1970) major contribution to history and philosophy of science (for details see Chap. 2). According to Bird (2010), “Indeed Feyerabend regards Lakatos’s view as being closet anarchism disguised as methodological rationalism” (p. 76). It is plausible to suggest that combining common sense standards of scientists with the standards based on the rationalist MSRP of Lakatos, facilitates an approximation between scientific practice and standards espoused by Lakatos and hence the incorporation of some elements of epistemological anarchism. According to Motterlini (2002a): Whether or not Lakatos has really gone a long way towards epistemological anarchism is a problem we can better solve in the more general framework of Lakatos’s late struggle against the “modern intellectuals’ betrayal of reason” (p. 37).

In his “Intellectuals’ Betrayal of Reason” Lakatos emphasized the importance of the Newtonian research program in the following terms: “Thus, since 1687 [year Newton published his Principia] the whole history of the philosophy of science— and indeed of rationality—has revolved around a single central issues: what are the epistemological merits, if any, of Newtonian physics?” (Reproduced in Motterlini, 1999, p. 397). Newtonian physics formed an important part of the Lakatosian research agenda. In order to respond to Feyerabend’s Against Method, Lakatos intended to write a book titled: “The changing logic of scientific discovery”, which was never finished. However, one chapter of this book (Newton’s effect on scientific standards, Lakatos, 1963–1964/1978) was published, which provides some ground for comparing the evolving nature of the Feyerabend-Lakatos debate with respect to standards for appraising scientific theories and its relationship to epistemological anarchism. According to Lakatos (1978): “In this sense one may say that Newton’s theory of method created modern philosophy of science. But this turn separated science and philosophy of science from 1686 to 1905 [In 1905, Einstein’s theory superseded Newton’s] … Philosophy, unaware of the split between the wonderful Newtonian method as practiced and the mad Newtonian method as professed, tried to clarify the professed method …” (Reproduced in Motterlini, 2002a, p.  40). One such attempt to clarify the professed method was by P. Duhem (1914), who first asked a very didactic question: “Does logic require our hypotheses to be simply experimental laws generalized by induction?” (p.  219), and then responded in very forthright terms: Now, we have recognized that it is impossible to construct a theory by a purely inductive method. Newton and Ampère [André-Marie Ampère, 1775–1836, known for his theory of

1.1  Origins of Epistemological Anarchism

5

electrodynamics] failed in this, and yet these two mathematicians had boasted of allowing nothing in their systems which was not drawn entirely from experiment. Therefore, we shall not be averse to admitting the fundamental bases of our physics postulates not furnished by experiment (Duhem, 1914, p. 219).

This clearly shows the difference between what Newton professed and practiced, which led Duhem (1914) to conclude that the “Newtonian method” although it may appear attractive, was more of a dream. Similarly, according to Kuhn (1977), when Newton enunciated his theory in the late seventeenth century, only his third law could be directly investigated by experiment. Convincing demonstration of the second law had to await the development of Atwood’s machine, almost a century after the appearance of the Principia. Consequently, Newtonian method is more of an idealization or abstraction and does not describe the behavior of actual bodies (cf. Cartwright, 1983; Giere, 1999). Interestingly, it was such considerations that provided a backdrop to Lakatos’s conceptualization of scientific progress and its relationship to epistemological anarchism. This relationship is explored further in Chap. 2.

1.1.1  Science Education and Feyerabend Ron Good (1993), Editor of the Journal of Research in Science Teaching, recommended to science educators the need to avoid the slippery slopes of postmodernism (as represented by Feyerabend and Foucault) and concluded: “To question the objectivity of observation or the truth of scientific knowledge, one does not need to travel to the wispy world of postmodernism. Logical positivism and postmodernism are at the extremes of a long continuum of positions taken by scholars of the nature of science. It is not necessary to carry along the unwanted (and unwarranted) baggage of either logical positivism or postmodernism to place oneself, as did the authors of Science For All Americans, in a more ‘scientifically’ defensible position” (p. 427, italics in the original). Science For All Americans refers to the document published by the National Research Council in 1992. Loving (1997) considers Feyerabend to be a relativist and postmodern philosopher and provides the following background: Although a rationalist and a realist, Popper … questioned acutely the ability of science to verify or confirm. Instead, he promoted the proliferation of bold theories and conjectures and their subsequent downfall and elimination through vigorous testing and refutation. His former student, Feyerabend …, abandoned the rationalist mode of his mentor, along with Hanson … and, of course, Kuhn, further defining and directing the postmodern agenda in science (p. 442).

Furthermore, Loving (1991) considers that Feyerabend has endorsed that, “… the winning theories could not have arrived at their position without conceit, passion, and prejudice” (p. 830). Geelan (2006) has suggested that Feyerabend’s epistemological anarchism provides a more useful, powerful and productive referent for research in education than

6

1  Introduction: Exploring Epistemological Anarchism

do the philosophies of Popper, Kuhn and Lakatos. Most science educators may not agree with this statement and this precisely makes the exploration of Feyerabend’s contributions important. Geelan endorses Feyerabend’s thesis that producing genuinely new knowledge, such as the Galilean revolution in astronomy could not have been possible had the strict rules of Popper, Kuhn and Lakatos been enforced. Based on this perspective Geelan recommends that educational research needs to be characterized by pluralism, disciplined eclecticism, reflexive interaction, and epistemological tolerance for rival and competing perspectives. Finally, Geelan (2006) concluded: “Epistemological anarchy should not, however, be confused with a bland relativism—there will continue to be controversy and passionate disagreement. Feyerabend is a great exponent of stinging polemical writing” (p.  26). Furthermore, this perspective facilitates the possibility of controversies between competing paradigms / programs in educational research leading to robust research experiences. Hodson (2009) has cautioned that a too literal interpretation of Feyerabend’s dictum “anything goes” constitutes a disservice to students’ understanding of science. Actually scientists are very careful with respect to the appropriate method for a particular kind of research, which has to be scrutinized and critically appraised by the scientific community. However, Hodson recognizes that Feyerabend’s free-­ wheeling approach is more suitable for the creative phase of scientific inquiry, often referred to by philosophers of science as the context of discovery. Similarly, Feyerabend’s anarchistic view may even be necessary during the revolutionary phase of scientific inquiry rather than during the conduct of normal science (cf. Kuhn, 1970), which requires a more dispassionate and systematic approach. After having pointed out such caveats, Hodson (2009) finally concluded: In other words, doing science is an untidy, unpredictable activity that requires each scientist to devise her or his own course of action. In that sense, science has no one method, no set of rules or sequence of steps that can, and should, be applied in all situations—as Paul Feyerabend (1975a) so eloquently argued. His case against methodologists like Karl Popper (1959) and Imre Lakatos (1978) is that their arguments for a fixed method of science rest on too naïve a view of what is involved in conducting investigations and building theories, and that the only principle that is always applicable in conducting the complex and sometimes chaotic business of science is the principle Anything Goes (p. 198, italics in the original).

The crux of Hodson’s argument rests on the premise that “doing” science is different from “teaching” science. Doing science generally refers to cutting-edge research when a scientist has no clear strategy or method, stakes are high, controversy is in full sway and it is difficult to foresee the final results. It is under such circumstances that Feyerabend’s “Anything goes” makes sense and can provide plurality of methods so essential for the scientific endeavor. At this stage it is important to clarify that “doing” and “teaching” science need not be conceived as two antagonistic activities. During the last 25 years the science education community has been actively engaged in infusing school science with the history of science in order to provide a semblance of how science is done (cf. Matthews, 2014a, 2014b, 2015). Similarly, Bazghandi and Hamrah (2011, p.  972) have also recognized the importance of Feyerabend’s methodological pluralism for science education.

1.2  Chapter Outlines

7

Most philosophers of science and science educators have characterized Feyerabend as a relativist, postmodern philosopher of science who espoused anything goes, based on epistemological anarchism. In order to provide a better ­understanding of Feyerabend’s thinking and its importance for science education, this book has the following objectives: 1. Explore the evolving nature of Feyerabend’s philosophy of science, especially on issues such as relativism and postmodernism. A review of the recent literature on history and philosophy of science shows that he was more of a perspectival realist and recommended methodological pluralism. 2. Based on this exploration of Feyerabend’s philosophy of science elaborate criteria for evaluating research (epistemological anarchism) in science education, within a history and philosophy of science framework. 3. Based on these criteria, evaluate research published in the following sources: Science & Education (Springer journal), Journal of Research in Science Teaching (Wiley-Blackwell journal), Interchange (Springer journal), and International Handbook of Research in History, Philosophy and Science Teaching (2014, Springer). These sources cover a period of over 30 years. 4. Evaluate science textbooks based on Feyerabend’s principle of counterinduction (accepting unsupported hypotheses) in the history of science, and explore teaching strategies. The rationale behind these four objectives is the importance of the changing nature of Feyerabend’s views on controversial issues and it is significant to note that many of these are now consonant with current philosophy of science. Furthermore, evaluation of science textbooks in the context of counterinduction adds a new dimension to our understanding of the history of science.

1.2  Chapter Outlines The objective of the chapter outlines is to provide the reader an overview of the different chapters by including some salient features. Introduction: Exploring Epistemological Anarchism (Chap. 1). Many philosophers of science consider Feyerabend to be anti-science and even perhaps the worst enemy of science. In his epistemological anarchism Feyerabend was particularly concerned with the practice of science, which generally refers to cutting-edge research when a scientist has no clear strategy or method, stakes are high, controversy is in full sway and it is difficult to foresee the final results. It is under such circumstances that Feyerabend’s “Anything goes” makes more sense and can provide a plurality of methods so essential for the scientific endeavor. Actually, based on Mill’s On Liberty he espoused proliferation of theories in science. Feyerabend has often been criticized for questioning the existence of objectively true theories. In this context how shall we explain to our students that before 1905, many physicists also thought that Newtonian classical mechanics was objectively true. In contrast to

8

1  Introduction: Exploring Epistemological Anarchism

skepticism, Feyerabend emphasized that epistemological anarchism does not have an everlasting loyalty towards any ideology. The possibility of “transgression” of method is embedded in the idea that even a well confirmed theory could have an alternative. Given the methodological debate, and the continuous exchange of ideas between Feyerabend and Lakatos, it is plausible to suggest that even Lakatos, although considered a rationalist, included in his philosophy of science some elements of epistemological anarchism. Lakatos was particularly influenced by the development of the Newtonian method and its appraisal in the history of science. Just like the philosophers of science most science educators also consider Feyerabend to be a relativist and postmodern. For science education it is important to note that arguments for a fixed method of science are based on a very simplistic comprehension of the scientific enterprise and this precisely led Feyerabend to postulate proliferation of methods. This book is based on the premise that a historical reconstruction facilitates a better understanding of the changing nature of Feyerabend’s views on controversial issues and many of these are now consonant with current philosophy of science. A major objective of this book is to explore important aspects of Feyerabend’s philosophy of science in the context of different sources (journals, handbook and science textbooks) and draw implications for science education. Epistemological Anarchism and How Science Works (Chap. 2). Theoretical framework of studies reported in this book is based on the following aspects of Feyerabend’s philosophy of science: (a) It is desirable to introduce hypotheses which are inconsistent with highly confirmed theories; (b) The idea of a method based on unchanging principles cannot be sustained as eventually all rules are violated; (c) A pluralistic methodology leads to proliferation of different points of view that are essential for science; (d) There is a need to admit unsupported hypotheses (counterinduction); (e) Adhoc hypotheses play an important role in science; (f) There is no such thing as the scientific method; and (g) Theory-laden nature of observations. Feyerabend was critical of both Popper and Lakatos for advocating standards based on a rational understanding of progress in science. According to Feyerabend, Lakatos oscillated between two extremes: On the one hand he endorsed methodological standards and on the other made his standards so flexible that they amounted to “anything goes” and hence Lakatos may even be considered an epistemological anarchist. Feyerabend was skeptical of scientific expertise as he considered that it should be subject to critical appraisal. Despite the differences between the position of Feyerabend and recent attempts at understanding scientific expertise (Collins and Evans), Sorgner considers that there are some common elements between the two positions, especially Feyerabend’s demand that science and society be separated. According to Feyerabend, Galileo was faced with a dilemma: accept the well-supported Ptolemaic hypothesis—a rational decision, or accept the ill-­ supported Copernican hypothesis—a non-rational decision. Thus, according to Feyerabend in order to confound the Ptolemaists, Galileo was using counterinduction, by using the relatively unreliable telescope to support the anomaly-ridden Copernicanism. Throughout his intellectual career Feyerabend was considered as a relativist and postmodernist, and philosophers of science questioned his eccentricity, eclecticism and especially his infamous “Anything goes”. Recent research in

1.2  Chapter Outlines

9

philosophy of science, however, has shown that Giere’s (2006a, 2006b) perspectival realism, Dupre’s (1993) promiscuous realism and Cartwright’s (1999) local realism in a “dappled world” all fit well with Feyerabend’s metaphysical views. Interestingly, many of the themes espoused by Feyerabend that were considered to be anti-science and irrationalist (summarized as epistemological anarchism), are now considered to form part of mainstream philosophy of science, such as: science is pluralistic, disunified, value-laden and bound up with social and political concerns. According to Giere’s perspectival realism no scientific theory can claim to present a complete and literally a true picture of the world itself, and this in my opinion comes quite close to Feyerabend’s principle of counterinduction. Giere himself endorsed the position that Feyerabend was a perspectival realist. Feminist epistemology (Haraway, Harding, Keller, and Longino) has also generally endorsed Feyerabend’s philosophy of science. How science is practiced, namely challenging methodological rules forms an important part of Feyerabend’s epistemological anarchism. This aspect of Feyerabend’s philosophy of science has been recognized by both philosophers of science (e.g., Kidd) and practicing scientists (e.g., Roald Hoffmann, who referred to it as “violating categories”). Understanding of Epistemological Anarchism (Feyerabend) in Research Reported in the Journal Science & Education (Springer) (Chap. 3). Based on a website search with the keywords, “Epistemological anarchism and Feyerabend” 78 articles in the 26-year period (1992–2017) referred to some aspect of Feyerabend’s philosophy of science and were classified according to the following criteria: Level I, traditional understanding of Feyerabend as an anti-realist, and postmodern philosopher of science who espoused “anything goes”; Level II, recognition of Feyerabend’s criticism of a unitary/single scientific method and other contributions; Level III, criticism and controversy with respect to the social responsibility of the scientist, scientism and the need to go beyond the scientific method; Level IV, different perspectives on a subject (such as anything goes) reflect also diversity rather than a Unitarian view of science; and Level V, breaking rules in scientific practice means transgression of categorization which leads to a plurality of perspectives. Of the 78 articles evaluated 7 were classified in Level V. Based on the treatment of the subject by the authors 17 categories were developed to report and discuss the results. Following are some of the findings. Feyerabend’s “anti-science” stance in most cases could be attributed to his rhetorical excesses. For science education, on the one hand we have to criticize the ‘dogmatic view of science’ (scientific method), and then go beyond and accept Feyerabend’s advice that scientific work involves a variety (plurality) of strategies (methods). Hence, both science and teaching science require the understanding that scientific work is vague, uncertain, intuitive and tacit. The self-corrective nature of science depends more on diversity and plurality in a discipline, rather than the scientific method. The legacy of logical positivism requires that the topic of acid-base equilibria (models of Arrhenius, Brønsted-Lowry and Lewis) should preferably facilitate students’ understanding that of the three models only “one” is “correct” or “true”. Feyerabend is critical of not only logical positivism but even some of the recent philosophers of science, as this precisely leads to epistemological pluralism (viz., more than one “correct” model).

10

1  Introduction: Exploring Epistemological Anarchism

Feyerabend’s anarchistic methodology has one simple rule: there is no rule and the methodological strategy is not decided a priori. In the case of constructivism ­opposing theories do not supplant each other but are rather necessary. Furthermore, the search for a single best approach is self-defeating. Anything goes means that one should feel free to try any and all approaches that may seem promising, but then with the clear understanding that many such attempts may not in fact succeed. Feyerabend argued that Galileo arrived at the modern theory of inertia by a critical examination of the tower experiment in the light of two alternative frameworks, namely that of Ptolemy and Copernicus. Despite his criticism of Galileo, in some of his writings, Feyerabend also supported such an approach in his philosophy of science. For example, in his reconstruction of the role for experiment, Feyerabend endorsed Galileo’s method namely an example of counterinduction. In teaching controversial issues (e.g., intelligent design, ID and evolutionary theory) the role played by culture as envisaged by Feyerabend is important. Debate between ID and evolution would provide the setting that approximates to what Feyerabend considered as counterinduction. It seems that although students may understand evolution, they generally do not believe in it. Leaving aside Feyerabend’s rhetoric, it is no surprise that if creationism and evolution were given equal time and opportunities, those teaching the science curriculum (even in the industrialized countries) would be hard pressed to convince students of the ‘scientific’ nature of evolutionary theory. Teaching evolution as a series of facts may lead to dogmatic scientism, whereas Feyerabend would instead recommend a contradictory approach based on alternative ways of thinking. This does not necessarily imply that Feyerabend is recommending creationism. It is important not to lose sight of the fact that science is basically not only counter intuitive but “unnatural” and hence the difficulties involved in learning it. Furthermore, there is some concern that if presented only one perspective, students may stop thinking. Based on the concept of incommensurability, Feyerabend drew attention to the difficulties involved in comparing theories with different frameworks. To overcome this difficulty one strategy could be the introduction of both theories, for example mechanics and thermodynamics—this coincides with Feyerabend’s advice for allowing diversity. Again, given the partial overlap between quantum and classical physics, that is in some limit the new theory can reproduce the results of the theory it intended to replace, leads to understanding: the new grew out of the old. Furthermore, scientific theories are neither “true” nor “false” but rather continue to evolve and progress. Similarly, alternative interpretations of historical events is recognized as an important part of philosophy of science and hence for understanding nature of science. Did Newton formulate his famous four rules before or after having postulated his law of gravitation? Most philosophers of science consider that Newton could not have formulated his rules a priori. In other words, strictly speaking Newtonian method was not an inductive generalization and has been the subject of considerable controversy in the history and philosophy of science literature. Falsification has been endorsed by Popper as an important part of the scientific enterprise and that inevitably leads to recognizing observations as crucial for testing a theory (induction). On the other hand, Feyerabend, (along with Lakatos and Kuhn) was particularly critical of the role of

1.2  Chapter Outlines

11

falsification in scientific progress. According to Kuhn, science education does and should distort the history of science in order to inculcate the dominant paradigm of the day and textbooks are designed precisely to perpetuate ‘normal science.’ Replacing normal science with the dynamics of science-in-the-making is difficult due to the dogmatic and authoritarian nature of science education and textbooks. Teaching normal science leads to memorization of science content with little understanding and its elimination could facilitate the inclusion of controversies, namely unfolding of the historical episodes based on rival interpretations—thus providing students a glimpse of what Feyerabend referred to as how science really works, namely proliferation of theories. Scientists often proceed intuitively based on their personal understanding and tacit knowledge of how to do science. It is plausible to suggest that this shows a proximity between Polanyi’s tacit knowledge and Feyerabend’s philosophy of science. The controversy between Lavoisier and Priestley to determine the composition of air is well known in the history of science. Priestley had difficulty in replicating some of Lavoisier’s experimental results, who was guided by the law of conservation of matter. Such controversies can help us to understand Feyerabend’s advice that, “history of a science becomes an inseparable part of the science itself”. Furthermore, in order to understand the controversy it is important that students become knowledgeable in both oxygen and phlogiston theories, facilitating not only a plurality of theories but also counterinduction, as suggested by Feyerabend. Situated learning is considered as not appropriate as a guiding principle for science education, as a child becomes an apprentice in a particular subject/field at an early age. In this context, anything goes means a teacher can adapt according to the needs, aspirations and motivation of the students. Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal of Research in Science Teaching (Wiley-Blackwell) (Chap. 4). Based on a website search with the keywords, “Epistemological anarchism and Feyerabend” 21 articles in the 37-year period (1970–2017) referred to some aspect of Feyerabend’s philosophy of science and were classified according to the following criteria: Level I, Level II, Level III, Level IV and Level V (same as in Chap. 3). Of the 21 articles evaluated only one was classified in Level V. Based on the treatment of the subject by the authors 9 categories were developed to report and discuss the results. Following are some of the findings. Based on Latour’s Actor Network Theory, researchers working with alternative literary forms have recommended that in order to understand science it is important to study how scientists arrive at conclusions and not the products of scientific inquiry. Furthermore, it is essential to investigate how controversies are resolved. A basic premise of such an approach is the “dialogue”, which has been used throughout human history—from Plato to Galileo to Feyerabend. Although many aspects of Feyerabend’s philosophy of science are controversial, his recognition that science is a creative endeavor continues to enjoy wide support. The use of empirical data in scientific research is a valid method. However, history of science shows that to infer objectivity from entirely empirical approaches is questionable. The masculine worldview which considers science as being objective, rational, individualistic, unemotional, and value-free needs a critical appraisal. Most philosophers of science would agree that just like scientific

12

1  Introduction: Exploring Epistemological Anarchism

knowledge itself, nature of science is tentative. The rich complexity of science can be illustrated with its practice and history. Feyerabend’s concept of proliferation of theories means that students can learn to work with many theories as working hypotheses in a given domain of inquiry (e.g., evolutionary theory). In other words, having more theories to work with, there is less chance that one will persist with just one theory, which may need changes. A major concern of Feyerabend was to understand how science is actually practiced and this in part explains why his interpretations differed from those of other philosophers of science. Comparing the philosophical work of Popper and Feyerabend is instructive for science educators, as despite the common features, the scientific enterprise goes far beyond the Popperian refutability of theories as it involves uncertainty and critical appraisal by peers. The traditional lecture, lab, and demo approach to science education, systematically obscures these elements of the scientific enterprise. Teacher demonstrations form an important part of school science, and still there problematic nature is ignored, namely students observe what they already know. Thus a student who does not know the relevant scientific principle will not benefit from the teacher demonstration. It is important for science educators to understand that logical positivism and postmodernism are two extreme positions in philosophy of science and that in his later work Feyerabend explicitly renounced postmodernism. Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal Interchange (Chap. 5). Based on a website search with the keywords, “Epistemological anarchism and Feyerabend” 15 articles, published since 1982, referred to some aspect of Feyerabend’s philosophy of science and were classified according to the following criteria: Level I, Level II, Level III, Level IV and Level V (same as in Chap. 3). Of the 15 articles evaluated, two were classified in Level V. Based on the treatment of the subject by the authors 10 categories were developed to report and discuss the results. Following are some of the findings. In most African and developing countries women’s education is neglected and even if they overcome the odds and graduate they are discriminated against in their professional careers. One study interviewed ten women from Cameroon in order to find out how they had managed to become scientists and science educators. Responses of the participants showed the difficulties involved in integrating Western medical science and traditional medicine of Cameroon. Providing in-service teachers an opportunity to become familiar with the controversial nature of progress in science (growth of knowledge) is an important objective of most innovative educational systems. In one study in-service teachers were asked to select a topic from their area of work and explain the growth of knowledge through any of the following conceptualizations: Kuhn, Lakatos, Campbell and Erickson. Differences between the four conceptualizations are rather subtle and can even be interchangeable. For example, a Kuhnian conceptualization is characterized by abrupt changes that leads to the displacement of one paradigm by another. A Lakatosian conceptualization is characterized by competition between rival research programs instead of paradigms. A conceptualization based on the work of Erickson is characterized by the coexistence of the old and the new paradigms. It is difficult to replace an old paradigm by falsification. A conceptualization based on the work of Campbell is characterized by

1.2  Chapter Outlines

13

rivalry between different hypotheses that appear plausible for explaining a phenomenon. It is important to note that these in-service teachers were neither historians of science nor philosophers of science, and had very little previous experience in understanding growth of knowledge critically. However, their teaching experience and a critical perspective helped most of them to understand the nuanced differences between the different conceptualizations. For example, Kuhn emphasizes displacements, Lakatos emphasizes rivalries and Erickson emphasizes coexistence. Interestingly, participants also provided competing interpretations of the same developments in their area of work. Actually, according to Feyerabend, this is how science works—alternative approaches to understanding growth of knowledge. Just as the scientist designs experiments around rival hypotheses, students can be encouraged to do the same. A parallel can be drawn between G. Kelly and Feyerabend as both considered that any event is open to as many reconstructions of it as our imagination will allow (hence Feyerabend’s advice “anything goes”). Again, both deplored the traditional teaching methods based upon the cultural transmission approach that emphasizes the student’s role as the passive receiver rather than the active participant. Precisely, Feyerabend argued cogently that science education should strengthen the minds of the young against any easy acceptance of comprehensive views and to be receptive to alternative counter-suggestive views, namely counterinduction. At the beginning of the twentieth century, Einstein, Perrin, and others offered an explanation of Brownian motion based on the kinetic theory—a rival of classical thermodynamics. The success of this explanation for the kinetic theory put classical thermodynamics in a rival position. Thus Brownian motion became relevant evidence for classical thermodynamics, only because a rival had explained it, which shows the role of diversity of rival theories, as espoused by Feyerabend. In the twentieth century, there has been a gradual encroachment of the nineteenth century conception of the artist as a genius into interpretations of science as an activity. This role of the genius in science has been endorsed by Koestler, Polanyi and to a lesser extent by Feyerabend (considered as an egalitarian romantic). Brush has argued that traditional science teaching looks for objective facts instead of recognizing that scientific research does not provide immutable truths but rather working hypotheses useful for future research. Furthermore, although objective facts are valid and useful knowledge they are subject to change and revision namely the tentative nature of scientific knowledge. Feyerabend’s framework of counterinduction provides alternative working hypotheses that can help to strike a balance between the two extremes. Such considerations can make science education more meaningful for both students and teachers. According to Winchester, history of science is important not only for science education but for science itself. For example, Galileo and Darwin on the one hand, and even Newton and Einstein were fully aware of the work of their predecessors and the controversies surrounding earlier work in their respective fields of expertise. Daston and Galison have recognized the intricate relationship between objectivity and subjectivity (both define each other) and that there is no objectivity without subjectivity to suppress and vice versa. More recently, Niaz has explored the evolving nature of objectivity in the history of science and its implications for science education. In most parts of the

14

1  Introduction: Exploring Epistemological Anarchism

world science teachers have presuppositions of the nature of science that can be represented by Karl Pearson’s The Grammar of Science. Based on the work of Popper, Kuhn and Feyerabend these teachers were introduced by Winchester to postpositivist philosophy of science. After teaching such courses for 20  years, Winchester considered himself to be successful. Despite this success, he also became skeptical as the teachers may have taken home the lesson that there was no “truth” in science and in the process may have “killed” science itself. According to Giere’s perspectival realism, scientific theories are neither true nor false. In the light of this discussion, it is plausible to suggest that Winchester’s (1993) qualms with respect to “truth” in science and history and philosophy of science are not warranted or at least misplaced. Feyerabend is famous for his Farewell to Reason. However, he clarified that he was opposed only to rationalism that was rigid, pompous and harsh. The positivist myth has led some scholars to believe that conformity with the rules of the scientific method will yield scientific knowledge that is “true” and perhaps even beyond criticism. Should inductive procedures be taught to students as a significant part of the method that scientists use to discover and test hypotheses about the observable world? This is an important question for science educators. Interestingly, Dewey responded affirmatively to this question and Popper in the negative. However, if we accept the liberal idea of toleration of diversity as a guiding principle, then it would be possible to incorporate Feyerabend’s methodological pluralism in school science to study the ideas of Dewey, Popper, or anyone else who might help to advance learning. Understanding Epistemological Anarchism (Feyerabend) in Research Reported in Reference Work (Chap. 6). This chapter evaluates research reported in the International Handbook of Research in History, Philosophy and Science Teaching (HPST), Springer. Based on the subject index of the handbook, I found 6 chapters that referred to “Feyerabend” or “epistemological anarchism” and were classified according to the following criteria: Level I, Level II, Level III, Level IV and Level V (same as in Chap. 3). Of the 6 articles evaluated, none was classified in Level V. Based on the treatment of the subject by the authors 6 categories were developed to report and discuss the results. Following are some of the findings. Based on the work of Hacking, a scientific experiment is an act of intervention in which material and theoretical entities interact within a cultural and societal context. With this perspective, science should be taught as an exemplar of how knowledge is generally acquired in the empirical sciences. Furthermore, drawing support from Feyerabend, Collins and Polanyi it can be argued that experimental procedures cannot be explicitly communicated because of their tacit nature. Kuhn’s “normal science” has been the subject of considerable controversy in both science and science education as it leads to a distortion of the history of science. However, it has also been argued that some controversies in science concern issues that are not linked to core epistemological commitments within a research tradition, and hence would not undermine what Kuhn had to say about the research training of individual new scientists. Schulz has asked a very pertinent question: who defines science for science educators? It is suggested that the classroom teacher needs the collaboration of the historian (for correcting pseudo-history in textbooks) and the philosopher of science (for

1.2  Chapter Outlines

15

correcting misleading epistemology). This seems to be a sound strategy and has been referred to it as the integration of domain-specific and domain-general aspects of the curriculum for introducing nature of science (NOS). The oil drop experiment forms an important part of the determination of the elementary electrical charge in science curricula of most countries (domain-specific aspect). However, the pseudo-­ history taught in most textbooks does not refer to the difficulties involved (as the experiment was not simple and straightforward, as suggested by the textbooks), nor that a bitter controversy ensued between Millikan and Ehrenhaft that lasted for many years. In his published scientific papers Millikan did not refer to his discarding data. Consequently, a teacher’s conclusion that he did not follow or respect the scientific method rigorously, clearly shows the domain-general aspect. In other words, it is the integration of domain-specific and domain-general aspects that facilitates an understanding of NOS. Postmodernism generally rejects metanarratives which makes it difficult to understand complex phenomena based on idealization that requires control of variables (e.g., Galileo on falling bodies and Newton on planetary motion). Contrary to postmodernists and even perhaps Feyerabend, sometimes Western and indigenous science can agree to an extraordinary extent. The role of objectivity, reason and observation in the standard view of science is the subject of considerable controversy in both science education and the philosophy of science. A historical reconstruction of science shows that it is replete with controversies. If scientists were absolutely objective, the reasons advanced by one group of scientists to understand observations would be accepted by the rival groups, and consequently there would be no controversy. History of science, however, does not provide support to the standard view of science. At present, in most parts of the world, the science curriculum endorses “Baconian methods.” In contrast the new style of history of science, which emphasizes the dynamics of scientific progress and its relation to the philosophical, technological, and social background, is much more suitable for science education than the empirical tradition that stressed the accumulation of facts and the assignment of credit for discoveries. Feyerabend’s claim that if modern science has found the “truth” then that would inhibit freedom of thought, has been the subject of considerable criticism. According to Giere’s perspectival realism scientific theories are neither “true” nor “false”, and thus approximates to Feyerabend’s thesis. For science education it is important to note that if modern science has found the “truth”, then students can be led to believe that perhaps there is not much left to do and that can limit freedom of thought. Feyerabend’s Counterinduction and Science Textbooks (Chap. 7). All scientists working in a field of knowledge do not necessarily agree with respect to all “observations”, “experimental results” and “theories.” Consequently, if we tell students to accept only those theories which are consistent with the available and accepted facts, we shall be left without any theory. To solve this dilemma, Feyerabend suggested a change in methodology by admitting counterinduction, namely accepting unsupported hypotheses. The objective of this chapter is to explore the following historical episodes as examples of counterinduction and draw implications for science textbooks and possible teaching strategies. In the late nineteenth and early twentieth century, atomic theory was still being questioned by some leading physicists (Duhem

16

1  Introduction: Exploring Epistemological Anarchism

& Ostwald) and consequently Brownian motion could not be explained by the kinetic-molecular theory, or in other words support could be accepted p­ rovided we considered it as an example of counterinduciton. Some general chemistry textbooks explicitly referred to the role played by Einstein’s work on the kinetic-­molecular theory and experimental support by the work of Perrin, which led to an explanation of Brownian motion and was classified as an example of counterinduction in the context of Feyerabend’s epistemological anarchism. In his kinetic theory of gases, J.C. Maxwell’s simplifying assumptions were precisely the ceteris paribus clauses, which helped him to progress from simple to complex models of the gases. Taking our cue from Feyerabend’s counterinduction and Galilean idealizations, it is plausible to interpret Maxwell’s basic assumptions as a move from the complexity of nature to the specially contrived order of the experiment. One general chemistry textbook emphasized the following aspects of Maxwell’s assumptions: speculative, models, approximate, and that models develop (tentativeness) in order to explain the behavior of gases. It is particularly important to note the following assertion by the textbook, “a model can never be proved absolutely true.” Interestingly, not only Feyerabend but also recent philosophy of science would endorse such a thesis (cf. Giere). The Michelson-Morley experiment was first conducted in 1887 and provided a “null” result with respect to the ether-drift hypothesis, namely, that there was no observable velocity of the earth with respect to the ether. According to Lakatos starting in 1905, it took almost 25 years for this hypothesis to be refuted and recognized as the “greatest negative experiment in the history of science”. Feyerabend presents a different interpretation of the events: The special theory of relativity (STR) was retained, despite D.C.  Miller’s decisive refutation. Lakatos implies that empirical evidence was necessary for refuting the ether-drift hypothesis, whereas Feyerabend would imply that despite empirical evidence to the contrary, STR was not refuted (counterinduction). Physics textbooks published in different countries even today generally emphasize that it was the Michelson-­Morley experiment that led Einstein to postulate his STR. With Feyerabend’s perspective of counterinduction it would be interesting to find if science textbooks explore: Despite empirical evidence to the contrary, STR was not refuted. The oil drop experiment is generally considered to be simple, beautiful and straightforward, that unambiguously led to the determination of the elementary charge. A historical reconstruction, however, shows although Robert Millikan and Felix Ehrenhaft obtained very similar experimental data, their interpretations were entirely different, leading to considerable controversy. Millikan postulated the existence of a universal charged particle (the electron) whereas Ehrenhaft postulated the existence of sub-electrons based on fractional charges. Later, in 1978 Holton found that Millikan in his 1913 article did not include data from 59% of the experimental drops. Thus, Millikan discarded data that did not support his guiding assumption—this coincides with Feyerabend’s claim that scientific theories are not consistent with all the experimental data. Furthermore, Millikan supported a theory that was not supported by at least 59% of his data, and hence accepted a theory that was at least partially unsupported, and Feyerabend would consider this as counterinduction. On the contrary, many general chemistry and physics textbooks state that Millikan included all the experimental data. It is generally ignored that both

1.2  Chapter Outlines

17

J.J. Thomson and E. Rutherford performed alpha particle scattering experiments in their respective laboratories. Although, results from both laboratories were similar, interpretations of Thomson and Rutherford were entirely different. Thomson propounded the hypothesis of compound scattering, according to which a large angle deflection of an alpha particle resulted from successive collisions between the alpha particle and the positive charges distributed throughout the atom. Rutherford, in contrast, propounded the hypothesis of single scattering, according to which a large angle deflection resulted from a single collision between the alpha particle and the massive positive charge in the nucleus. The rivalry between Rutherford’s hypothesis of single scattering based on a single encounter and Thomson’s hypothesis of compound scattering, led to a bitter dispute between the proponents of the two hypotheses. Given, Thomson’s credentials most scientists could argue that Rutherford’s hypothesis of single scattering, perhaps constituted Feyerabend’s counterinduction. Historical events, however, turned out to be otherwise and the scientific community eventually accepted Rutherford’s arguments. It is plausible to suggest that Thomson’s hypothesis of compound scattering provided counterinduction and thus helped Rutherford to strengthen his arguments. In 1913 when Bohr first published his model of the atom, it had a fairly adverse reception in the scientific community. Lakatos, however, considered it as acceptable scientific practice—an example of growth on inconsistent foundations, which can easily be construed as counterinduction as postulated by Feyerabend, namely postulating hypotheses that are not entirely supported by experimental evidence. In contrast, Popper a falsificationist had recommended that Bohr’s article should not have been published in 1913 as it was based on inconsistent foundations. A general chemistry textbook considered Bohr’s contribution as a bold new assertion that initially did not seem to follow from the data (Feyerabend’s counterinduction). During the latter half of the nineteenth century nature of the photoelectric current was not clear and led to considerable controversy. In 1905 Einstein proposed that ordinary light behaves as though it consists of a stream of independent localized units of energy that he called lightquanta. According to Einstein, if light consists of localized quanta of energy, an electron in an atom will receive energy from only one lightquantum at a time. Based on Einstein’s photoelectric equation, in 1916 Millikan determined the experimental value of Planck’s constant h. However, Millikan in the same publication accepted Einstein’s photoelectric equation, but questioned the hypothetical lightquantum, the theoretical base of the photoelectric theory. Based on his presupposition (i.e., wave theory of light), Millikan considered Einstein’s hypothesis as reckless—a clear example of counterinduction, viz., accepting an unsupported hypothesis (wave theory). For textbook authors and classroom teachers, Holton has raised a very pertinent question: Millikan’s determination of h is not, as we might now naturally consider it to be, an experimental proof of the quantum theory of light. Interestingly, one study found that almost 87% of general physics textbooks (published in U.S.A., between 1950s to 2000s) considered Millikan’s determination of h, as an experimental proof of the quantum theory of light. It is plausible to suggest that the inclusion in science textbooks of the following aspects related to the photoelectric effect can facilitate a better understanding of the dynamics of scientific progress: (a) Millikan considered Einstein’s hypothesis as

18

1  Introduction: Exploring Epistemological Anarchism

reckless—in other words accepting Einstein’s photoelectric equation constituted a clear example of counterinduction, viz., accepting unsupported hypothesis; (b) Einstein’s hypothesis was not accepted by the scientific community, including Planck, the ‘originator’ of the quantum hypothesis, for many years; (c) Millikan presented experimental evidence to support Einstein’s photoelectric equation and still rejected his quantum hypothesis; (d) scientific theories are underdetermined by experimental evidence, that is, no amount of experimental evidence can provide conclusive proof for a theory (these aspects can be included in the textbooks by presenting Millikan’s experimental determination and at the same time pointing out that this was not considered as sufficient evidence for Einstein’s theory). Starting in 1922, De Broglie applied the wave-particle duality hypothesis to existing problems in physics and provided the first physically plausible explanation for the Bohr-Sommerfeld stability rules. De Broglie even suggested a possible experimental confirmation of his controversial hypothesis based on diffraction phenomena. Despite Einstein’s support, duality remained a controversial hypothesis, until conclusive experimental evidence was presented by Davisson and Germer in 1927. Most general chemistry textbooks present atomic structure by referring to the work of J.  J. Thomson, R. Millikan, E. Rutherford and N. Bohr. Following this, Einstein’s interpretation of the photoelectric effect is presented as an application of quantum theory. Bohr’s model of the atom was the first to depart from the classical wave theory of light by introducing the ‘quantum of action’. Next, in order to introduce the wave mechanical model of the atom (E. Schrödinger), De Broglie’s contribution is mentioned by posing the question: if light can have both wave and particle properties then why particles of matter (argument of symmetry in nature) cannot also have both properties. Furthermore, experimental work of C. Davisson and L. H. Germer is reported based on diffraction of electron beams by metal foils. One general chemistry textbook presented almost a historical reconstruction of wave-­particle duality, starting with the contributions of Planck in 1900 to Einstein in 1905, Bohr in 1913, de Broglie in 1924, and finally 3 years later Davisson and Germer in 1927. Furthermore, it referred to the fact that De Broglie’s conceptualization of wave-particle duality preceded its experimental determination by Davisson and Germer, and emphasized the argument of symmetry in nature, a bold hypothesis. It is plausible to suggest that this can be considered as an example of counterinduction. Indeed, Feyerabend would endorse that science involves not only “hypothesizing” but also “pondering” that leads to “counterinduction.” On comparing such presentations in textbooks, a thoughtful student may wonder if these are presenting chemistry or history of chemistry. Indeed, this is the dilemma faced by most science teachers and textbook authors. However, a critical appraisal of most of our current textbooks would show that if we want to understand science, its history cannot be ignored. In other words, the history of chemistry is ‘inside’ chemistry and according to Feyerabend history of science becomes an inseparable part of the science itself. Despite some ambivalence, D. Mendeleev acknowledged the role played by the atomic theory to explain periodicity in the periodic table. The placing of tellurium and iodine (I) caused considerable problems. In order to solve the problem, Mendeleev reduced the atomic mass of Te from 128 to 125, which thus preceded that of I, and could then be placed in the

1.2  Chapter Outlines

19

appropriate group. This change and others can be considered as examples of counterinduction? In other words, Mendeleev’s periodic law (ascending order of atomic weights) was not supported by the experimental data (atomic weights of Te and I), and consequently he changed the data (128 to 125), and thus accepted an unsupported hypothesis—counterinduction. Despite Mendeleev’s considerable expertise in predicting new chemical elements, for which he left empty spaces in his periodic system, he was taken by surprise by W. Ramsey’s discovery in 1894 of a new element, later named as argon. Mendeleev at first refused to accept argon as an element. Instead, he postulated the hypothesis that argon could be triatomic nitrogen (N3), in analogy to ozone (O3), and that would also explain why it did not react with any other element. This is a clear example of counterinduction as Mendeleev had no empirical evidence to postulate argon as triatomic nitrogen. Despite the controversy with respect to the placing of argon, some general chemistry textbooks not only ignore it but instead state that the periodic table did not need any revision. This clearly shows how history of science can help to understand science. It is generally recognized that G.N.  Lewis in 1916 presented the first satisfactory model of the covalent (shared electron pair) bond based on the cubic atom. However, it is important to note that the genesis of the cubic atom can be traced to an unpublished memorandum written by Lewis in 1902. Given the hegemony of the ionic (transfer of electrons) bond, pairing (sharing) of electrons seemed at first glance to be a bizarre idea. It was in the early 1920s that Lewis’s covalent bond started to gain support. Lewis’s cubic atom was first conceived as a teaching device to illustrate the octet rule and can be considered as primarily ‘speculative’. Clearly, in 1902, Lewis did not have all the empirical data and hence the hypothesis of the cubic atom was at best only partly supported by experimental evidence—a clear example of counterinduction, as suggested by Feyerabend. The existence of the planet Neptune was predicted in 1846 before it was discovered in order to explain observed aberrations in the orbit of another planet (Uranus). Postulation of the planet (Neptune) remained an unsupported hypothesis until it was discovered, namely counterinduction. Lakatos not only endorsed a similar interpretation but even went beyond by postulating repeated counterinductions. By the end of the 1920s it was found that energy conservation does not seem to hold for beta decay reactions (changing a neutron into a proton and an electron in radioactivity). In order to overcome this difficulty, in 1929 W. Pauli postulated the elementary particle neutrino—counterinduction, as there was no experimental evidence of their existence. Neutrinos were widely accepted by the scientific community before they were finally discovered in 1956. Starting in 1963, it was in 1979 that Martin Perl and colleagues could claim that all confirmed measurements agreed with the hypothesis that the tau is a lepton produced by a known electromagnetic interaction and that it decays through the conventional weak interaction. It is important to note that for almost 16 years, the hypothesis that the Tau is a Lepton remained unsupported (lacked experimental evidence) and still the scientists continued to work on it—a clear example of counterinduction. Conclusion: Feyerabend and challenges of the twenty-first century (Chap. 8). Despite Feyerabend’s hyperbolic flourishes, such as anti-rationalism and anything goes, his main objective was to present a picture of “how science really works.” For

20

1  Introduction: Exploring Epistemological Anarchism

example, anything goes also means that the methodological strategy is not decided a priori. Recent philosophy of science has endorsed Feyerabend’s vision as appropriate for the twenty-first century. Of all the 120 articles related to science education, evaluated in this book only 9% approximated to an understanding of Feyerabend as a philosopher of science trying to investigate how science really works. Michelson-Morley experiment has been the subject of considerable controversy. According to Lakatos, it took 25 years for the ether-drift hypothesis to be refuted, whereas Feyerabend, has claimed that Einstein’s special theory of relativity (STR) was retained despite empirical evidence (Michelson, Miller) and including Lorentz contraction, to the contrary. Lakatos implies that empirical evidence was necessary for refuting the ether-drift hypothesis, whereas Feyerabend would imply that despite empirical evidence to the contrary, STR was not refuted. This has educational implications especially for writing textbooks. In other words, a review of the literature at present shows that researchers were interested in finding: Michelson-Morley (MM) experiment led Einstein to postulate his special theory of relativity (STR). With Feyerabend’s perspective of counterinduction it would be interesting to find if science textbooks explore: Despite empirical evidence to the contrary, STR was not refuted. The role of counterinduction has also been explored in this chapter with respect to other historical episodes, such as Brownian motion, Kinetic theory, Oil-­ drop experiment, Alpha particle scattering experiment, Bohr’s “quantum of action”, Photoelectric effect, Wave-particle duality, Mendeleev’s periodic table and Lewis’s postulation of the covalent bond. Given Kuhn’s influence some science curricula and especially the textbooks emphasize “normal science”. In contrast, Feyerabend would recommend alternative and competing approaches to understanding science. Actually, all accepted theories may eventually change, showing that science is tentative. This precisely leads to understanding the tentative nature of science and hence Feyerabend’s advice to students that “current view of science may soon be voted out of office.” The legacy of logical positivism leads science education to emphasize that there is only one “true” or “correct” theory to understand a phenomenon in science. Some of the examples discussed in this book show the contrary (e.g., acid-base equilibria, valence bond and molecular orbital theories to understand covalent bonding and Copenhagen and Bohmian interpretations of quantum mechanics). Thus, it is plausible to suggest that science can provide more than one “correct” theory of the same phenomena and that would represent what Feyerabend referred to as epistemological pluralism. There is some concern in science education that introduction of epistemological anarchism may lead to lifting the lid off the Pandora’s Box. On the contrary, if we want to go beyond “two centuries of empiricism” and understand how science really works, then the inclusion of alternative approaches is essential. On the contrary a deeper understanding of Feyerabend’s philosophy of science shows that it does not rule out stringent testing. For example, the strategy of counterinduction can lead to the consideration of hypotheses with even no empirical support and thus precisely facilitating conceptual clarity (e.g., evolutionary theory and intelligent design; Lavoisier’s oxygen theory and Priestley’s phlogiston theory, for details see Chap. 7). In order to understand the controversy between the oxygen and phlogiston theories, it is important to note that the latter too

1.2  Chapter Outlines

21

was a comprehensive theoretical system. The “clash” between Lavoisier’s oxygen theory and a scientifically unacceptable theory (Priestley’s phlogiston theory) today, approximates to Feyerabend’s counterinduction. In order to understand the controversy successful students must become “knowledgable phlogistonists” and this precisely leads to Feyerabend’s (1993) suggestion that the history of science becomes an inseparable part of a science itself. A basic idea of the modern natural sciences is bound with an appreciation that they are objective rather than subjective accounts. This has been questioned on the ground that both objective and subjective aspects play an important role in scientific progress. Furthermore, empirical data generally presuppose a theoretical background. This provides an opportunity to understand the constant confrontation between objectivity and subjectivity in the history of science and thus the need to go beyond an exclusively empirical approach. Feyerabend’s methodological pluralism means that the scientist is free to employ all methodological approaches that seem fruitful, with the understanding that some of the attempts may not succeed, and the scientist has to be prepared to accept the results. Feyerabend was generally both critical and skeptical of scientific expertise (especially under the aegis of the state), and hence his recommendation that science and state be separated. Recent research has shown some support for Feyerabend’s views with respect to scientific expertise. The traditional scientific method based on the lecture, lab and demo approach is generally followed by most science educators. Instead of stockpiling of experimental data, it is the tacit knowledge (Polanyi), among other sources, that helps a scientist to understand the real significance of science. In contrast, Feyerabend’s approach would emphasize how science really works, namely uncertainty, consideration of wasted efforts and critical encounters with peers. Metaphysical criticism must not be allowed to make us reject inductive proofs, is generally endorsed by school science. Duhem, however, has argued that Newton formulated his method a posteriori to protect and strengthen the law of gravitation, which revealed some of the skeletons in the Newtonian cupboard. For a science teacher it is important to note that if the new grew out of the old (Bunge), very soon the new itself may be voted out of office (Feyerabend), and consequently it sounds good policy to follow some of the guidelines of the old (Cartwright). This sounds a plausible strategy despite the different philosophical perspectives of Bunge, Cartwright and Feyerabend. Teaching and doing science is a very complex issue, especially when these are enmeshed with various socio-political and cultural controversies, and this is precisely the context in which Feyerabend situated the problems (e.g., evolution and intelligent design controversy). No wonder, Feyerabend considered that astrology and voodoo are attractive alternatives to evolution, especially if it is presented as based on “simple inspection of phenomena” or “dogmatic scientism.” Indeed, most science curricula and textbooks reduce “scientific practice” to a “simple inspection of phenomena” and thus ignore the unnatural nature of science. Feyerabend’s philosophy of science was based on how science really works and recent literature in philosophy of science has recognized that his oeuvre was more modern (in the Enlightenment tradition) than postmodern. Furthermore, progress in science is perspectival, rather than a quest for “correct or true theories.” In a historicized vision of scientific judgment, there is no objectivity without ­subjectivity

22

1  Introduction: Exploring Epistemological Anarchism

to suppress and vice versa (Daston & Galison). It is precisely in such a context that Feyerabend claimed that “true belief limits freedom.” Thus, Feyerabend is not only a perspectivist but also a perspectival realist.

Chapter 2

Epistemological Anarchism and How Science Works

2.1  Introduction It is well known that most of Feyerabend’s ideas on epistemological anarchism developed as a constant dialogue with Imre Lakatos (cf. Motterlini, 1999). Furthermore, in developing his ideas, Feyerabend was constantly critiquing both Karl Popper and Lakatos. In order to facilitate a better understanding of the debate between Feyerabend and Lakatos, here I present a brief description of the Lakatosian Methodology of Scientific Research Programs (MSRP). (I am grateful to David Geelan for this suggestion). According to Lakatos (1970) the basic unit of appraisal must not be an isolated theory or conjunction of theories but rather a “research program,” with a conventionally accepted “hard core” and with a “positive heuristic” which defines problems, outlines the construction of a “belt of auxiliary hypotheses,” foresees anomalies and turns them victoriously into examples, all according to a preconceived plan. The “negative heuristic” represents the “hard core” of the program, consisting of basic assumptions considered “irrefutable” by the methodological decision of its protagonists, and does not allow modus tollens to de directed at this hard core. The positive heuristic represents the construction of a “protective belt” consisting of a partially articulated set of suggestions or hints on how to change, develop the “refutable variants” of the program. The positive heuristic saves the scientist from becoming confused in the “ocean of anomalies” by directing the modus tollens at the “auxiliary hypotheses”. The scientist lists anomalies, but as long as his research program sustains its momentum, he may freely put them aside, and it is primarily the positive heuristic of the program, and not the anomalies, which dictate the choice of problems. There is no such thing as crucial experiments, in the sense that a theory is falsified (refuted) when a statement that expresses the result of an observation is in contradiction with a statement of the theory—that is, naive falsificationism. A theory is refuted, by a rival research program which explains the previous success of its © Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_2

23

24

2  Epistemological Anarchism and How Science Works

rival and supersedes it by a further display of heuristic (explanatory) power, and not by a crucial experiment. Lakatos considers this to be his sophisticated falsificationism. In contrast to Popper, both Feyerabend and Lakatos agreed on the role of crucial experiments, in the history of science. A research program is progressing if it frequently succeeds in converting anomalies into successes, that is, explainable by the theory—referred to as “progressive problemshifts”. The classic example of a successful research program is Newton’s gravitational theory (however, for a criticism of the Newtonian method by Lakatos, see Chap. 3). Lakatos also deals extensively with Bohr’s atomic theory in order to illustrate various aspects of his philosophy. According to Lakatos, Bohr’s famous postulates constitute the negative heuristic of his research program, which helped him to develop his theory. On the contrary, most general chemistry and physics textbooks consider the postulates in themselves as Bohr’s theory. The textbook presentations lack the appreciation of how scientists face difficulties and resort to various methodological strategies.

2.2  Feyerabend’s Epistemological Anarchism In this section, I present some of the ideas that guided Feyerabend to develop his epistemological anarchism: 1. According to Feyerabend (1970b): “… it is not only possible but also desirable to introduce and elaborate hypotheses which are inconsistent with highly confirmed theories and with evidence [for example], evidence which refutes a theory can often be found only with the help of an alternative so that the advice to postpone alternatives until the first refutation has occurred puts the cart before the horse” (pp.  275–276). After reading a preliminary version of this chapter, Kalman (2019a) suggested: “This is very important. Feyerabend maintained that Galileo only arrived at his new theories by considering such alternatives. Evaluation of a theoretical framework does not occur until there is an alternative (principle of counterinduction). This principle is very different from anything proposed by Lakatos or Popper.” 2. Based on historical episodes such as the Copernican Revolution, atomism (both in antiquity and present), kinetic theory, dispersion theory, stereochemistry, quantum theory, Feyerabend (1970a) has concluded: “The idea of a method that contains firm, unchanging, and absolutely binding principles for conducting the business of science gets into considerable difficulty when confronted with the results of historical research. We find, then, that there is not a single rule, however, plausible, and however firmly grounded in epistemology, that is not violated at some time or other. It becomes evident that such violations are not accidental events, they are not the results of insufficient knowledge or of inattention which might have been avoided. On the contrary, we see that they are necessary for progress” (pp. 21–22).

2.2  Feyerabend’s Epistemological Anarchism

25

3. Based on J.S. Mill’s On Liberty, Feyerabend espouses a pluralistic methodology for the natural as well as the social sciences. Actually, Motterlini (2002b) recognizes that even Lakatos accepted that the interaction between philosophy and history of science can be more fruitful if “enough pluralism is tolerated” (p. 501), especially if we want to compare and learn from different historical reconstructions that leads to the proliferation of various points of view. According to Farrell (2003, p. 133), pluralism is the “hard-core” of the Feyerabendian philosophical program and it came to permeate all aspects of his thought. 4. According to Feyerabend, science as we know it cannot be based entirely on theories that are consistent with all the facts and hence the need to admit counterinduction and unsupported hypotheses (Feyerabend, 1975a, p.  43). Counterinduction requires the postulation of a new theory that clashes with established observational results and confounds plausible theoretical principles (p. 45). 5. As opposed to Popper, Lakatos considers that new ideas in science are almost entirely ad hoc. Feyerabend (1975a) supports this Lakatosian stance (p. 64), and asserts that adhocness is neither despicable nor absent from the body of science. After reading a preliminary version of this chapter, Kalman (2019a) suggested, “Sir John Herschel maintained that theoretical statements derived inductively from experiments and wild guesses are equally acceptable provided that their deductive consequences are confirmed by observation.” 6. Feyerabend (1970a) emphasized that, “… the idea of a fixed method, or of a fixed (theory of) rationality, arises from too naïve a view of man and of his social surroundings. To those who look at the rich material provided by history, and who are not intent on impoverishing it in order to please their lower instincts, their craving for intellectual security as it is provided, for example, by clarity and precision, to such people it will seem that there is only one principle that can be defended under all circumstances, and in all stages of human development. It is the principle: anything goes” (pp. 25–26). In an end note Feyerabend clarifies that “anything goes” reflects the changes in theories of knowledge, namely we find new principles and abandon old ones. The idea of a fixed method has also been questioned by James Bryant Conant (practicing organic chemist and President of Harvard University): “I believe almost all modern historians of the natural sciences would agree … There is no such thing as the scientific method. If there were, surely an examination of the history of physics, chemistry, and biology would reveal it …” (Conant, 1951, p.  45). Among other, Feyerabend draws on the following historical episodes: (a) Copernicanism; (b) Newton’s theory of gravitation; (c) Bohr’s atomic model (the model was retained despite precise evidence to the contrary). A practicing physicist and Nobel Laureate has referred to Bohr’s dilemma in the following terms: “In 1913 Niels Bohr proposed his famous theory of the hydrogen atom. One cannot say that he resolved the problem raised by Rutherford. In a sense he crystallized the dilemma in an even more dramatic form. Focusing his attention entirely on the construction of

26

2  Epistemological Anarchism and How Science Works

a nuclear atom, Bohr took what principles of classical physics he needed and added several nonclassical hypotheses almost without precedent; the mélange was not consistent. But they formed a remarkably successful theory of the hydrogen atom” (Cooper, 1970, p. 325); d) Special and the general theory of relativity (especially, in the context of D.C. Miller’s experimental refutation and the earlier Michelson-Morley experiments). In this section I have tried to show that the difficulties involved in the historical episodes referred to by Feyerabend have been recognized even by practicing scientists. 7. Theory-laden nature of scientific observations holds an important place in Feyerabend’s philosophical framework. This has important implications for science education as it emphasizes the difference between observation and inference. In contrast, Feyerabend (1962) considers that observation statements are precisely those that are accepted with relative ease without much discussion as they are based on theories that have acquired consensus in the scientific community.

2.3  Feyerabend Versus Popper and Lakatos Feyerabend criticizes both Popper and Lakatos for on the one hand advocating standards that facilitate a rational understanding of progress in science and at the same time recognizing that these standards need to be flexible and even perhaps liberalized. For example, Popper has himself recognized that standards are not always adopted on the basis of argument (Feyerabend, 1975a, p. 79). After surveying possible sources for standards, Popper (1963b), The Open Society and its Enemies, asks a pertinent question, “What, then, are we to trust?” and responds in the following terms: The answer is: whatever we accept we should trust only tentatively, always remembering that we are in possession, at best, of partial truth (or right-ness), and that we are bound to make at least some mistake or misjudgement somewhere—not only with respect to facts but also with respect to the adopted standards; secondly, we should trust (even tentatively) our intuition only if it has been arrived at as the result of many attempts to use our imagination; of many mistakes, of many tests, of many doubts, and of searching criticism (p. 391).

Feyerabend (1975a) considers that even such an understanding of the standards (with all its pitfalls) is far from being applicable in order to guarantee the rationality of science and at best remains a “verbal ornament” (p. 79). For an understanding of how Feyerabend’s views on rationalism developed see Shaw (2019). In the case of Lakatos’s (1970) Methodology of Scientific Research Programs (MSRP), Feyerabend (1975a) considers that a strict principle of falsification (or even naïve falsification) combined with the demand for maximum testability and non-adhocness would not only wipe out science as we know it but also make it difficult for it to even start (p. 77).

2.4  Was Lakatos an Epistemological Anarchist?

27

Having recognized this difficulty, Lakatos has suggested a change in his critical standards, so that a theory needs to be evaluated in the long run after it has had an opportunity to show its merits (explanatory power). Consequently, if the theory gives rise to interesting new developments (progressive problem shifts in the terminology of Lakatos) over time then it can be retained despite its initial drawbacks. At this stage, Feyerabend (1975a) raises the following thorny issue: “Now it is easily seen that standards of this kind have practical force only if they are combined with a time limit. What looks like a degenerating problem shift may be the beginning of a much longer period of advance, so—how long are we supposed to wait?” (p. 77, italics in the original). The recurring question is of course that why not wait a little longer. Next, Feyerabend provides examples of two theories (heliocentric theory and atomic theory) that were considered as degenerating for centuries and later returned to the stage in full bloom.

2.4  Was Lakatos an Epistemological Anarchist? Given Lakatos’s formulation of MSRP, this question may sound strange. However, given the close relationship between Feyerabend and Lakatos, this is a plausible thesis. Most of Lakatos’s ideas are based on a peculiar philosophical combination of Hegelian historicism and Popperian fallibilism, which led Feyerabend to suggest that he was a philosophical bastard, namely a “Pop-Hegelian” born from a Popperian father and a Hegelian mother (Motterlini, 1999, pp.  184–185). In his later work, Lakatos (1971a, 1971b) explicitly differentiated between methodology and heuristic principles. In contrast to methodological rules, the latter instruct the scientists how to proceed within a particular research program. In his “Replies to Critics” Lakatos (1971b) stated: “I can judge: I can say whether they have made progress or not. But I cannot advise them—and I do not wish to advise them—about exactly what to worry and in which direction they should seek progress” (p. 178). Motterlini (2002b) considers such a statement as, “The last nail in the coffin of the early Lakatos is that methodology (divorced from normative heuristic) aims to advise scientists neither about how to arrive at good theories nor even about which of two rival programmes they should work on” (p. 497). Feyerabend (1975a, 1975b) considers that this inclusion of the role of the owl of Minerva (a term introduced by G.W.F. Hegel who stated that the owl of Minerva spreads its wings only with the falling of the dusk, meaning that philosophy comes to understand a historical condition just as it passes away), leads Lakatos to oscillate between two extremes: a conservative position that emphasizes methodological standards of appraisal (hence the status quo) and a revolutionary position which amounts to “anything goes.” This leads to the following conundrum: “Whether or not Lakatos has really gone a long way towards epistemological anarchism” (Motterlini, 2002b, p. 498).

28

2  Epistemological Anarchism and How Science Works

In order to solve this problem Farrell (2003) has suggested that if the rational prescriptions (MSRP) are interpreted as values, rather than rules then that would help to remove contradictions in the Lakatosian methodology and concluded: If Lakatos were to have dropped the formal machinery from MSRP, then Feyerabend would have applauded Lakatos as a ‘fellow anarchist’: Lakatos recognized the inherently conflicting nature of rationality and if Lakatos had interpreted this conflicting rationality in terms of non-rules-based values, then Lakatos would have upheld a position virtually identical to Feyerabend’s” (Farrell, 2003, p. 206).

However, for Lakatos to have accepted such changes in his MSRP explicitly, was very difficult. Musgrave (1976a, 1976b) a close friend, student and colleague of Lakatos would agree that not only he had gone a long way towards epistemological anarchism, but also abdicated the methodologist’s throne: Some, of course, welcome ‘the abdication of the Lakatosian methodologist.’ It is with glee that Feyerabend exclaims “scientific method, as softened up by Lakatos, is but an ornament which makes us forget that a position of ‘anything goes’ has in fact been adopted” [Feyerabend, 1970c, p. 229 ]. But I do not welcome it, and I regard it as an aberration given Lakatos’s general philosophical position. How can his methodology help us to stem intellectual pollution, if it has no repercussions for intellectual practice? We must try to extricate Lakatos from anarchism, by taking a critical look at the arguments which have led him, unwittingly, towards it (pp. 476–477, italics added).

Given Musgrave’s close intellectual relationship with Lakatos for many years this reaction is understandable. On the one hand it shows his disagreement and even consternation, and on the other a call for “extricating” Lakatos from epistemological anarchism. The intellectual relationship between Feyerabend and Lakatos (cf. Motterlini, 1999), as revealed in their debate with respect to Against Method, clearly shows an underlying coincidence as to how science is practiced. Interestingly, Musgrave also refers to “intellectual practice” and surprisingly despite his close relationship with Lakatos he did not foresee this crucial aspect of his philosophy. Finally, Musgrave clarifies that although Feyerabend is right that Lakatos came to adopt an anarchistic position with respect to advice to be given to scientists for developing a particular research program, the same is not true for existing theories. In other words, Lakatos did provide criteria in the light of which one program could be judged superior to another.

2.5  Feyerabend and Scientific Expertise Feyerabend’s views on the role of scientific experts in the resolution of problems that affect the lives of the general public are perhaps the most controversial. In this context his views stand in sharp contrast to those of Collins and Evans (2002, 2007). According to Feyerabend (1982) in a democratic and pluralistic society the social, religious and other beliefs of the people are most important. Contrary to Feyerabend,

2.5  Feyerabend and Scientific Expertise

29

Collins and Evans consider the notion of scientific expertise as an indispensable resource for democratic culture. According to Sorgner (2016, p. 115), Collins’ and Evans account can even be considered as a hypothetical response to Feyerabend’s attack on scientific expertise. Although some of Feyerabend’s arguments can be rejected within the Collins/ Evans framework, especially with regard to the concepts of interactional expertise and experience-based experts, there are also some common elements between the two approaches (Sorgner (2016, p. 115). Interactional expertise according to Collins and Evans involves tacit knowledge, which is acquired by embedding in the relevant group of specialists. Without tacit knowledge successful participation in technical debate is not possible (Collins & Evans, 2007, p.  22). Furthermore, Collins and Evans introduce the Periodic Table of Expertises as a tool to differentiate and analyze various kinds of expertise. This leads Collins and Evans to assert that ordinary citizens also possess to a certain degree the ability to discriminate between experts and non-experts, as this forms part of ubiquitous expertise that comes naturally with living in society. According to Sorgner (2016, p. 118) this comes quite close to what Feyerabend (1982) referred to as “natural shrewdness of the human race” (p. 98). According to Collins’ and Evans (2007), “Democracy cannot dominate every domain—that would destroy expertise—and expertise cannot dominate every domain—that would destroy democracy” (p. 8). Feyerabend (1987), on the other hand, has argued throughout his later work that science must also be protected from the uninformed application of external standards, for example in Farewell to Reason: “[S]cience, within our democracies, needs protection from non-scientific traditions (rationalism, Marxism, theological schools, etc.) and non-scientific traditions need protection from science” (p.  41). After comparing the two perspectives, Sorgner (2016, p. 119) has rightly suggested that this insistence by Collins and Evans on distinguishing between a technical and a political phase of decision-making closely resemble Feyerabend’s demand that science and society be duly separated. The study by Wynne (1989) dealing with “Sheepfarming after Chernobyl” is a good example of lay expertise and illustrates the differences between the approaches of Feyerabend and Collins and Evans. Wynne describes the conflicts between Cumbrian sheep farmers and British agriculture authorities dealing with the radioactive contamination of pastures and sheep after the Chernobyl fallout. Wynne shows that the ignorance of the authorities towards the local knowledge and customs resulted not only in distrust and frustration, but also financial hardship for the farmers; much of which could have been avoided had the official scientists recognized the farmers’ specific expertise and consulted it. For Collins and Evans, this case does not provide a justification for laypeople to participate in technological decision-making in general, but rather illustrates that technical expertise can be found among the non-professional population. In contrast to Collins’ and Evans’ reading, according to Sorgner (2016, p. 118), in Feyerabend’s terms, the issue at stake was not that the sheep farmers’ specialist contributory expertise remained unrecognized by the authorities, but that two different traditions—the abstract sci-

30

2  Epistemological Anarchism and How Science Works

entific and the pragmatic of day-to-day farm work—failed to engage in an open dialogue. This case study cogently illustrates the differences between the two approaches. Finally, Sorgner (2016) emphasizes that Feyerabend’s approach is still helpful in resolving the issue of scientific expertise: Feyerabend and Collins & Evans share the view that scientists should be treated asspecialists in a very narrow domain, not as universal authorities. The insight that true competence and reliable expertise can only be found within the small group of specialized researchers dealing with a specific problem is one of the reasons why Collins and Evans suggest that experts should be identified with regard to their experience rather than their affiliation. Also, quite similar to Feyerabend, they oppose any kind of scientism that cleaves to some canonical model of scientific method, ignores the political embeddedness of technological debates in the public domain or holds that science alone has the right answer to any question (p. 119).

Motterlini (2016) comes quite close to this understanding of scientific expertise by stating: “Feyerabend’s corrosive skepticism is here directed towards the uncontrolled and uncritical, yet all powerful, authority of ‘scientific expertise’” (p. 5). In some areas of everyday life (e.g., climate change) the role of scientific expertise is complex as it may require specialized knowledge that lay people may not have and thus lead to “argument from authority” (I am grateful to David Geelan for this suggestion).

2.6  Feyerabend Versus Galileo and Copernicus Similar to his views on scientific expertise, Feyerabend was critical of Galileo for his support of the heliocentric and geokinetic hypothesis of Copernicus. Galileo’s main argument was that the Church was not a scientific authority and hence it should not be invoked to invalidate astronomical claims that are proved or provable. Of course, Galileo fully realized that the new data provided by the telescope (1609–1612) was not conclusive evidence to support the Copernican hypothesis of earth’s motion. Feyerabend in contrast argued that Galileo advocated the uncritical acceptance by society of the view of experts, whereas the Church advocated the evaluation by society of the views of experts in the light of human and social values. According to Finocchiaro (1997), Galileo’s defense of Copernicus was characterized by the willingness to know and understand the arguments, evidence and reasons against one’s own views that is, in nut-shell open-mindedness. Actually, Feyerabend (1985) went so far as to endorse the views of Cardinal Robert Bellarmine a controversial figure in the Galileo affair, “The Church would do well to revive the balance and graceful wisdom of Bellarmine, just as scientists constantly gain strength from the opinions of … their own pushy patron saint Galileo” (p.  164). Bellarmine made no secret of his views and in a letter written (dated 12 April 1615) to Carmelite friar Foscarini, a supporter of Copernicus, stated: “However, it is different to want to affirm that in reality the sun is at the center of the world and only turns on itself without moving from east to west, and the earth is in the third heaven

2.6  Feyerabend Versus Galileo and Copernicus

31

and revolves with great speed around the sun; this is a very dangerous thing, likely not only to irritate all scholastic philosophers and theologians, but also to harm the Holy Faith by rendering Holy Scripture false” (Reproduced in Finocchiaro, 1989, p. 67, italics added). Clearly, Bellarmine’s main objective was to defend Scripture. Also, it is important to recall that Bellarmine played an important role in sending Giordano Bruno to the stake in 1600, for having supported Copernicus. However, it is important to note that Bruno was also being tried for heresy by the Roman Inquisition on charges of denial of several core Catholic doctrines (I am grateful to David Geelan for this suggestion). Furthermore, it is well known that Bellarmine requested Galileo to provide proof of the movement of the earth as a condition for revising the interpretation of the Holy Scripture. Interestingly, this led Segre (1997) to ask, “Why was Galileo required to “prove” his heliocentric views when his opponents were not asked to “prove” their geocentric ones? Given the normal ­conventions of seventeenth-century exegesis, on the one hand, and Bellarmine’s (a priori) geocentric conviction, on the other, could it be that requesting proof from Galileo was nothing more than a pretext for enforcing Bellarmine’s own views and, more generally, for subjugating science to theology?” (p. 498, original italics). Actually, Galileo faced a formidable challenge, as in order to provide “proof” for Copernicanism, stellar parallaxes should have been observed, which was not possible with the naked eye and nor with the telescope. It took almost another two centuries when these were observed by Friedrich Bessel in 1838. With this background it is important to note what Galileo wrote in 1615, with respect to his opponents, in a Letter to the Grand Duchess Christina: “These people seemed to forget that a multitude of truths contribute to inquiry and to the growth and strength of disciplines rather than to their diminution or destruction…” (Reproduced in Finocchiaro, 1989, p. 87, italics added). This clearly recognizes the need for a pluralist methodology. Interestingly, Bellarmine’s letter to Foscarini and Galileo’s letter to Christina, were both written in 1615 and show the wide gulf that separated the epistemological views of the two rival philosophers. Now, let us see how Feyerabend conceptualizes this aspect of research. In a letter written to Lakatos, dated 24 January 1971, Feyerabend stated: “The principle of proliferation not only recommends the invention of new alternatives, it also prevents the elimination of older theories which have been refuted. The reason is that such theories contribute to the content of their victorious rivals” (Reproduced in Motterlini, 1999, p. 237, original italics). It is plausible to suggest that Galileo’s “multitude of truths” and Feyerabend’s “principle of proliferation” resemble quite closely. This makes Feyerabend’s criticism of Galileo all the more difficult to understand. It seems that Feyerabend (1993) tried to understand Galileo’s problem by recognizing the historical importance of the “non-rational” method of counterinduction in which hypotheses are introduced that are inconsistent with well-established facts and theories. According to Feyerabend, Galileo was faced with a dilemma: accept the well-supported Ptolemaic hypothesis—a rational decision, or accept the ill-­ supported Copernican hypothesis—a non-rational decision. Thus, according to Feyerabend in order to confound the Ptolemaists, Galileo was using counterinduction, by using the relatively unreliable telescope to support the anomaly-ridden

32

2  Epistemological Anarchism and How Science Works

Copernicanism (cf. Thomason, 1994). Feyerabend’s treatment, however, has been the subject of criticism in the literature (cf. Chalmers, 1985; Machamer, 1973; also see Feyerabend’s 1974 response to Machamer). Thomason (1994), however, has provided support for Feyerabend’s interpretation: “Further, at least in some situations, counterinduction might be necessary for scientific progress. Feyerabend’s reconstruction of Galileo’s procedure is a plausible example. Sometimes it is not only rationally acceptable to work counterinductively, it might also be rationally obligatory to do so” (p. 264).

2.7  Feyerabend and Recent Philosophy of Science In his Against Method, Feyerabend (1993, p. 40) clearly stated that he was not criticizing science but rather the superficial conceptions of science that emerge when philosophers ignore the history of science and its practice. Throughout his intellectual career Feyerabend was considered as a relativist and postmodernist, and philosophers of science questioned his eccentricity, eclecticism and especially his infamous “Anything goes”. Although during the 1970s and 1980s, Feyerabend could be considered to subscribe to some form of relativism and postmodernism (see Farewell to reason), starting from the 1990s (leaving his rhetoric aside) he rejected most forms of relativism and postmodernism (Kidd, 2016). Actually, Feyerabend’s main commitment throughout his career is to pluralism and anti-­ dogmatism (Oberheim, 2016), that led him to question methodological rules in science, and to emphasize the role played by science in human welfare (cf. especially his early paper: Science, Freedom, and the Good Life, Feyerabend, 1968). Feyerabend’s pluralism emphasized: underdetermination of theory by evidence, incommensurability and the non-neutrality of evidence, which led him to suggest that, “Science is, and should be permanent revolution” (cf. Oberheim, 2016, p. 21). In his political philosophy, Feyerabend (1978) was particularly critical of the traditional epistemology of science (i.e., scientific materialism or realism) as it denies disunity and pluralism in science. Brown (2016, p.  150) has drawn a similarity between the views of Feyerabend (ontological pluralist and abundant realist) and Nancy Cartwright, as both are opposed to the idea of reality as a uniform, determinate, universal order or structure that is the object of science, to which all phenomena could be reduced. Furthermore, Brown (2016, p.  151) considers that Ronald Giere’s (2006a, 2006b) perspectival realism, Dupre’s (1993) promiscuous realism and Nancy Cartwright’s (1999) local realism in a “dappled world” all fit well with Feyerabend’s metaphysical views. Interestingly, many of the themes espoused by Feyerabend that were considered to be anti-science and irrationalist (summarized as epistemological anarchism), are now considered to form part of mainstream philosophy of science, such as: science is pluralistic, disunified, value-laden and bound up with social and political concerns (cf. Sankey, 2012, p. 475). For example, science is value-laden as the creation

2.7  Feyerabend and Recent Philosophy of Science

33

of scientific facts is a selection or a valuation of certain observations over others and Feyeraband (2011) expressed this in cogent terms: Turn now to the sciences as they present themselves today. They are free of values, it is said. But that is simply not so. An experimental result or an observation becomes a scientific fact only when it is clear that it does not contain any “subjective” elements—that it can be detached from the process that led to its announcement. This means that values play an important role in the constitution of scientific facts. (pp. 94–95).

Similarly, with respect to pluralistic methodology, Feyerabend (1975a) expressed views that would be endorsed by many present day philosophers of science: A scientist who wishes to maximize the empirical content of the views he holds and he wants to understand them as clearly as he possibly can must therefore introduce other views; that is, he must adopt a pluralistic methodology … knowledge so conceived is not a series of self-consistent theories that converges towards an ideal view; it is not a gradual approach to the truth. It is rather an ever increasing ocean of mutually incompatible (and perhaps even incommensurable) alternatives, each single theory, each fairy tale, each myth that is part of the collection forcing the others into greater articulation and all of them contributing, via this process of competition, to the development of our consciousness (p. 30, italics in the original).

It is interesting to compare this with what Holton (1986) has referred to as the “final temple of science.” Indeed, history of science is replete with examples that illustrate that in some instances it was difficult to separate the “subjective” elements in order to establish the scientific fact, and following are some of the examples: (a) Determination of the elementary electrical charge and the ensuing controversy between R. Millikan and F. Ehrenhaft (cf. Holton, 1978a, 1978b); (b) Interpretation of alpha particle scattering experiments and the controversy between J.J. Thomson and E. Rutherford (cf. Wilson, 1983); (c) Interpretation of observational data based on the bending of light in the 1919 eclipse experiments (cf. Earman & Glymour, 1980); (d) Interpretation of the experimental evidence that led to the discovery of the Tau Lepton (cf. Perl, 2004). For further details and slightly different interpretations see: Daston and Galison (2007), and Niaz (2009, 2018). According to Dupré (1993) epistemological anarchism was “… intended above all as a therapy against the antidemocratic and oppressive consequences of the monopoly of epistemic authority sustained by science” (pp.  262–263). Based on such considerations, Brown and Kidd (2016) have suggested the need for a reappraisal of Feyerabend’s status within the history and philosophy of science (p. 2), especially with respect to postmodernism (generally associated with philosophers such as Lyotard, Rorty, Derrida, Baudrillard, among others). In order to evaluate the status of Feyerabend’s philosophy of science Kidd (2016) considered the following three characterizations of postmodernism: (a) “relativism”, (b) “incredulity to metanarratives” and (c) “depthlessness.” According to relativism a knowledge-claim’s epistemic value is not its relation to reality, but rather its strategic value in relation to the interests of a community. Even in Feyerabend’s later work, according to Gattei (2016): “An often repeated slogan of Feyerabend’s, ‘potentially every culture is all cultures’, is commonly referred to as the epitome of his relativistic views

34

2  Epistemological Anarchism and How Science Works

(Feyerabend, 1994). However, as it is clear from the reading of the short essay that bears this very title, these words actually mean that every culture may change and be transformed through interaction with one another” (p. 89). Gattei (2016) goes on to explore the changing nature of Feyerabend’s views on relativism and show that these words negate relativism and rather affirm that cultures cannot be closed units and their practices and values may be assessed from outside within of course a humanitarian perspective. Actually, relativism in contrast would espouse that cultures are relatively closed units. It is important to note that relativism is not an overarching presupposition of Feyerabend’s philosophy and over his career he manifested various attitudes towards relativism, such as: radical realism, radical agnosticism, radical relativism and finally some sort of in-between realism/relativism mix. According to Farrell (2003): “Relativism, of some sort, could never be denied of Feyerabend’s philosophy, but it is not usually the simplistic relativism of the sort which falls victim to charges of incoherence. In fact, Feyerabend is a realist in many respects and a relativist in others: as Feyerabend stated, ‘I confess to be a fervent relativist in some senses, I am certainly not a relativist in others.’ (Feyerabend, 1991a, p. 507)” (p. 106). Postmodernists consider metanarratives to be paradoxical, authoritarian and dogmatic. Instead, postmodernism favors an eclectic plurality of local and particular micro-narratives. During his early career Feyerabend did endorse some form of relativism and local micronarratives (e.g., Galileo’s case in Against Method). His main argument was that metanarratives could distract our attention away from complex interplay of factors that influence progress in science. For postmodernists, “depthlessness” means that if there is no “true” underlying order to the world, then there is nothing “deep” to explore and an artist/scholar is free to dwell on the surfaces of things. For example, Rorty (1979) rejects the search for deep grounds or foundations for our values and practices. Based on his evaluation of the three characterizations of postmodernism, Kidd (2016) concluded: “It emerges that none of these characterizations offers a strong justification for classifying Feyerabend as “postmodern” in any significant sense. Instead, what does emerge is that Feyerabend’s work was informed by a humanitarian vision of the value of science, that is, strikingly modern” (p.  1, italics in the original). Overall, Kidd (2016) suggested that in his later work Feyerabend decisively rejected relativism and micronarratives and his philosophy of science is more modern (inherited from the Enlightenment) than postmodern. Farrell (2003, p. 209) has suggested that instead of considering Feyerabend as a postmodern, it is better to understand him as a tight-rope walker, namely to be optimally rational it is important to continuously balance incompatible rational demands. This metaphor just like the tight-rope walker masks intense conflict between different rational demands. It has even been suggested that Philip Kitcher’s (2011) Science in a democratic society follows in the footsteps of Feyerabend’s (1978) Science in a free society by incorporating pluralism and dissent (Brown & Kidd, 2016, p.  5). Furthermore, Feyerabend also espoused the construction of an epistemically alert “critical citizenry”, endorsed by many modern writers on science and democracy (e.g., Barnes,

2.8  Feyerabend and Perspectivism

35

Bloor & Henry, 1996; Collins & Pinch, 1998), and even science educators (Hodson, 2009). In an effort to look for the origin and antecedents of Feyerabend’s thinking, Munévar (2016) has argued that Feyerabend extended John Stuart Mill’s epistemological arguments for political freedom to the practice of science. Furthermore, it seems that just as universal laws refer to the mean and are thus bound to have exceptions, so should universal methodological rules, as Feyerabend argued in Against Method and most of his other books (p. 11). According to Oberheim (2016, p. 23) Feyerabend defends his infamous epistemological anarchism, according to which there is no fixed scientific method (‘anything goes’), by citing Einstein: … the external conditions, which are set for [the scientist] by the facts of experience, do not permit him to let himself be too much restricted in the construction of his conceptual world by the adherence to an epistemological system. He therefore must appear to the systematic epistemologist as a type of unscrupulous opportunist (Einstein, 1949, p.  684). Cited by Feyerabend (1975a), pp. 10–11.

Finally, Oberheim (2016) concluded that Einstein by emphasizing underdetermination of theory by evidence, rational disagreement, and that empirical evidence is not always neutral with respect to competing theories, anticipated substantial components of Feyerabend’s controversial views.

2.8  Feyerabend and Perspectivism According to Giere (2016), Feyerabend himself seems never to have used the term “perspectivism” to designate a philosophical position. Despite this, in retrospect, Giere thinks that Feyerabend’s views can be characterized as perspectival and hence the need to explore how his work can contribute to current thinking about perspectivism. What is perspectivism? According to Giere (2006a): “… scientific knowledge claims are perspectival rather than absolutely objective. It follows almost immediately that some contingency is always present in any science. That human observation is perspectival, a function of an interaction between the world and human cognitive capacities seems to me indisputable … Looking back historically, we can examine and understand the perspectival nature of earlier theoretical perspectives” (p. 93). Historical contingency refers to the order in which events take place that can determine which of two observationally equivalent scientific theories is accepted by the scientific community (Cushing, 1989, one example being the Copenhagen and Bohmian interpretations of quantum mechanics). Next, it is important to understand what constitutes perspectival realism: For a perspectival realist, the strongest claims a scientist can legitimately make are of a qualified, conditional form: “According to this highly confirmed theory (or reliable instrument), the world seems to be roughly such and such.” There is no way legitimately to take the further objectivist step and declare unconditionally: “This theory (or instrument) provides us with a complete and literally correct picture of the world itself.” (Giere, 2006a, pp. 5–6).

36

2  Epistemological Anarchism and How Science Works

Indeed, this perspective comes quite close to Feyerabend’s thesis of counterinduction, when he argued that even a very well corroborated theory like that of Brownian motion can benefit from alternative hypotheses (see Chap. 7 for various other examples of counterinduction in the history of science). At this stage it is important to consider how Giere (2016) establishes a relationship between Feyerabend’s conception of “Being” and perspectivism (based primarily on Feyerabend’s, 1999b, Conquest of abundance): The concept of Being plays a key role in Feyerabend’s exposition, sometimes referred to as “Basic Reality” (215), “Ultimate Reality” (215), or, more simply, just as “Reality” or “Nature.” The connection to perspectivism is direct because, for Feyerabend, “Being as it is, independently of any kind of approach, can never be known.” (205). For “kind of approach,” read “perspective.” So, in my terminology, Feyerabend is a perspectivist. Furthermore, he is also a perspectival realist (Giere, 2016, p. 138; the material in double quotes is from Feyerabend, 1999b).

Giere (2016) drawing on his advocacy of “scientific perspectivism” has argued that the later Feyerabend (1999b) as reflected in Conquest of abundance is not only a perspectivist but also a perspectival realist, as he recognizes the plurality and permeability of perspectivism upon the world.

2.9  Feyerabend and Feminist Epistemology According to feminist epistemology (Sandra Harding, Donna Haraway, Evelyn Fox Keller, Helen Longino) hegemonic (masculine) power relations lead to dualisms, such as: rational-emotional, logical-intuitive, objective-subjective, and abstracted-­ holistic. Hildebrand (1998) has argued that both sides of these dualistic concepts are present in science (in agreement with Feyerabend) and should be portrayed within the science curriculum. Feminist research literature has generally endorsed Feyerabend’s philosophy of science. Longino (1990) has criticized traditional philosophy of science for having fetishized objectivity and concluded: “Feyerabend … has rejected the relevance to science of rationality or of general criteria of theory acceptance and defends a positive role for subjectivity in science” (p. 65). Although objectivity is not synonymous with truth or certainty, it has eclipsed other epistemic virtues and to be objective is often used as a synonym for scientific in both science and science education. According to Daston and Galison (2007) the history of objectivity is nothing less than the history of science itself and that there is no objectivity without subjectivity to suppress and vice versa (p. 33). Similarly, science educators have endorsed a similar epistemological stance: This reporting style [of research] masks the subjectivity of all science in a false guise of objectivity. Research that matters is motivated by deep commitments and passions to learn. Feminism insists that we acknowledge these passions and the emotional as well as intellectual lives of researchers. Researchers should make visible their personal biases, values, and commitments in reporting research (Cavazos, et al. 1998, p. 342, italics added).

2.10  Feyerabend and the Practice of Science

37

History of science shows that science in the making (i.e., how science works) is characterized by the work of researchers who are deeply committed to their passion to learn. Interpretation of data and events always has an element of bias and even perhaps prejudice. Building consensus in science is a complex process of competitive cross-validation by the peers (cf. Campbell, 1988a, 1988b).

2.10  Feyerabend and the Practice of Science An important part of Feyerabend’s epistemological anarchism is based on how science is practiced. This aspect of his philosophy of science has been recognized by both modern philosophy of science and scientific practice itself. Kidd (2016) a philosopher of science has argued cogently that challenging methodological rules form an important part in the practice of science: Despite the entrenched perception of Feyerabend as a critic of science, closer attention to his writings, and due disregard for his rhetoric shows quite the opposite. Central to his project is a deep faith in science as a source of cognitive and other goods for the amelioration of life. That is why it matters to challenge methodological rules that would stultify inquiry and to urge deep sensitivity to the history and practice of science and to defend a thoroughgoing pluralism that checks the dangers of dogmatism. Science matters because, if conceived and pursued properly, it can serve the human good— a theme made fully explicit in the neglected, early paper, ‘Science, Freedom, and the Good Life’ (Feyerabend, 1968) (pp. 9–10, italics added).

At this stage it is interesting to compare this vision of Feyerabend from a philosopher of science with that of a practicing scientist (Roald Hoffmann, 2012, p. 36)— similarities between the two are striking. Hoffmann (2012), Nobel Laureate in chemistry and also an active researcher in various fields of science, including philosophy of science has recognized the importance of Feyerabend’s epistemological anarchism for scientific practice in the following terms: I want to claim that people are unlikely to make the new (art, science, religion, new people) without violating categories. I am here beyond philosophical bricolage, close to Feyerabend’s prescription for “epistemological anarchism.” Must be what too much poetry does (p. 36).

Interestingly, Hoffmann endorses Feyerabend’s epistemological anarchism not only for scientific practice but a wide range of activities related to human endeavor. Based on these considerations I asked Roald Hoffmann to further elaborate his understanding of epistemological anarchism, and he responded in the following terms: I see Feyerabend, whom I sadly never met, as a malevolent genius, intent on destruction of method. But underneath I sense in him an admiration for science, for the knowledge we gain, by hook or by crook. I think of epistemological anarchism as the expression Feyerabend uses for the way science really works, namely

38

2  Epistemological Anarchism and How Science Works

(a) that there is [are] no general rules governing scientific method, (b) that there may be some protocols for gaining reliable knowledge (often learned by copying the attitudes of one’s elders) in a given field, but that these are not necessarily recognized as foolproof or reliable by neighboring communities (even ones as close as chemical specialties – organic, physical).

A metaphor I have found useful recently is of scrabblers after knowledge. I mention it in the attached Tensions of Scientific Storytelling (Hoffmann, 2016, as part of an email sent to me dated 23 February, 2016). It is important to note that Hoffmann emphasizes the relationship between epistemological anarchism and how “science really works.” The metaphor of “scrabblers after knowledge” is particularly helpful to understand the scientific endeavor.

Chapter 3

Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal Science & Education (Springer)

3.1  Method The journal Science & Education (Springer, http://www.springer.com/11191) started publishing in 1992 with Michael R.  Matthews (University of New South Wales, Australia) as its Editor. This journal specifically deals with the contributions of history, philosophy and sociology of science to science education, and is indexed in the Social Sciences Citation Index (Thomson-Reuter). Consequently, it seems that an evaluation of literature published in this journal related to epistemological anarchism can help science educators to better understand the evolving nature of Feyerabend’s philosophy of science. In September 2017, I made an online literature search on the website of Science & Education, with the key words “epistemological anarchism” and “Feyerabend” (http://www.springer.com/11191). This gave a total of 102 articles published between 1992 and September 2017. All articles were downloaded and a preliminary examination showed that 24 articles could not be included in the study due to the following reasons: (a) Book reviews in which the reviewer refers to the subject (with no elaboration) and not the original author; (b) Book notes, for the same reason as for book reviews; (c) Golden oldies, which included articles by famous historians/philosophers of science with only a simple mention of Feyerabend; and (d) In some articles the authors provided a reference and the word “epistemological anarchism or Feyerabend” appeared in the title of that reference.

© Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_3

39

40

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

3.1.1  Grounded Theory Grounded theory (Glaser & Strauss, 1967) provides a set of guidelines that helps to focus on data collection procedures, based on successive levels of data analysis and conceptual understanding. In the present study, I first classified the selected articles from Science & Education in different levels (details are presented below), which were later assigned a category, and finally in Chap. 8, categories from different studies (Chaps. 3, 4, 5, 6, and 7) are compared and synthesized to facilitate conceptual understanding. This procedure can be summarized in the following steps: (a) Comparison of data sources (articles) to assign a level (I to V); (b) comparison of these levels (presented later) which facilitated their classification in categories; and (c) comparison of categories from different studies to facilitate understanding and draw conclusions. Following guidelines were used while developing the different steps of the procedure (based in part on Charmaz, 2005, p. 528): 1 . Familiarity with the setting and topic of study in each of the selected articles. 2. Evaluate classification of the selected articles to see if they are based on appropriate evidence. 3. Systematic comparisons between the classifications and the categories. 4. The need for the categories to represent a wide range of experiences represented in the classifications. 5. Establish a conceptual link between the classifications, categories and arguments for the analyses. Although the guidelines presented above were of considerable help in different stages of data analysis, a word of caution is necessary: “… grounded theory does not refer to some special order of theorizing per se. Rather, it seeks to capture some general principles of analysis, describing heuristic strategies that apply to any social inquiry independent of the particular kinds of data: indeed it applies to the exploratory analysis of quantitative data as much as it does to qualitative inquiry” (Atkinson & Delamont, 2005, p.  833, italics added). The emphasis on heuristic strategies is particularly important in the present study, as they facilitated conceptual understanding.

3.1.2  Classification of Articles Finally, a total of 78 articles were evaluated based on the following criteria: Level I Traditional view of Feyerabend as an anti-realist, and postmodern philosopher of science who espoused “anything goes”. Level II Recognition of Feyerabend’s criticism of a unitary/single scientific method and other contributions (e.g., physical concept, scientific progress, pseudoscience, naïve falsificationism, indoctrination, instsrumentalism, incommensurability, and objectivity).

3.2  Results and Discussion

41

Level III Criticism and controversy with respect to the social responsibility of the scientists, scientism, dogmatism, incommensurability, genius in science experimental verification of a theory, theory-laden nature of data and the need to go beyond the scientific method. Level IV Different perspectives on a subject (such as anything goes) reflect also diversity rather than a Unitarian view of science. It implies the absence of an algorithm rather than the absence of methods. Level V Interaction and competition between opposing theories leads to a flexible and hence anarchic epistemological perspective. Breaking rules in scientific practice means transgression of categorization, and counterinduction, which leads to a plurality of perspectives—hence proliferation of theories (methodological pluralism). Following the guidelines presented above (cf. Charmaz, 2005), and in order to facilitate credibility, transferability, dependability and confirmability of the results I adopted the following procedure: (a) All the 78 articles from Science & Education were evaluated and classified in one of the five levels; (b) After a period of approximately 3 months all the articles were evaluated again and there was an agreement of 91% between the first and the second evaluation; and (c) After another period of 3 months all the articles were evaluated again, and there was an agreement of 94% between the second and the third evaluation. This procedure was particularly helpful in understanding the underlying issues as according to Denzin and Lincoln (2005): “Terms such as credibility, transferability, dependability, and confirmability replace the usual positivist criteria of internal and external validity, reliability, and objectivity” (p. 24, original italics). A complete list of all the 78 articles from Science & Education that were evaluated is presented in Appendix 1. In the section on Results and Discussion, examples of the different levels (I–V) are provided. These examples provide an understanding of how the subject of epistemological anarchism has been discussed by authors in this journal. It is important to note that all the articles evaluated in this study referred to epistemological anarchism (or Feyerabend) in some context, which may not have been the primary or major subject dealt with by the authors. Distribution of all the 78 articles according to author’s area of research, context of the study and level (classification) is presented in Appendix 2.

3.2  Results and Discussion Each of the 78 articles from Science & Education was evaluated (Levels I–V) with respect to the context in which they referred to epistemological anarchism. Based on the treatment of the subject by the authors 17 categories were developed to report and discuss the results (cf. guidelines presented above from Charmaz, 2005). These categories along with examples are presented in alphabetical order. It is important to note that some of the articles could easily be placed in more than one category.

42

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

The idea behind the creation of 17 categories is to facilitate the reader to find the subject of her/his interest. Science & Education has a readership and contributors that include science educators, historians, philosophers of science and sociologists that cover many areas of expertise in the science curriculum. Given the wide range of subjects discussed by the authors over a period of almost 25 years, it is difficult to create the semblance of a continuous storyline. Complete information about each article and the author is provided in Appendices (1 and 2), which can be consulted by the interested readers. Following categories (n = 17) with detailed examples of all five levels are presented (in alphabetical order) in the next section: 1. Acid-base equilibria 2. Anarchistic methodology 3. Constructivism 4. Critical thinking 5. Diversity of methods 6. Enlightenment 7. Evolution, knowledge and belief (to give meaning to life) 8. Gender 9. Incommensurability 10. Nature of science 11. Newtonian method 12. Normal science, dogmatism and science education 13. Polanyi’s tacit knowledge 14. Science and religion 15. Scientific expertise and Galileo 16. Scientific method 17. Situated learning

3.2.1  Acid-Base Equilibria Acid-base reactions and the resulting equilibria form an important part of high school and introductory university level courses. Based on his theory of electrolytic dissociation, in 1887 S.  Arrhenius presented his model of acid-base reactions. Electrolytic dissociation at that time was not well received by the scientific community (for a review see De Berg, 2003). According to Arrhenius, an acid produces H+ and a base produces OH− when dissolved in water. Arrhenius (1912) used the idea of ionic dissociation to explain the similar values for the heats of neutralization that hydrochloric acid or nitric acid gave when mixed with sodium hydroxide, and thus concluded that each acid-base reaction was equivalent to the same ionic reaction: H+ + OH− ➔ H2O. A major shortcoming of the Arrhenius theory is that many substances yield OH− when they dissolve in water, although they do not contain OH in their formula, e.g., ammonia: NH3(g) + H2O (l) ↔ NH4+ (aq) + OH− (aq). Another

3.2  Results and Discussion

43

limitation of the Arrhenius model was that water had to be the solvent for acid-base reactions. A different conceptualization of acids and bases was presented in 1923 by J.N. Brønsted and T.M. Lowry, according to which an acid-base reaction is a proton transfer process. One species donates a proton and another species accepts it. An acid is a proton donor, that donates a H+ ion, e.g., HNO3. All Arrhenius acids are Brønsted-Lowry acids. A base is a proton acceptor, any species that accepts an H+ ion. However, a base must contain a lone pair of electrons to bind the H+ ion, and following are some examples: NH3, CO32−, and F−. Brønsted-Lowry bases are not Arrhenius bases, but all Arrhenius bases contain the Brønsted-Lowry base OH−. In 1923 G.N.  Lewis presented a much broader conceptualization of acid-base reactions by postulating that a base is a substance that donates an electron pair and an acid is a substance that accepts an electron pair (e.g., B: + H+ ↔ B-H+). The Lewis model, like the Brønsted-Lowry model, requires that a base have an electron pair to donate, so it does not expand the class of bases. However, it greatly expands the class of acids. Many species such as CO2 and Cu2+, do not have H and still function as Lewis acids. Based on this historical reconstruction of the acid-base conceptualization, Kousathana, Demerouti, and Tsaparlis (2005) consider that it illustrates the tentative nature of science and concluded: Some of the central questions in philosophy of science such as the distinctive features of science that set it apart from other endeavors have been traditionally addressed in terms of what is considered to be the paradigm science: physics. Even though the emphasis on the logical analysis of scientific theories has been challenged by philosophers, such as Popper, Kuhn, Feyerabend, Lakatos, and other more recent philosophers, the legacy of logical positivism as “physics” dominating in philosophical analysis persists even today … (p.  175, italics added). (Classified as Level III).

The legacy of logical positivism, would of course, require that the topic of acid-base equilibria (models of Arrhenius, Brønsted-Lowry and Lewis) should preferably facilitate students’ understanding that of the three models only “one” is “correct” or “true”. It is precisely under such circumstances (historical episodes) that Feyerabend is critical of not only logical positivism but even some of the recent philosophers of science, as this precisely leads to epistemological pluralism (viz., more than one “correct” model). As a follow-up of the study by Kousathana et  al. (2005), a similar study was designed by Niño (2009) to evaluate 37 general chemistry textbooks (mostly published in U.S.A., between 1950s and 2000s, see Appendix 9). A major assumption of the study was based on the Lakatosian thesis that empirical evidence does not necessarily refute a theory but rather leads to rival and competing theories (Lakatos, 1970). There is a fair amount of agreement between Feyerabend and Lakatos on this point. Of the 37 textbooks evaluated by Niño, almost all dealt with the three acid-­ base models. Textbooks were evaluated on five criteria based on history and philosophy of science, and one of the criterion dealt with the tentative nature of science, as suggested by Kousathana et al. (2005). Seven of the textbooks presented a satisfactory perspective of the three acid-base models with respect to the tentative nature of science and following is one example:

44

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported… The classical (Arrhenius) definition, the first attempt at observing acids and bases on the molecular level, is the most limited and narrow of the three definitions. It applies only to species that contain H or OH groups in their structure, which are released as ions when the species dissolve in water… The Brønsted-Lowry definition is more general, seeing acid-­ base reactions as proton transfer processes and eliminating the requirement that they occur in water … The Lewis definition has the widest scope of the three. The defining event of an acid-base reaction is the donation and acceptance of an electron pair to form a new covalent bond. Lewis bases still must have an electron pair to donate, but Lewis acids—electron-pair acceptors—can be very different from the species encompassed by the other two definitions, including electron-deficient compounds, compounds with polar double bonds, metal ions, and the proton itself (Silberberg, 2000, p. 794, original italics).

This comparative overview of the three acid-base models helps to understand the changing nature of our understanding, and hence the tentative nature of science. Besides other aspects, it emphasizes the increasing generality of the models. Furthermore, although the heuristic power of the models increases progressively, still depending on the context all three can help understand different aspects of the acid-base equilibria. This, of course, leads to the logical positivist perspective: which model is “correct” or “true”? None of the textbooks evaluated by Niño (2009) referred to this issue. The quest for an answer to this question led us to one general chemistry textbook that provided a plausible scenario: What are the main characteristics of the molecules and ions that exhibit acid and base behavior? In this chapter, we examine three different definitions: the Arrhenius definition, the Brønsted-Lowry definition, and the Lewis definition. Why are there three definitions, and which one is correct? … there really is no “single” correct definition. Rather, different definitions are convenient in different situations” (Tro, 2008, p. 666).

Does the scenario provided by Tro comply with the scientific method and is it appropriate for discussion in introductory university level courses and textbooks. Most chemistry teachers and textbooks will have considerable difficulty in accepting Tro’s scenario. Epistemological anarchism would of course endorse Tro’s scenario. Another example that leads to a similar scenario is the one provided by the valence bond and molecular orbital models of covalent bonding (see Niaz, 2016 for details). After reading a preliminary version of this chapter, Geelan (2019) commented: “The discussion of the three models of acid-base is great, but there may be value in addressing the word ‘model’: It leads to a similar conclusion, that there is no single true acid-base theory but different ones serve our interests in different contexts. It has the advantage that it can be taught to students right from primary/ elementary school. Science creates models of the world that allow us to serve human interests, rather than certain knowledge. It avoids the problem where a Bohr model of orbits is taught as though true then supplanted by a quantum model with orbitals which is now taught as though true. Teaching each model ‘as’ a model is more powerful.”

3.2  Results and Discussion

45

3.2.2  Anarchistic Methodology In a section entitled “Epistemology and Nature of Science”, Kalman (2009b) first compares and contrasts the philosophies of Bacon, Popper, Kuhn, Lakatos and Feyerabend. According to this historical perspective, discovery learning is basically Baconian due to its emphasis on observation and induction. The idea that regularity in nature yields a law is Baconian. Kalman (2009b) provides various examples of such laws, such as Kepler’s discovery of Bode’s law and Balmer’s law for the visible hydrogen line spectrum. Popper’s criterion for a successful scientific theory is that besides passing experimental tests, it could also be subjected to falsification. Feyerabend would counter by pointing out that strictly speaking no theory can be falsified, as it can be resurrected by invoking new auxiliary hypotheses. In Kuhn’s view a mature science has only one paradigm, whereas Feyerabend would counter by suggesting that it is desirable to have as many competing theories as possible. For example, Michelson-Morley experiment was conducted in 1887 in order to distinguish between the rival ether theories of Stokes and Fresnel. Lakatos, on the other hand, differentiates between progressive and degenerating research programs. Feyerabend has countered by suggesting that a degenerating program often takes a progressive turn and in the process stimulates other programs to move forward. According to Kalman, based on this critical appraisal of Bacon, Popper, Kuhn, Lakatos and Feyerabend, some of these difficulties can be overcome in epistemological anarchism: Feyerabend’s anarchistic methodology has one simple rule: there is no rule. The successful and creative scientist breaks rules, reverses rules, defends ad hoc hypotheses, works inductively then deductively, and works sometimes for unity and sometimes for plurality i.e. anything goes. Moreover, such efforts are liable to produce the kind of plurality of theories that Feyerabend considers to be essential for maximum progress in the sciences (Kalman, 2009b, p. 335). (Classified as Level V).

It is important to note that Feyerabend’s famous “anything goes” has to be understood in the context of what the scientist is trying to resolve and the methodological strategy is not decided a priori (also see Chap. 2, for a recent appraisal of Feyerabend by philosophers of science). After considering the critical analyses of philosophers of science, with respect to the role of observation as a source of reliable knowledge, the validity of inductive and deductive strategies, the existence of a single scientific method, the status of scientific knowledge as truth or infallible, and the superiority of a theoretical model over others, Guerra-Ramos (2012) concluded: “By the middle 1970s, Paul Feyerabend proposed what came to be known as ‘epistemological anarchism’, defending that there is not such a thing as scientific method and the provocative idea that in scientific enquiry ‘anything goes’” (p.  649). Such considerations led Feyerabend to propose a diversity of perspectives among researchers rather than a Unitarian view of science. (Classified as Level IV).

46

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

3.2.3  Constructivism Geelan (1997) has explored the dialectical tension between the various forms of constructivism, such as: Personal constructivism (G.  Kelly & J.  Piaget), Radical constructivism (von Glasersfeld), Social constructivism (J. Solomon), Social constructionism (K. Gergen), Critical constructivism (P. Taylor) and Contextual constructivism (W. Cobern). During the 1990s these different forms of constructivism were the source of considerable controversy in the science education community. After a critical appraisal of the various forms of constructivism espoused by science educators, Geelan (1997) concluded: Opposing theories do not damage or supplant one another, in Feyerabend’s view: they are necessary to one another, since their dialectical interaction throws each theory into sharper focus, making it more useful and powerful … the flexible, anarchic, context-sensitive use of a variety of selected epistemological perspectives …, is the most powerful theoretical ‘engine’ which can be used for the development of educational theory and practice. The search for a single best approach is self-defeating: the most important effect of choosing any single perspective is that it blinds us to an enormous range of other valuable possibilities (pp. 26-27, italics in the original, underline added). (Classified as Level V).

As an example of dialectical tension Geelan cites the example of the controversy between Fosnot (1993), who espoused cognitive constructivism based on Piaget’s model of equilibration, and O’Loughlin (1992) who espoused sociocultural and contextual factors (social constructivism). Geelan (1997) recommended that such differences, “… can be a great source of creativity and productivity if it is not allowed to deteriorate into antagonism and confrontation” (p. 22). Precisely, according to Feyerabend such opposing theories are necessary to advance our knowledge. In the case of constructivism, there are two rival methodological approaches, namely the internalist and the externalist research programs. The former (e.g., Piaget) would emphasize the endogenous development of intellectual structures, whereas the latter (e.g., social constructivists) analyzed the social, economic and cultural changes. Actually, Fosnot (1993) herself recommends that science educators need to hear the story from both sides (for further details on the Fosnot-O’Loughlin controversy see Niaz, 2018, Chap. 4). Furthermore, given the uncertainty and tentative nature of scientific knowledge, Geelan goes on to caution that, “the search for a single best approach is self-defeating” and avoiding such a search may even avoid needless controversies (italics suggested by Geelan, 2019). Such plurality of models and perspectives is supported not only by Feyerabend but also Giere (2006a, b) and Kellert, Longino and Waters (2006). In a similar vein, drawing on the history of science Niaz (2011b, Chap. 11) has drawn an analogy between the progress in atomic structure (science) and educational practice (constructivism). History of atomic structure in the twentieth century is based on a series of atomic models that developed by including some aspect of the earlier models, that is, dialectical tension (e.g., atomic models of Thomson, Rutherford, Bohr, Bohr-Sommerfeld, and wave mechanical). Feyerabend’s thesis of counterinduction is particularly helpful in understanding these atomic models (for details see Chap. 7). Similarly, ­constructivism

3.2  Results and Discussion

47

has evolved through a sequence of models based on the work of various scholars (e.g., Piaget, Ausubel, von Glasersfeld, Vygotsky and Perkins). Continuing this line of reasoning, Niaz et al. (2003) have argued that constructivism in science education (like any scientific theory) will continue to progress and evolve through continued critical appraisals (based on various aspects of the different forms of constructivism). The science education community has generally ignored the history and philosophical issues involved with scientific progress. Publication of the journal Science & Education has redressed this problem to some extent. However, the social and cultural milieu still seems to be outside the purview of most science educators. In this context while analyzing radical social constructivism, Cobern (1995) has pointed out that, “… my immediate concern is that science educators will come to associate the social and cultural study of science education with an anything goes perspective reminiscent of Feyerabend (1993). It simply is not necessary to equate cultural influences on the process of constructing meaningful knowledge with relativism” (p. 290, original italics). (Classified as Level IV). Clarification of anything goes notwithstanding, Cobern’s concern is well placed as in a certain sense Feyerabend did sympathize with some form of relativism (see Chap. 2 for details on changes in Feyerabend’s thinking on this issue). Radical constructivism is generally criticized for claiming that as all knowledge is constructed, and consequently it does not tell us anything certain about the world. Ernest (1993) has claimed that this criticism does not lead to new implications for education beyond those of trivial constructivism. Furthermore, based on Popper any good modern scientific theory must admit that it is refutable. Based on this perspective, Ernest (1993) concluded: There is a long sceptical tradition in philosophy which denies the existence of certain knowledge, which in the modern era finds expression in the work of Dewey and the pragmatists; Wittgenstein, Putnam, Rorty; and which constitutes a central tradition in the philosophy of science, in the work of Kuhn, Feyerabend, Toulmin and Lakatos … Thus the denial of certain knowledge is far from unique to radical constructivism (p. 91). (Classified as Level III).

Uncertain nature of scientific knowledge can of course become an obstacle, especially for students. More recently, Giere (2010) has argued that knowledge claims are perspectival rather than absolutely objective and hence cannot provide a “true” or “correct” answer to a problem, which leads to a pluralism of perspectives. After reading a preliminary version of this chapter, Geelan (2019) suggested: “I would agree that the ‘radical’ part of radical constructivism is philosophically interesting but bootless for science education. It doesn’t really imply doing anything differently. I do very much like von Glasersfeld’s image of ‘finding our way out of a dark forest blindfolded’: the physical world will allow certain pathways and block others, but finding one does not mean it is the only possible pathway.” Radical constructivism is frequently criticized for its relativist stance of knowledge, and that may lead some to suggest that both physics and astrology would offer equally good descriptions of the interactions between certain celestial bodies and the Earth. According to Quale (2007) this leads to an extreme epistemic thesis of

48

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

solipsism, namely the idea that we are all free as individuals to invent our knowledge of the world. However, this criticism can be countered on the grounds that the “invented” knowledge needs to be shared and experienced with others and for this interaction to take place, in the first place knowledge has to be constructed. In this context Quale (2007) draws attention to Feyerabend’s dictum “anything goes”, which has been understood by some as suggesting that all knowledge is equally good and can also be interpreted as: … asserting that science is a very complex and difficult subject matter to deal with, and that one should therefore feel free to try any and all approaches that may seem promising, in order to generate scientific knowledge—but then with the clear understanding that many (perhaps most) such attempts may not in fact succeed. In other words: one should not be dogmatic about what is or is not permissible in the activity of scientific research; one should feel free to do one’s own thing, but then be prepared to find that this may not lead to any useful results for science. (Indeed, later statements by Feyerabend indicate that such an interpretation may be more in line with what he meant: a plea for methodological pluralism—an attitude that resonates well with his own designation of himself as an “epistemological anarchist”.) (p. 243, original italics). (Classified as Level V).

Actually, Quale goes beyond by suggesting that this methodological pluralism may even lead to an irreverent attitude that may help us to understand that we need not be complacent of the “truth” of our existing theories but rather look for major breakthroughs in the evolution of science. It is plausible to suggest that irreverent attitude could mean “transgression of categorization” and major breakthrough approximates to “breaking rules.” Furthermore, Quale’s understanding of Feyerabend comes quite close to that of recent philosophy of science (for details see Chap. 2). In this context it is important to note that R. Hoffmann (2012, p. 36), a working chemist, considers that Feyerabend’s epistemological anarchism approximates to “violating categories” (that is “transgression of categorization”), and that at first he was “torn” about this as he was drunk on logic, mathematics and symmetry (for further details on Hoffmann’s views, see Niaz, 2018, Chap. 6).

3.2.4  Critical Thinking Chang (2011) has recognized the importance of critical thinking for both liberal education and the training of scientists: So the aim of science education must be considered within the larger framework of general or liberal education, and critical thinking is vital in that context. Secondly, even for the training of normal scientists, there is a strong argument for making a place for critical inquiry. There are the old arguments by Karl Popper, John Watkins and Paul Feyerabend (see papers in Lakatos and Musgrave 1970) against the Kuhnian line on the necessity of dogma in scientific training (p. 336). (Classified as Level III).

Indeed, Feyerabend was particularly concerned about the role of dogma in scientific progress, and how that stifles future research.

3.2  Results and Discussion

49

Kalman (2002) has designed undergraduate physics courses in order to facilitate conceptual understanding through critical thinking (for further details see Kalman, 2017), which is in turn based on the work of Feyerabend (1975a). Feyerabend emphasized the importance of a pluralistic methodology in which theories are critically compared with other theories and not with ‘data’ or ‘facts.’ Interestingly in his early work Galileo was a follower of Ptolemy and argued against the motion of the earth. Based on a reconstruction of Feyerabend’s experience of Galileo, Kalman (2002) concluded: “Only later did he [Galileo] take the Copernican point of view and assume that the stone dropped from the mast [of a rapidly moving ship] would fall toward its foot. Feyerabend’s conclusion is that Galileo only arrives at the modern theory of inertia by a critical examination of the tower experiment in the light of two alternative frameworks; that of Ptolemy and that of Copernicus … The idea that examining alternatives will enhance critical thinking skills and help to produce conceptual change is put to test …” (p. 86). (Classified as Level V). It is plausible to suggest that this approximates to Feyerabend’s thesis of counterinduction (for details see Chap. 7). This central idea helped Kalman to facilitate conceptual change in physics courses and students came to understand that, “… older ‘replaced’ theories were not necessarily ‘bad’ and that presently accepted theories may very well change” (Kalman, 2002, p. 91). Furthermore, this can help students to understand the tentative nature of science.

3.2.5  Diversity of Methods According to Park, Nielsen and Woodruff (2014): Science allows for variations of research methods depending on subject matter and individual scientists. However, almost all scientists observe, compare, measure, test, speculate, hypothesize, create ideas and conceptual tools, and construct theories and explanations. For philosophers of science, methods of science are not so straightforward. Philosophers classify methods into inductivism, deductivism, hypothetico-deductivism, and relativism … No single sequence of prescribed activities will guarantee for philosophers or scientists the discovery of new theories or solutions. So some argue that no specific method can be universally applied and may suggest a case that “anything goes” (Feyerabend, 1975a, p. 28). (Classified as Level IV).

This presentation first outlines the need in science to use a series of different methods (depending on the problem to be solved) and then draws attention to the requirements of philosophers of science. A transaction between the different perspectives leads to an understanding of Feynerabend’s “anything goes” and hence the diversity of methods.

50

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

3.2.6  Enlightenment In a critical appraisal of the role of worldviews in understanding nature of science, Matthews (2009) has recognized that: The traditional Enlightenment position was one that embraced and defended the methodologies and claims of science. Some wish to relax this connection and have the personal, cultural and political elements of the Enlightenment without a robust commitment to science (Feyerabend, 1978); while others want science without these personal, cultural and political positions (most rulers and technocrats in authoritarian cultures such as China, Saudi Arabia, Iran, Pakistan and so on). These anti-science, and anti-Enlightenment sentiments have, unfortunately, been widely embraced in science education circles … (p. 644). Classified as Level III.

Matthews has referred to various issues which are a cause of tension in science education. First, recent research in philosophy of science (see Chap. 2) has shown that Feyerabend was not anti-science but rather tried to elucidate how science really works. Second, the political philosophy of authoritarian regimes was a cause of considerable concern to Feyerabend. With respect to the acceptance of postmodernism in some science education circles, Feyerabend’s later work has a more nuanced perspective of scientific progress. Kalman (2019a) provides further insight with respect to Feyerabend’s style of thiking, “He certainly was not anti-science. Feyerabend cannot always be taken at face value. He is always provocative and often makes seemingly absurd statements to challenge us.”

3.2.7  E  volution, Knowledge and Belief (to Give Meaning to Life) Mugaloglu (2014) has explored the relationship between constructivist pedagogy and pseudoscientific ideas (e.g., Intelligent Design, ID) while teaching evolution: “The radical postmodern view holds that knowledge can never be more than just a story … Feyerabend highlights the limits and irrationalities in science and discusses that ‘anything goes’. Knowledge claims are socially embedded and depend on culture” (p. 831). Next, Mugaloglu reports the results of a study in which prospective science teachers at a state university in Turkey were asked the following question: “Would you teach ID and/or evolution theory if there were no curriculum restrictions on what to teach?” (p. 836). For anyone familiar with the present sociopolitical situation in Turkey this is a very difficult and controversial question. Of the 48 respondents, 19 declared that they would teach both ID and evolution, two opted for teaching only ID and the remaining decided in favor of evolution. This shows that at least 21 teachers considered ID to be a legitimate topic in a science class and following is one example of a teacher’s response: I would like to teach both evolution and ID theories in class … There is no debate that the best explanation scientist[s] have for maintaining life is evolutionary theory. On the other

3.2  Results and Discussion

51

hand, ID is proposed by other scientists as well … Teaching only one perspective would cause students to accept everything that they were told and soon they will stop thinking. I would like my future students to give meaning to life with science … That’s why I will try to do my best to teach both theories so that they have different ideas (Reproduced in Mugaloglu, 2014, p. 836, italics added). (Classified as Level II).

This example clearly shows that in such controversial issues the role played by culture as envisaged by Feyerabend is important (for Feyerabend, teaching only one theory could lead to indoctrination. Debate between ID and evolution would provide the setting that approximates to what Feyerabend considered as counterinduction (for details see Chap. 7). Furthermore, this has pedagogical implications if we consider the teacher’s concern with respect to, “soon they will stop thinking”. At this stage it is important to note that the ID/evolution controversy has been the subject of considerable controversy in the science education literature. Centrality of Darwinian theory to biological thought has been recognized in the literature (Dobzhansky, 1973; Gould, 1977; Mayr, 1982; Ruse, 2013). However, despite this recognition science educators (in USA and other parts of the world) continue to face considerable difficulties in teaching biological evolution to students from orthodox (Christian, Jewish, Islamic & others) background. In this context it is important to consider Kalman’s (2013) advice that science and religion are totally independent subjects. He then refers to Maimonides who suggested that to be a scientist and to find scientific reasons for everything in the world is important. However, none of that negates the existence of God. Nevertheless, we are still faced with the dilemma of a teacher while teaching evolution as some students may want to compare it with intelligent design (ID), or some other similar idea. Most teachers would perhaps agree that the goal for students is to acquire knowledge about evolution. Cobern (1994) considers this answer as simplistic and understandable only within a scientistic view of science, which is a myth in school science: The myth is a scientistic view roughly embracing classical realism, philosophical materialism, strict objectivity, and hypothetico-deductive method. Though recognizing the tentative nature of all scientific knowledge, scientism imbues scientific knowledge with a Laplacian certainty denied all other disciplines, thus giving science an a priori status in the intellectual world. The certainty of scientism can make life easy for the science teacher. Scientism allows the teacher to say to students that this is the way things are, for science provides the one reliable source of objective knowledge (p. 585, original italics).

Cobern’s concern lies with the fact that although students may seem to understand evolution they generally do not believe in it. In other words we are faced with the dilemma: Science educators need to facilitate conceptual understanding and/or persuasion for belief. Furthermore, as learning takes place in a social context, controversial topics like evolution cannot ignore the significance of the cultural milieu. Interestingly, Cobern suggests that the issue of belief cannot be ignored, and that belief is the place where instruction should begin (p. 587). Finally, Cobern (1994) states, “Today’s teacher of evolution faces a situation very much like Darwin presenting the Origin of Species to a public that historically held a very different view of origins” (pp. 587–588). This scenario is crucial in teaching not only evolution but

52

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

also all controversial topics of the science curriculum. Leon Cooper (1992), Nobel Laureate in physics, has provided cogent advice to solve the dilemma: “A question often very puzzling to students is why such a thing was done at such a time. Frequently, the answer can only be given in the milieu of the time—the problems that seemed important, the opinions of the people involved” (p. xii, Preface, emphasis added). Cooper goes beyond by suggesting that if the Michelson-Morley experiment (late nineteenth century) had been done at the time of Copernicus (sixteenth century), its result would have no significance for the astronomers, as they considered the earth to stand still and at the center of the universe. It seems that reference to milieu of the time can help to facilitate a better understanding of the beliefs of students in a particular topic. At this stage it is important to note that students’ beliefs are closely enmeshed with their alternative conceptions of a topic, which have been investigated intensively. Smith and Siegel (2004) have criticized Cobern’s (1994) approach to teaching evolution that focuses on belief in evolution, on the grounds that students may understand the term belief as synonymous with faith, opinion or conviction. Furthermore, it may lead the students to understand that accepting evolution is a matter of personal faith that has no evidential basis. Actually, the role of evidential basis based on empirical evidence is controversial even in the history of science. For example, J.J. Thomson and E. Rutherford had very similar experimental evidence (alpha particle experiments) and still there interpretations were entirely different and in part based on their prior beliefs, theories, models or theoretical frameworks (for details on this and other historical episodes see Niaz, 2012, 2016). Prior theoretical beliefs play a crucial role in scientific progress and the controversy between Thomson and Rutherford lasted for many years although they were well known to each other and could easily have met over dinner and resolved the controversy. Based on his critique with respect to teaching biology, Smith (1994) concluded: Although the distinction between believing and accepting may be a subtle one for many, it is crucial to understanding the nature of science; moreover, drawing carefully the distinction between belief (or faith) in the absence of objective evidence and acceptance that is based on evidence provides an excellent opportunity for helping students to understand what science is. In my view, in fact, the primary reason for including evolution in the curriculum, other than the obvious value of a meaningful understanding as a basis for understanding the rest of biology, is that it provides the wonderful opportunity for addressing pervasive misconceptions about the nature of science (p. 595).

More recently, Laats and Siegel (2016) have argued that a student does not need to believe in evolution in order to understand its tenets and evidence. In other words, a student can be fully literate in modern scientific thought and still maintain contrary religious or cultural views. Both Laats (a historian) and Siegel (a philosopher of science) agree that as a science creationism is flawed. However, given that creationism represents a form of religious dissent it is important to disentangle belief from knowledge. Interestingly, in the history of science even scientists (fully literate in scientific thought) can ignore experimental evidence and continue to believe in their prior theoretical frameworks. One example (Thomson versus Rutherford) of such a case was cited above. Similarly, Robert Millikan provided experimental evidence to

3.2  Results and Discussion

53

determine Planck’s constant h based on Einstein’s photoelectric equation and still rejected Einstein’s theoretical framework and continued to believe in the classical wave theory of light (for details see Chap. 7 of this book and Niaz, 2012). On reading a preliminary version of this chapter Kalman (2019a) suggested, “Consider, Einstein’s famous comment that if the experiments challenging special relativity (which eventually proved to be erroneous) were correct, he would feel sorry for nature. The conclusion is that scientists NEED to hold on to their theoretical frameworks as long as possible” (emphasis in the original). Indeed, the role played by empirical evidence in the history of science is much more complex and controversial. The historical evolution of objectivity itself as studied by Daston and Galison (2007) is a good representation of how empirical evidence was cast in different ways, depending on the epistemological orientation of the scientists involved. Interestingly, it is not only the teachers but even researchers have faced difficulties while dealing with the topic of evolution. Jackson, Doster, Meadows and Wood (1995) have reported the difficulties faced by a science educator (from a secular-­ humanist background in the northern USA), in trying to communicate biological evolutionary theory to scientists, science educators, and science teachers in the religious-influenced culture in the southern USA. This experience shows the limitations of the cognitively-oriented conceptual change theory. Instead the authors used a heuristic inquiry approach (Patton, 1990) in which an overtly personal and subjective viewpoint is acknowledged. Elaborating on the methodology used, Jackson et al. (1995) concluded: First, this topic [biological evolution] elicits strong emotional reactions in many people, including several of the researchers … In such circumstances, the use of a method which explicitly strives for objectivity is probably futile and definitely presumptuous. Second, the primary researcher/first author was conscious of a profound initial ignorance of the religious points of view on the issues raised. This situation calls for a method which anticipates and values an adaptive process by which specific research questions and methods evolve in response to data gathered and analyzed earlier in the inquiry (p.  590, italics added).

This statement clearly shows the difficulties involved in doing research on topics in which both the participants and researchers have strong prior epistemological views that produce conflicting situations in the classroom. Indeed, the authors recognize that in such studies participants need to be considered as co-researchers as they posed incisive questions that provided a stimulus to reflect and reevaluate the basic assumptions and goals of the study. In a sense these findings can be seen as the two poles of the subjectivity-objectivity interface. In other words, based on his professional training in evolutionary biology the primary researcher thinks that he is being objective and at the same time in his interactions with the participants he is forced to understand their views and hence the need for a subjective understanding. At this stage it is important to summarize the two epistemological positions: a) First according to Smith and Siegel (2004) teachers ought not to strive to shape students’ beliefs but rather only their knowledge and understanding of evolution; a contrary strategy would lead the teacher to be an indoctrinator rather than a teacher (p. 565). b) Second according to Cobern (2004) who holds a contextual, pluralistic

54

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

(non-relativistic) view of epistemology, for classrooms that are increasingly multicultural teachers need to facilitate discussions with students that hold diverse views, namely both knowledge and belief about evolution (p. 588). With this background let us consider the following response Ron Good (2001) received from a student: I have to disagree with the answers I wrote on the exam. I do not believe that some millions of years ago, a bunch of stuff blew up and from all that disorder we got this beautiful and perfect system we know as our universe … To say that the universe “just happened” or “evolved” requires more faith than to believe that God is behind the complex organization of our solar system … (p.7, italics added).

Many students in different parts of the world would perhaps agree with this student and Good considers that meaningful learning cannot take place without belief in evolution. Both Smith and Siegel (2004) and Cobern (2004) discuss this student’s statement and the latter coincides with Good to some extent. Perhaps, it is plausible to suggest that the part in italics “just happened” represents a yearning for certainty and all the more so as it deals with such an important theory as evolution. This “yearning” has been endorsed by the American Association for the Advancement of Science (AAAS, 1993b) in cogent terms: “The notion that scientific knowledge is always subject to modification can be difficult for students to grasp. It seems to oppose the certainty and truth popularly accorded to science, and runs counter to the yearning for certainty that is characteristic of most cultures, perhaps especially so among youth” (p. 5, italics added) (also see AAAS, 1993a). Interestingly, this yearning may have cultural overtones. Now let us consider the following statement (and the ensuing dilemma) from a student teacher (cited above): “Teaching only one perspective would cause students to accept everything that they were told and soon they will stop thinking. I would like my future students to give meaning to life with science … That’s why I will try to do my best to teach both theories [Evolution and Intelligent Design] so that they have different ideas (Reproduced in Mugaloglu, 2014, p. 836, italics added). To give meaning to life with science is perhaps an important part of the educational endeavor and generally ignored in most science curricula. It is plausible to suggest that Feyerabend’s perspective (counter induction, see Chaps. 2 and 7) leading to diversity and epistemological anarchism approximates to a better understanding of knowledge and belief, in the context of teaching both evolution and intelligent design. On reading a preliminary version of this chapter Kalman (2019a) commented, “I think that Feyerabend would be horrified at this idea. It is a perversion of his principle. We absolutely should not teach intelligent design. It is not science”. However, given Feyerabend’s provocative style it is important to recall that he applauded California science teachers for having included intelligent design (ID) in their curriculum. Recent research in science education has supported a similar teaching strategy: “The knowledge that has already been acquired allows the researchers to raise new questions because there is uncertainty; a given study aims to decrease this uncertainty and then new questions emerge, again pointing out new uncertainty. This dynamics of uncertainty based on knowledge is a way of ­developing

3.2  Results and Discussion

55

knowledge. We also consider that, in the students’ processes of construction of knowledge, uncertainty can drive the learning process of knowledge” (Tiberghein, Cross, & Sensevy, 2014, p. 931). This clearly shows that uncertainty with respect to scientific knowledge need not be a constraint in learning science (evolutionary theory being a good candidate), but rather can even facilitate construction of new knowledge.

3.2.8  Gender A gender plural science education has been the subject of considerable research and is also supported by Feyerabend’s post-empiricist epistemology. Based on this assumption Ginev (2008) has endorsed: A multi-gendered research process would be a cognitive reality that is envisaged in Feyerabend’s celebrated principles of counterinduction and proliferation. Clearly science education has a role in science’s gender pluralization (p. 1153). (Classified as Level III).

3.2.9  Incommensurability Based on the Duhem-Quine thesis, Lamont (2009) has referred to the difficulty involved in defining theoretical terms in an observational language, namely if the observations associated with a term change as a result of further research, then the meaning of the term changes. This in turn makes it difficult to explain how different theories can disagree, leading to the incommensurability thesis that is, a rational justification of the preference of one theory over another. Such difficulties led Feyerabend to his ideas on comparing theories: Paul Feyerabend, indeed, drew from the notion of incommensurability the conclusion that such preference cannot be justified, and denounced the imperialism of scientists who claimed that their theories were in any way superior to the views of believers of witchcraft (Lamont, 2009, p. 878). (Classified as Level III).

The reference to “witchcraft” is puzzling in this context. However, most philosophers of science would attribute it to Feyerabend’s rhetorical style. Drago (1994) has traced the origin of the debate between Newtonian mechanics and thermodynamics, especially during the last decades of the nineteenth century, and argued that the two are incommensurable because of the kind of mathematics used. Mach considered mechanics to be a very abstract and idealistic theory, in which time is a reversible parameter. In contrast, thermodynamics is an empirical and operative theory in which time is irreversible. Furthermore, Mach considered thermodynamics to be more basic for theoretical physics than mechanics. Mach’s

56

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

thesis was also supported by some of the leading scientists, such as Duhem, Ostwald, and Planck. Based on these considerations, Drago (1994) concluded: We have to remember the suggestion by Feyerabend and Kuhn, that is, two physical theories may be incommensurable. In such a case almost any notion belonging to a theory is radically different from the corresponding notion, if any, belonging to the other theory. Really, the traditional teaching, which starts from mechanics as first theory, manifests an incommensurability when it is not able to introduce thermodynamics by means of its basic notions (p. 197, original italics). (Classified as Level IV).

This leads to the recommendation that physics teaching should not monopolize theoretical physics under only one theory, namely mechanics or thermodynamics, but rather use a plural (instead of Unitarian) approach, in which two strategies are possible. This coincides with Feyerabend’s advice with respect to following diversity, rather than a Unitarian approach. According to Bunge (2003) because of partial overlap, quantum and classical physics are not incommensurable or incomparable. For example, when calculating the spectra of atoms immersed in an electromagnetic field, one uses Maxwell’s theory rather than quantum electrodynamics. Bunge further argues that this is based on the correspondence principle (first formulated by Einstein and Bohr), practiced by all theoretical physicists, namely: A new theory should contain, perhaps in some limit, such as c ➔ ∞ or h➔ 0, the true results of the theory it is intended to replace. This is all the more important as quantum mechanics inherited the classical concepts of space, time, mass, and charge, as well as the universal concept of energy. It seems that Bunge (2003) seemed to be aware of Feyerabend’s criticism and thus concluded: “Like everything else in history, the new grew out of the old. As a consequence—Gaston Bachelard, Thomas S. Kuhn, Paul K. Feyerabend notwithstanding—there have been continuities along with discontinuities” (p. 590, italics added). (Classified as Level III). On reading a preliminary version of this chapter Kalman (2019a) commented: “Bunge is such a great philosopher but here he is unfortunately wrong. Mathematicians have gone to great lengths to show that the mathematics of classical physics is incompatible with the mathematics of Quantum Mecahnics so that taking limits such as c ➔ ∞ or h ➔ 0 do not yield classical mechanics. It only takes you from one equation to another but not from one theory to the other.” Of course, Feyerabend conceded that under certain limiting conditions the equations of relativity theory yield values that approach those calculated within Newtonian mechanics (Losee, 2001, p. 186). Feyerabend’s (1975a, p. 83) position is, however, much more nuanced, and he illustrated by comparing the philosophical positions of an instrumentalist and a realist—the former would consider all theories with the same observation language as commensurable, whereas the latter would not. The transition from Newtonian mechanics to Relativity theory involves a change of meaning of spatiotemporal concepts. For example, “classical length” and “relativistic length” are incommensurable notions. In Relativity theory, length is a relation whose value is dependent on signal velocity, gravitational fields, and the motion of the observer. An influential philosopher of science has made the case for cooperation between classical and quantum mechanics more compelling and hence drawn a blue-print for the presentation of this topic in science education:

3.2  Results and Discussion

57

Let us grant that quantum mechanics is a correct theory and that its state functions provide true descriptions. That does not imply that classical state ascriptions must be false. Both kinds of descriptions can be true at once and of the same system. We have not learned in the course of our work that quantum mechanics is true and classical mechanics false. At most we have learned that some kinds of systems in some kinds of situations (e.g., electrons circulating in an atom) have states representable by quantum state functions, that these quantum states evolve and interact with each other in an orderly way (as depicted in the Schrödinger equation), and that they have an assortment of causal relations with other non-­ quantum states in the same or other systems (Cartwright, 1999, pp. 231–232, italics added).

This seems to be a fairly reasonable solution of the incommensurability problem, as recent scholarship in philosophy of science has shown that scientific theories are neither “true” nor “false” (Giere, 2006a, b), but rather continue to evolve and progress.

3.2.10  Nature of Science Changing nature of science and scientific thinking has considerable importance for understanding scientific progress and hence science education. Based on this consideration and a history and philosophy of science perspective, Lederman (1995) has concluded: Furthermore, although Popper’s (1959) conception of science held favor for an extended period of time, Kuhn’s (1962) alternative conceptualization was quickly followed by a series of reconceptualizations (Feyerabend, 1975a; Giere, 1988; Lakatos, 1970; Laudan, 1977) significantly differing from one another. Consequently, it appears that conceptions of the nature of science are quite variable and tentative (House, 1991) (p. 373, italics added). (Classified as Level IV).

With this perspective Lederman suggested that there was an emerging recognition that the nature of science is not universal or unchanging and nor there was a ‘final’ answer to the nature of scientific thought. If this was an emerging perspective in 1995, recent research in philosophy of science has recognized the importance of alternative interpretations of historical events even more acutely (see Chap. 2 for details). In this context it is important to consider Lakatos’s proposed book The changing logic of scientific discovery, which he never finished. However, one chapter of the announced book was published in Lakatos (1978) after his death.

3.2.11  Newtonian Method While reviewing Peter Achinstein’s Evidence and method: Scientific strategies of Isaac Newton and James Clerk Maxwell, Karam (2014) has evaluated the claim that by using his rules for the study of natural philosophy Newton was able to establish a universal law of gravitation to explain a range of phenomena. Of the four rules,

58

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

perhaps the following (Rule 4) is more controversial: “In experimental philosophy, propositions gathered from phenomena by induction should be considered either exactly or very nearly true notwithstanding any contrary hypotheses, until yet other phenomena make such propositions either more exact or liable to inspection” (reproduced in Achinstein, 2013, p. 44; Karam, 2014, reproduces on pp. 2138–2139). Interestingly, not only Rule 4 is problematic, even its wording and translation has also been the subject of controversy. Following translation of Rule 4 is provided by Worrall and Currie (1978): In experimental philosophy we are to look upon propositions inferred by general induction from phenomena as accurately or very nearly true, notwithstanding any contrary hypotheses that may be imagined, till such time as other phenomena occur, by which they may either be made more accurate, or liable to exceptions. This rule we must follow, that the argument of induction may not be evaded by hypotheses (pp. 204-205; italics added. Worrall and Currie were editors of this volume of Lakatos’s papers, published posthumously).

The reader will note that the last sentence (in italics) in the translation presented by Worrall and Currie is missing in the version of Rule 4 presented by Achinstein. Actually, Lakatos (1978) presents the original version from Newton, in which this sentence is included, and refers to it in the following terms: “According to his famous Rule IV, metaphysical criticism must not be allowed to make us reject inductive proofs” (p.  204). Furthermore, there was considerable discussion with respect to the word “contrary”, and after some hesitation Newton crossed it out. However, later it was reinserted by the publishers (Bentley and Halley). In his review of Achinstein’s book, Karam (2014) concluded: By this point it should be clear to the reader that the author’s [Achinstein] intention is to sustain that Newton’s four rules served as a methodological guide for him to discover the law of gravitation. This view has received strong criticism from philosophers (e.g., Whewell, Feyerabend and Norton) who argue that Newton formulated the rules a posteriori to protect and strengthen the inference of gravitation, and that they are worthless as a means of testing or justifying inferences (p. 2139). (Classified as Level III).

Let us consider the major point of contention, namely Newton formulated his rules after having postulated his law of gravitation. In other words, strictly speaking Newtonian method was not an inductive generalization and has been the subject of considerable controversy in the history and philosophy of science literature (Cartwright, 1983; Duhem, 1914; Feyerabend, 1970d; Giere, 1999; Kuhn, 1977; Lakatos, 1978; Popper, 1957; for a detailed discussion see Niaz, 2009). According to Kuhn (1977), when Newton enunciated his theory in the late seventeenth century, only his third law could be directly investigated by experiment. Convincing demonstration of the second law had to await the development of Atwood’s machine, almost a century after the appearance of the Principia. Giere (1999) has questioned the claim that the law of gravitation could have been based entirely on experimental observations, as Newton did not consider bodies carrying a net charge. Lakatos (1978) explicitly recognizes the contributions of Duhem, Popper and Feyerabend in breaking the myth of Newtonian foundations and inductive ‘logic’: “The aim and structure of physical theory published in 1905 [Duhem, 1914], contains a brilliant and crushing criticism, which reveals some of the skeletons in the Newtonian

3.2  Results and Discussion

59

­cupboard. It is amazing how this criticism was ignored until resuscitated by Popper and his school. Popper, in his crusade against inductivism, revived and improved Duhem’s arguments in two papers published in 1948 and 1957. His papers were ignored just as were Duhem’s; they were finally given wider circulation by Feyerabend, who took up their main point in 1962” (p. 213, underline added). The controversial nature of Newton’s Rule 4 (see Karam above) is explicitly recognized by Feyerabend (1970a) in the following terms: Even Newton, who explicitly advises against the use of alternatives for hypotheses which are not yet contradicted by experience and who invites the scientist not merely to guess, but to deduce his laws from “phenomena” (cf. his famous rule IV), can do so only by using as “phenomena” laws which are inconsistent with the observations at his disposal. (As he says himself [stated, i.e., Newton]: “In laying down … Phenomena, I neglect those small and inconsiderable errors” (Feyerabend, 1970a, p. 106, original italics).

Feyerabend provides the following reference for the above quote from Newton: Sir Isaac Newton’s Mathematical Principles of Natural Philosophy and His System of the World (trans. A. Motte and rev. F. Cajori). Berkeley: University of California Press, 1953, p. 405. The point of this discussion (provided by Karam’s book review cited above) is that despite their many differences some of the major philosophers of science (Kuhn, Lakatos, Popper) agreed with Feyerabend’s thesis with respect to the Newtonian method. At this stage it would be interesting to recapitulate how Duhem (1905/1914), being one of the first to do so, conceptualized the construction of a physical theory: “Now, we have recognized that it is impossible to construct a theory by a purely inductive method. Newton and Ampѐre [André-Marie Ampѐre, known for his theory of electrodynamics] failed in this, and yet these two mathematicians had boasted of allowing nothing in their systems which was not drawn from experiment. Therefore, we shall not be averse to admitting among the fundamental bases of our physics postulates not furnished by experiment” (p. 219). Even today this statement may be considered controversial in some science education and research communities (e.g., Achinstein) and still Feyerabend along with major philosophers of science endorsed it.

3.2.12  Normal Science, Dogmatism and Science Education While reviewing Feyerabend’s (2011) The tyranny of science, Rowbottom (2013) has endorsed his position on scientism but at the same time raised the issue of whether science and science education benefit from dogmatism, and questions Feyerabend’s claim that his philosophy of science has important consequences for education: Given the venue in which this review appears, [Science & Education] I should also mention Feyerabend’s dubious claim that his findings have “important consequences for education” (p. 125) … His [Feyerabend’s] main worry … is that students are given a false picture of science, in so far as they are not told that the current view may soon be voted out of office … But Feyerabend does not discuss the possibility that dogmatism is crucial for science

60

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported… (and therefore in science education), as Kuhn (1963) argued (Rowbottom, 2013, p. 1230, italics are reproduced from Feyerabend, 2011, p.  125, underline added). (Classified as Level III).

In the light of the publication of Science & Education and recent evaluation of Feyerabend’s oeuvre by philosophers of science (see Chap. 2), it is a good opportunity to reconsider Feyerabend’s “dubious” claim (see Chap. 8). Furthermore, there is a close relationship between Feyerabend’s criticism of scientism, namely the thesis that science has all the answers about the way the world is, and his advice to students, that “the current view may soon be voted out of office”. It is important to note that this advice precisely leads to understanding the tentative nature of scientific knowledge, considered by most science educators to be an essential part of nature of science. A recent study has endorsed this thesis in forthright terms: “Scientific knowledge is reliable and durable, but never absolute or certain … Scientific claims change as new evidence, made possible through conceptual and technological advances, is brought to bear; as extant evidence is reinterpreted in light of new or revised theoretical ideas; or due to changes in the cultural and social spheres or shifts in the directions of established research programs” (Abd-El-­ Khalick, Myers, et al., 2017, p. 26, italics added). Feyerabend would have particularly liked the part in italics. With respect to dogmatism, Feyerabend (1970c) had critiqued Kuhn, “The recipe, according to these people [e.g., Kuhn], is to restrict criticism, to reduce the number of comprehensive theories to one, and to create a normal science that has this one theory as its paradigm. Students must be prevented from speculating along different lines and the more restless colleagues must be made to conform … Is this what Kuhn wants to achieve” (p. 198, italics in the original). Siegel (1979), a philosopher of science with close ties with the science education community has responded by arguing that according to Kuhn, science education does and should distort the history of science in order to inculcate the dominant paradigm of the day and textbooks are designed precisely to perpetuate ‘normal science.’ Collins (2000), a sociologist of science has gone further by pointing out that replacing normal science with the dynamics of science-in-the-making is difficult due to the dogmatic and authoritarian nature of science education. Niaz (2010) has concluded that teaching normal science leads to memorization of science content with little understanding and its elimination could facilitate the inclusion of controversies, namely unfolding of the historical episodes based on rival interpretations— thus providing students a glimpse of what science is all about, and how it really works. In order to facilitate conceptual change Kalman and Aulls (2003) have designed an introductory university level physics course with the following characteristics: In the course under study, students are presented with two alternative frameworks; pre-­ Galilean physics and Newtonian physics. The idea of the course design is that students would at first view the frameworks almost in a theatrical sense as a view of a drama involving a conflict of actors; Aristotle, Galileo, Newton and others occurring a long time ago. Based upon the final interviews, we came to the conclusion that some students gradually identify with the conceptual positions taken by the proponents of the alternative frameworks and become themselves a part of the action (p.  762, italics added). (Classified as Level V).

3.2  Results and Discussion

61

Rationale of this study is based on the epistemological advice of Feyerabend (1975a), namely evaluation and understanding of a theoretical framework requires an alternative (principle of counterinduction). Kalman and Aulls (2003) have clarified that traditional science teaching consists of exercises in problem solving (perhaps in the Kuhnian sense), and thus ignores the underlying theoretical frameworks. Precisely, this interaction and competition between alternative frameworks provides a more engaging classroom environment rather than the one based on Kuhnian normal science. Furthermore, during the course activities students discover that different theoretical frameworks are related to concepts drawn from various parts of the course.

3.2.13  Polanyi’s Tacit Knowledge Polanyi (1964, 1966) differentiated between two types of knowledge: (a) Explicit, articulated and formal knowledge; and (b) Tacit, unarticulated and non-formalized knowledge. Polanyi argued that the first cannot be achieved without the second. Based on this perspective, Glass (2013) has formulated a model of scientific thinking: Advances in science break from conventional (explicit) understanding. Therefore, a set of rules needs to exist regarding social practices and political norms, allowing for the development of new ideas (Feyerabend, 1975a; Latour, 1987). These conventions are not easy to articulate and are rarely made explicit. We can ask ourselves “what makes a good experiment?” and provide numerous characteristics, but these will not ensure a good experiment. This is because we cannot articulate all that goes into a good experiment. It is too deeply rooted in our sense of the field itself, too dependent on context to account for every instantiation and there are possibilities that as yet remain undone (pp. 2712–2713, italics added). (Classified as Level IV).

Except for the part in italics (set of rules) this is a fairly good representation of Feyerabend’s research agenda, precisely because aspects such as “sense of the field” and “context” are primarily tacit. It is plausible to suggest that Feyerabend’s philosophy of science and Polanyi’s tacit knowledge approximate to a certain degree. The ability of scientists to generate new knowledge does not depend on a set of rules and procedures, but rather on their ability to use different methods selectively and in this context: Feyerabend’s (1975a) famous assertion that ‘anything goes’ implies the absence of a prescribed method, the absence of an algorithm, rather than the absence of methods. It should not be taken too literally (Hodson, 1992, p. 131, original italics). (Classified as Level IV).

Furthermore, Hodson considers that it is a disservice to children to state that the world of the scientist is totally anarchic (p. 131) or for that matter to state that science has a single method. Finally, Hodson concluded that scientists proceed intuitively based on their personal understanding and tacit knowledge of how to do science (p.  132). It is plausible to suggest that this shows a proximity between

62

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

Polanyi’s tacit knowledge and Feyerabend’s philosophy of science. I will elaborate on this theme in the final (conclusions) chapter. Jacobs (2000) has drawn attention to the importance of Michael Polanyi’s (1958/1964) magnum opus, Personal Knowledge, and for science educators the leitmotif of his book, namely personal judgment as the paramount arbiter of all our intellectual performances. Not only Feyerabend but also Daston and Galison (2007) would endorse such a thesis. According to Jacobs, it is plausible to suggest that Polanyi’s ideas can be presaging and even catalyzing themes in thinkers as diverse as Popper, Kuhn, Ravetz, Knorr-Cetina and Feyerabend. The case of Feyerabend is particularly interesting as Jacobs considers that his views on the denial of “general method rules” were previously put forward by Polanyi. (Classified as Level III).

3.2.14  Science and Religion Cordero (2001) has highlighted the following features (among others) of Feyerabend’s (1975b, 1993) philosophy of science: (a) Science should be regarded as just another religion (p.74); (b) A formal separation between state and science just as there is now a formal separation between state and church (p. 74); and (c) To protect children against dogmatic scientism it is worthwhile to develop in them the spirit of contradiction by providing alternative ways of thinking such as creationism, voodoo and astrology (p.  74). Based on these and similar considerations, Cordero (2001) has concluded: Science receives state and public support primarily because it achieves truths that deserve to be preserved and privileged. Feyerabend disagrees with this policy. It is a mistake, he thinks, because science is not the only enterprise delivering valuable results, and—its pretensions to the contrary notwithstanding—science lacks a special method for achieving such results. In Feyerabend’s view nothing but political circumstance privileges what we currently take as the ‘scientific truth’ over claims from allegedly disreputable specialties like astrology, voodoo or magic. Build up the society he recommends, he says, and voodoo and astrology and the like will return in such splendor that you will have to work hard to maintain your own position and will perhaps be entirely unable to do so (pp. 73–74, italics added). (Classified as Level IV).

Leaving aside Feyerabend’s rhetoric, it is no surprise that if creationism and evolution were given equal time and opportunities, those teaching the science curriculum (even in the industrialized countries) would be hard pressed to convince students of the ‘scientific’ nature of evolutionary theory. Teaching evolution as a series of facts may lead to dogmatic scientism, whereas Feyerabend would instead recommend a contradictory approach based on alternative ways of thinking. This does not necessarily imply that Feyerabend is recommending creationism. Interestingly, at the end of the above quote, Cordero (2001) raises the following very thought provoking question: “But, isn’t science plainly better than those ‘primitive practices’?” (p. 74). Of course, the answer is in the affirmative. Nevertheless, it is important not to lose

3.2  Results and Discussion

63

sight of the fact that science is basically not only counter intuitive but “unnatural” and hence the difficulties involved in practicing and learning it. Meera Nanda (2003) in her book, Prophets facing backward: Postmodern critiques of science and Hindu nationalism in India, has espoused the thesis that contemporary India is in the grip of a reactionary modernism in which technological advance is situated within the context of nationalistic religion, namely “Vedic Science.” While reviewing the book, Nola (2004a, b) concluded: But Nanda tells us that they do cite some Kuhn and Feyerabend, and they do know that in the West there are those who decry modernity and debunk the enlightenment pretensions of defenders of the rationality of science. For such nationalists this is enough to justify their separation of science from broader aspects of enlightenment rationality and its relocation in a different socio-political and religious setting; in so doing they create a distinctive “Vedic Science.” (p. 244). (Classified as Level I).

Such relocation of Feyerabend’s thinking in an entirely different socio-political structure shows, all the more, the need for its reevaluation in the context of recent developments in philosophy of science (see Chap. 2). It is interesting to note that “Vedic Science” attributes India’s possession of the atomic bomb, and the science behind it, to God Krishna and the Bhagavad Gita, with its reference to a “thousand suns.” Actually, similar chauvinistic accounts based on the Quran are quite common in the Islamic countries and are not necessarily based on Feyerabend’s philosophy of science.

3.2.15  Scientific Expertise and Galileo Finocchiaro (2011) recounts how Feyerabend based on the views of Cardinal Bellarmine, supported the views of the church to criticize Galileo and thus neglected scientific expertise (for further details on Bellarmine and Galileo see Chap. 2). Despite efforts by Pope John Paul II to rehabilitate Galileo in 1979, later in 1994 Cardinal Ratzinger (later Pope Benedict) criticized Feyerabend’s views and thus reversed this trend. After a detailed analysis of the relationship between Bellarmine and Galileo, in the context of scientific expertise, Finocchiaro (2011) has recommended a Galilean approach to the Galileo affair in the following terms: … my over-arching thesis is that today, in the context of the Galileo affair and the controversies over science versus religion and over institutional authority versus individual freedom, the proper defense of Galileo should have the reasoned, critical, judicious, open-minded, and fair-minded character which his own defense of Copernicus had (p. 65, italics in the original). (Classified as Level IV).

More recently, Finicchiaro (2019) has elaborated further on the Galileo affair. Interestingly, despite his criticism of Galileo, in some of his writings, Feyerabend also supported such an approach in his philosophy of science. For example, in his reconstruction of the role for experiment, Kalman (2009a) has shown how Feyerabend endorsed Galileo’s method (viz., an example of counterinduction):

64

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported… Feyerabend (1993) has given the name counter induction to Galileo’s method. This procedure is to first accept the veracity of the theory (in this case the heliocentric model of the solar system) and then discover what changes in the theory (principle of inertia) are required to remove any contradiction between the theory and observation (p. 28). Classified as Level V.

This clearly shows that Galileo first compared theory with theory (counterinduction) and then went on to do experiments with inclined planes. Interestingly, according to Nola (2004a, b), “On Feyerabend’s understanding of Galileo’s method, we need to overcome, and even contradict, the commonsense deliverances of experience, especially when new, profoundly deep theories are being developed” (p. 352, Classified as Level III). Next, in order to go beyond, Nola cites the following from Galileo himself, “For just as I … have never seen nor ever expect to see the rock fall [tower experiment] any way but perpendicularly, just so do I believe that it appears to the eyes of everyone else. It is therefore better to put aside the appearance, on which we all agree, and to use the power of reason either to confirm its reality or to reveal its fallacy” (Galileo, 1967, p. 257). At this stage Nola clarifies that this passage from Galileo was also cited by Feyerabend (1975a, p. 7). Nola considers that Galileo used strong words to understand fallacy in experience and concluded that this led Feyerabend to his view that all observations are theory laden (p. 352). Clarke (2016) has reviewed J. Agassi’s Popper and his popular critics: Thomas Kuhn, Paul Feyerabend and Imre Lakatos, to highlight the fact that some of Feyerabend’s anti-rationalistic and hyperbolic flourishes (e.g., science is not superior to magic) were primarily teasers or challenges to defend diversity, and Feyerabend acknowledged this aspect in his correspondence with Agassi. In Clarke’s view Popper’s replacement of proof with empirical criticism opened the road to pluralism in science which was later endorsed by Feyerabend, and he cites Agassi (2014) to support his claim: Feyerabend did support pluralism but as anti-science, and with some justice. For, the classical idea of science was the set of proven assertions; Popper’s replacing proof with openness to empirical criticism opened the road pro-science pluralism. He developed pluralism before Feyerabend and I can testify to this as I sat at his feet when he did so … Popper presented his pluralism in his magnificent “Towards a rational theory of tradition” … where he offered a theory that put on a par the diversity of tradition, legislation and scientific hypotheses (pp. 110–111, italics added).

Interestingly, however, Agassi also acknowledges that Feyerabend’s “anti-science” stance in most cases could be attributed to his rhetorical excesses. Furthermore, Feyerabend’s proposal with respect to the separation between science and state is questioned and instead it is suggested that there is a need for organizations that should look after the public interest in science, that is specific controls and not separation. (Classified as Level IV).

3.2  Results and Discussion

65

3.2.16  Scientific Method Scientific method as a sequence of well-defined steps in which observation and rigorous experiments play a central role, has been the subject of criticism by Gil-­ Pérez et al. (2005). Furthermore, these authors rightly consider that, “… obsessive preoccupation with avoiding ambiguity and assuring the reliability of the evaluation process distorts the nature of the scientific approach itself, essentially vague, uncertain, intuitive” (p. 313, italics added). Interestingly, the part in the italics would be endorsed by Feyerabend’s epistemological anarchism. However, given the widespread misunderstanding of Feyerabend’s philosophical stance, these authors concluded: Some teachers, in rejecting this rigid and dogmatic view of science, may accept an extreme relativism, both methodological—‘anything goes’, there are no specific strategies in scientific work (Feyerabend, 1975a, b, c)—and conceptual: there is no objective reality which allows us to test the validity of scientific construction (p. 313). (Classified as Level I).

This has important implications for science education, namely on the one hand we have to criticize the ‘dogmatic view of science’ (scientific method), and then go beyond and accept Feyerabend’s advice that scientific work involves a variety (plurality) of strategies (methods). Hence, both science and teaching science requires the understanding that scientific work is vague, uncertain and intuitive—in Feyerabend’s terminology ‘anything goes.’ Two prominent philosophers of science have espoused the family resemblance approach to understanding the nature of science (NOS), and within that perspective they have considered that: Many advocates of the consensus view are dismissive of scientific methodology or methodological rules. They say that there is no algorithm, no fixed and universal set of rules that govern scientific activity at every stage of inquiry. This is certainly true, a point made forcefully by Paul Feyerabend in his book Against Method when he said “anything goes” … However, there is no reason to understand scientific methodologies such as various forms of inductivism, hypothetico-deductivism and Bayes’ Theorem and its associated methods in this strict manner (Irzik & Nola, 2011, p. 599). (Classified as Level II).

Consensus view refers to a wide range of science educators who consider that despite the complexity of the history and philosophy of science some degree of consensus can be achieved in order to introduce NOS in the classroom (e.g., Lederman, 2004). At first sight it appears that Irzik and Nola would seem to be endorsing “anything goes” not in its traditional sense of Feyerabend as a relativist. However, at the end of the paragraph, they once again reiterated the traditional criticism, namely “Without the notion of scientific method and methodological rules, the self-corrective nature of science becomes a mystery” (Irzik & Nola, 2011, p. 599). Actually, the self-corrective nature of science depends more on diversity and plurality in a discipline, rather than the scientific method (cf. Giere, 2006a, b; Niaz, 2018).

66

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

Heffron (1995) has critiqued the role played by induction in the science curriculum and pointed out that: If, as Karl Popper and others have argued, science itself does not advance “solely by inductive methods”, that is, by the simple stockpiling and ordering of observations, however repetitious, we cannot expect to make our children (often considered “natural scientists” because of their superior observational skills) better scientists by simply making them more observant. We must first make them more theoretical (p. 245).

This statement and the perspective that underlies has important implications. As it is generally believed that science progresses by stockpiling and ordering of observations (i.e., inductive perspective), the same holds for children—considered to be natural scientists. In this context, it is important to note that Heffron (1995) added an endnote to the above statement, in the following terms: “As the work of Kuhn, Feyerabend, Toulmin and Lakatos and others suggests, falsification is not an unproblematic model of science” (p. 248). (Classified as Level III). Precisely, falsification has been endorsed by Popper as an important part of the scientific enterprise and that inevitably leads to recognizing observations as crucial for testing a theory. On the other hand, Feyerabend, Lakatos and Kuhn were particularly critical of the role of falsification in scientific progress. Furthermore, inclusion of Feyerabend along with other philosophers of science is an indication of some consensus among these philosophers. Experiments to determine the composition of air conducted by A.  Lavoisier (1743–1794) and J. Priestley (1733–1804) in the eighteenth century have been the subject of considerable controversy (Kuhn, 1970). It is generally ignored that both Lavoisier and Priestley interpreted the experiments based on their worldviews. Before Lavoisier did his experiments in 1772, scientists generally followed G. Stahl’s (1660–1734) phlogiston theory according to which metals on being burnt in air must decrease in weight due to the loss of phlogiston (for example, formation of mercuric oxide on heating mercury in air). On the contrary, Lavoisier’s experiments showed that metals on being burnt in air increased in weight. According to the inductivist interpretation (also followed by general chemistry textbooks), such findings decisively refuted phlogiston theory. History of chemistry, however, shows that as early as 1630 (long before Lavoisier came on the scene) J. Rey (1630) had reported that it was common knowledge that metallic oxides weigh more than the metals from which they were prepared. According to Musgrave (1976b) who gives a detailed account of the episode, “… if Lavoisier’s 1772 experiment refutes phlogiston theory, then phlogiston theory was born refuted” (p. 183, original italics). De Berg (2014) presents a detailed account of the experiments conducted by Priestley and Lavoisier to determine the composition of air (for example by heating mercurius calcinatus, i.e., mercuric oxide). A major objective of De Berg is to compare the methodologies used by Priestley and Lavoisier in the context of the scientific method. For example, Lavoisier’s self-confessed emphasis on facts derived from observation and experiment and avoidance of speculation appears quite similar to Chalmers’ (1982) understanding of scientific knowledge. Next, De Berg (2014) goes on to compare the views of Chalmers and Feyerabend:

3.2  Results and Discussion

67

Feyerabend viewed science in a less prescriptive way than many [e.g., Chalmers]. He (Feyerabend, 1993, p.18) was adamant that “the idea of a fixed method, or a fixed theory of rationality, rests on too naïve a view of man and his social surroundings”. The boundary between facts and theories was fuzzy or blurred according to Feyerabend’s understanding of science … facts and theories are much more intimately connected than is admitted by the autonomy principle (1993, p. 27); facts and theories are never as neatly separated as everyone makes them out to be (1993, p. 51); theories, observations, and experimental results are not as well defined as we think (1993, p. 51); and facts that enter our knowledge are already viewed in a certain way and are, therefore, essentially ideational (1993, p. 11). It would seem that Feyerabend’s opposition to a fixed method would have resonated more with Priestley than Lavoisier, but in both Priestley and Lavoisier’s mind the boundary between facts and theory would have been more clearly defined than that described by Feyerabend. But were Priestley and Lavoisier simply unaware of their commitment to some overarching model or theory in their observation statements? (p. 2057). (Classified as Level III).

In order to respond to the query in the last sentence of the above quote, De Berg (2014, p. 2059) suggested that Lavoisier’s commitment to the law of conservation of mass guided the reporting of his results. Interestingly, however, De Berg raises another important issue by quoting the following from Gillispie (1960, p.  231): “Scientists have sometimes written that Lavoisier formulated the law of conservation of matter. The reality was simpler. He assumed it.” Gillispie, a historian of science has raised an important issue that has been the subject of considerable discussion. The difference between “formulate” and “assume” is important and is the subject of controversy in the history and philosophy of science literature. For example, in the seventeenth century did Newton have experimental evidence to formulate his laws or he assumed them (cf. Duhem, 1914, p. 219). Actually, Duhem was categorical that it was impossible to construct a theory by a purely inductive method and Newton failed in this attempt (for details see Niaz, 2009, Chaps. 2 and 3). More recently, in the twentieth century, Millikan-Ehrenhaft controversy with respect to the determination of the elementary electrical charge was also based on the assumptions brought to bear by the two in order to interpret their experimental data (cf. Holton, 1978a, b). Millikan assumed that electrons existed, whereas Ehrenhaft held on to the assumption of subelectrons. Priestley tried to duplicate Lavoisier’s experimental work but without success and thus accused Lavoisier of over-estimating the accuracy of his measurements (similarly, Ehrenhaft could not duplicate some of Millikan’s experiments). De Berg (2014) provides the following explanation from W.H. Brock, a historian of chemistry to understand the dilemma: “Priestley’s objections to Lavoisier’s chemistry were often, indeed, usually, perfectly valid … For example, in the decomposition of mercuric oxide Priestley consistently got less mercury back than he started with. In any case, he observed, Lavoisier’s pretence of measuring to four or five places of decimal was pure window dressing” (Brock, 2008, p. 75). It seems that besides the lack of better experimental apparatus, Lavoisier’s commitment to the law of conservation of mass compounded his difficulties. In contrast, De Berg provides the following quote from Priestley in order to provide insight to his methodology: “If we could content ourselves with the bare knowledge of new facts, and suspend our judgment with respect to their causes, till, by their analogy, we are led to the discovery of more facts, of a

68

3  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

similar nature, we should be in a much surer way to the attainment of real knowledge (Priestley, 1790, p. xxix, italics added). Interestingly, based on his seminal work with respect to the Millikan-Ehrenhaft controversy, Holton (1978a) provided similar advice, “… the graveyard of science is littered with those who did not suspend belief while the data were pouring in” (p. 212, italics added). Such controversies (Priestley-Lavoisier and Millikan-Ehrenhaft), can help us to understand Feyerabend’s (1993, p. 21) advice that, “history of a science becomes an inseparable part of the science itself” and De Berg has suggested their inclusion in current chemistry education. According to Koertge (1996), the real revolutionary advances in science take place when two comprehensive theoretical systems clash. Although, Koertge recognizes the importance of scientific method but still endorses Feyerabend’s thesis of a clash between comprehensive theories and provides the conflict between Lavoisier’s oxygen theory and the phlogistonists, as an example. Phlogiston theory’s claim that a metal loses weight when heated in air could be refuted with experimental evidence. However, as it was a comprehensive theory it treated more or less successfully subjects such as combustion, calcination, plant growth, acid-base behavior and the color of chemicals. Based on these considerations, Koertge (1996) concluded: This means that to understand and appreciate the chemical advances made by Lavoisier and other oxygen chemists, one must first thoroughly study both the empirical comprehensiveness and the metaphysical attractiveness of phlogiston chemistry. Then this system must be compared and contrasted with its eighteenth century competitor and successor, not modern chemistry … If this indeed be the case (and there is considerable evidence for the position—see especially the work of Feyerabend, Lakatos, and Kuhn ), and if we are really serious about wanting students to gain a sophisticated understanding of the nature and growth of science, then we must be prepared to include in the curriculum an intensive and sympathetic study of scientific systems that are scientifically unacceptable today. Students must become knowledgeable phlogistonists, caloricists or Lamarckians (p.  399, original italics, underline added). (Classified as Level IV).

It is important to note that Noretta Koertge is a prominent philosopher of science and a former Editor of the journal Philosophy of Science, official journal of the US-based Philosophy of Science Association. Her position as presented above not only comes quite close to Feyerabend’s thesis of counterinduction (see Chap. 7) but also presents a guideline for the science curriculum based on competing or plurality of theoretical systems—again very much in tune with Feyerabend’s philosophy of science. Furthermore, Koertge’s article was first published in 1969 and published again by Science & Education in 1996, as part of the Golden Oldies series, and antecedes the recent reevaluation of Feyerabend’s oeuvre by philosophers of science.

3.2.17  Situated Learning Lave and Wenger (1991) have proposed that all learning is situated and occurs by means of legitimate peripheral participation (LPP) within a community of practice. Traditional schools, in contrast, are based upon the premise that knowledge can be

3.2  Results and Discussion

69

imparted in a decontextualized sense. LPP is a form of apprenticeship in which the role of the community is important. Ben-Ari (2005) considers situated learning untenable as a guiding principle for science education, as it is not only deterministic but also a child becomes an apprentice in a particular subject/field at an early age. This leads Ben-Ari to the following dilemma: How do we choose a community of practice for a 12-year old child? Based on these considerations the author concluded: … an attempt to apply the principles of LPP to science education will lead to disaster: an appalling waste of time of highly trained specialists and unacceptable inequities in access to educational opportunities. To insist on “real” situations ignores both the diversity of the students and the dynamic nature of reality in our hi-tech world. Problem design must be based on considerations of motivation and pedagogy, rather than verisimilitude to the real world. To paraphrase Feyerabend (1975a/1993, p.  14), anything goes when it comes to choosing pedagogic methods that will improve learning (p. 374, italics added). (Classified as Level III).

Indeed, given the diversity of the students’ interests (especially at a young age) and the dynamics of classroom interactions, anything goes seems to be a particularly appropriate pedagogic strategy, namely a teacher can adapt according to the needs, aspirations and motivation of the students. This chapter provides examples of research reported in the journal Science & Education (17 categories) that facilitate a wide range of perspectives with respect to understanding Feyerabend’s epistemological anarchism. These examples provide a glimpse of research conducted in various parts of the world over a period of more than 25 years. Conclusions based on these findings along with those from Chaps 4, 5, 6 and 7 will be synthesized and presented in Chap. 8.

Chapter 4

Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal of Research in Science Teaching (Wiley-Blackwell)

4.1  Method The Journal of Research in Science Teaching (JRST) is the official journal of the US-based National Association for Research in Science Teaching (NARST), which has members in many countries around the world. JRST started publishing in 1963 and is indexed in the Social Sciences Citation Index (Thomson-Reuter). In October 2017, I made an online search on the website of JRST with the key words “epistemological anarchism” and “Feyerabend” (http://onlinelibrary.wiley.com/journal/10982736). This gave a total of 21 articles which were evaluated on the same criteria (Levels I–V) as in the previous study (see Chap. 3). Following the guidelines based on Charmaz (2005), presented in Chap. 3, and in order to facilitate credibility, transferability, dependability, and confirmability (cf. Denzin & Lincoln, 2005) of the results, I adopted the following procedure: a) All the 21 articles from the Journal of Research in Science Teaching, were downloaded and after evaluation were classified in one of the five levels, I–V (for levels see Chap. 3); After a period of approximately three months all the articles were evaluated again and there was agreement of 93% between the first and the second evaluation. It is important to note that all the articles evaluated in this study referred to epistemological anarchism in some context, which may not have been the primary or major subject dealt with by the authors. Detailed examples from different levels are presented in the next section. A complete list of all the 21 articles from JRST that were evaluated is presented in Appendix 3. Distribution of all the articles according to author’s area of research, context of the study and level (classification) is presented in Appendix 4.

© Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_4

71

72

4  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

4.2  Results and Discussion Based on the treatment of the subject by the authors, 9 categories were developed to report and discuss the results. These categories are presented in alphabetical order. It is important to note that the idea behind creating the 9 categories is to facilitate the reader to find the subject of her/his interest. Given the wide range of subjects discussed by the authors over a period of more than 30 years, it is difficult to create the semblance of a continuous storyline. Similarly, due to limitations of space it is not possible to present a detailed critical analysis of every article. The following are the 9 categories (presented in alphabetical order) that were created to present and discuss the results: 1 . Alternative literary forms 2. Creativity 3. Evaluation 4. Feminism 5. Nature of science 6. Proliferation of theories 7. Scientific method 8. Teacher demonstrations 9. Worldviews

4.2.1  Alternative Literary Forms Roth and McGinn (1998) have criticized the science education community for not bringing to the forefront unnoticed and uncomprehended assumptions in their research. As an alternative these authors have espoused alternative literary forms based primarily on Latour’s (1987) Actor Network Theory (ATN). Based on science and technology studies, ATN has recommended that in order to understand science it is important to study how scientists arrive at conclusions and not the products of scientific inquiry. Furthermore, it is essential to investigate how controversies are resolved based on the perspectives and relationships of the network in which the different actors take part. As examples of alternative literary forms, Roth and McGinn provide two examples: (a) Bjelic and Lynch (1992) argued that to understand what Goethe was doing one needs to look through a prism rather than employ discourse analysis on the texts he produced about his color experiments and theory; and (b) Latour (1992) used fact and fiction to investigate why an 18-year multibillion dollar project was ultimately cancelled. This literary style contains original documents of various sources, government briefs and industry advertisements. It is expected that in both cases alternative literary forms provide readers with enough information so that they can construct their own conclusions. Based on these considerations both authors (Roth and McGinn) then debate the merits of alternative literary forms for the science education community and concluded: “But the

4.2  Results and Discussion

73

d­ ialogue form that we are using just now really is old and has been used throughout human history from Plato to Galileo to Feyerabend and, befitting this article, by Trevor Pinch …” (Roth & McGinn, 1998, p. 224; among others, authors provide the following references: Galileo, 1960; Feyerabend, 1991a, 1991b). (Classified as Level IV).

4.2.2  Creativity Bransford and Donovan (2005) consider science to be ultimately a creative endeavor and thus science education should represent science not as “truths to be memorized” but rather as a creative process of observing, imagining and reasoning. Based on this perspective Beghetto (2007) has explored the relationship between students’ self-­ perceptions of their ability to generate creative ideas (creative self-efficacy) and their perceived competence in science and concluded: Paul Feyerabend, the philosopher of science, held as his deepest conviction that science is an aspect of human creativity (Godfrey-Smith, 2003, p.  111). Although, many of Feyerabend’s claims about science remain controversial, this one is not. Few would disagree with the assertion that science is a creative endeavor. Indeed, creative ideas fuel the work of science and ultimately lead to the development of new theories, new knowledge about the physical world, and scientific and technological breakthroughs (p.  802). (Classified as Level III).

Indeed, various aspects of Feyerabend’s philosophy of science (e.g., counterinduction, proliferation of theories) do emphasize human creativity.

4.2.3  Evaluation Technicist approaches to evaluation are generally based on the assumption that scientific method and empirical testing offer the only means to rationality, objectivity, truth and hence objective knowledge. Robottom (1989) has questioned this assumption: “Critics suggest that empirical testing, rather than enabling access to objective reality, produces inevitably personal and social constructs of reality. A number of philosophers including Popper (1963a, 1963b), Hanson (1958) and Feyerabend (1975a) claim that an observation, far from being prior to theory, logically presupposes theory. Researchers’ observations are ordered by their internal cognitive structures (Piaget, 1971)” (p. 439). (Classified as Level III). Based on these considerations, Robottom concluded that this is not an argument against being empirical but rather an argument against inferring objectivity from empirical approaches. Inclusion of Piaget along with the philosophers of science (Popper, Hanson & Feyerabend) is particularly interesting given Piaget’s interest in philosophy of science and its influence in science education (Piaget & Garcia, 1989). For further

74

4  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

details on the relationship between observation and objectivity and its evolving nature in the history of science, see Daston & Galison, 2007 and Niaz, 2018.

4.2.4  Feminism The role of gender has been the subject of considerable research in science education. Many feminist scholars consider that science is masculine at the surface level, at the deeper epistemological level, and in the nature of the knowledge that is accepted as scientific. Based on this premise, Roychoudhury, Tippins, & Nichols (1995) have referred to the masculine worldview which considers science as being objective, rational, individualistic, unemotional, and value-free. Although philosophers of science (Feyerabend, 1978/1982; Kuhn, 1970) have questioned this perspective of science, these authors consider that this vision of science still dominates the academia (Classified as Level II). Furthermore, feminist scholars have argued that the masculine worldview leads to girls’ avoidance of science.

4.2.5  Nature of Science After reviewing research conducted over the previous forty years, related to students’ and teachers’ understanding of nature of science (NOS), Lederman (1992) has particularly emphasized the tentative nature of scientific knowledge and concluded: Furthermore, when one considers the differences among the works of Popper (1959), Kuhn (1962), Lakatos (1970), Feyerabend (1975a), Laudan (1977), and Giere (1988) it becomes quite clear that there is no singularly preferred or informed nature of science and that the nature of science is as tentative, if not more so, than scientific knowledge itself (p. 352). (Classified as Level III).

In the case of biology, Lederman clarifies that current views of evolutionary processes (cf. punctuated equilibrium, Gould, 1980) have changed despite its origin in Darwin’s work and the theory of natural selection. This has important educational implications as we do not have to impose a particular view of the nature of science on teachers and students. In this context Feyerabend’s ideas can provide a more ample perspective. Despite widespread research with respect to students’ and teachers’ understanding of NOS, this still is a contentious area of research. For example, Alters (1997) considered that the criteria for elaborating NARST tenets for evaluation of NOS in both qualitative and quantitative instruments are universally absent from the literature (p. 40). Furthermore, there was no consensus among philosophers of science with respect to various NOS tenets used by science educators. Smith, Lederman, Bell, McComas, and Clough (1997) countered by arguing that although there was

4.2  Results and Discussion

75

no unanimity among philosophers of science still there was reasonable consensus to facilitate NOS understanding. As philosophers of science, Eflin, Glennan and Reisch (1999) have endorsed the following NOS tenets, with respect to which there is consensus among science educators: (a) The main purpose of science is to acquire knowledge of the physical world; (b) There is an underlying order in the world which science seeks to describe in a maximally simple and comprehensive manner; (c) Science is dynamic, changing, and tentative; and (d) There is no one, single scientific method. Of course there are areas of disagreement of which the following are important: (i) Observation alone cannot give rise to scientific knowledge in a simple inductivist manner; and (ii) We view the world through theoretical lenses built up from prior knowledge. Based on these considerations, Eflin et al. (1999) concluded: “These beliefs are closely associated with the work of Thomas Kuhn, who argued that it is impossible to make a firm distinction between observational and theoretical languages in the way that some logical positivists had hoped (Maxwell, 1962; Kuhn, 1970; Feyerabend, 1993)” (p. 109). (Classified as Level II). Eflin, Glennan and Reisch (1999) also suggested that science educators instead of appealing to philosophers of science can themselves become acquainted with philosophical debates about science as the philosophical positions are more than a list of tenets (pp. 111–112). Secondly, science educators can illustrate the rich complexity of science with its practice and its history (p. 112). A review of the literature shows that science educators have tried to follow the first suggestion. However, the second suggestion has generally been ignored, namely integration of domain specific (history of science) and domain general (philosophy of science) aspects (cf. Niaz, 2016).

4.2.6  Proliferation of Theories Epistemological problem of the effect of theoretical commitments on what is observed by the scientist is not independent of theory, is important for science education. This leads to a methodological problem: sometimes scientists overlook or misconstrue evidence that would count against their own theory. In Kuhn’s framework commitment to a particular scientific paradigm makes it difficult to recognize anomalies. On the other hand, empirical research without a theoretical framework to guide would be chaotic. One alternative could be the idea of “working hypotheses” and a multiplicity of hypotheses could lead to “conflicting working hypotheses” within a domain. Based on Feyerabend (1963) and these considerations, Martin (1970) recommends proliferation of theories in the following terms: Students of science should be taught a number of different theoretical approaches in a domain of research. If necessary, discarded theories from the history of science should be resurrected and re-examined. But not only should students be exposed to different theoretical approaches, they should learn to work easily with different theories, now seeing the domain from the point of view of one theory, now seeing the domain from the point of view

76

4  Understanding Epistemological Anarchism (Feyerabend) in Research Reported… of another, switching back and forth to get various theoretical perspectives and insights (p. 189, italics added). (Classified as Level V).

The role of discarded theories from the history of science has also been recognized by Koertge (1996) in the context of Priestley’s phlogiston theory as compared to Lavoisier’s oxygen theory (for details see Chap. 3). Feyerabend’s concept of proliferation of theories in this case means that students can learn to work with many theories as working hypotheses in a given domain of inquiry (e.g., evolutionary theory). In other words, having more theories to work with, there is less chance that one will persist with just one theory. Starting with three “keystone” philosophers (Kuhn, Popper and Hempel), Loving (1991), has developed a Scientific Theory Profile for possible use with preservice and inservice science teachers. A central question in the elaboration of the profile was the following: Do science teachers explore how philosophers differ in their interpretations of science as opposed to a course being taught with one perspective? (p. 825). Another important consideration was the role played by history of science to determine “how science is actually practiced—balancing the more traditional exclusive emphasis on how it should be practiced” (p. 826, original italics). The difference between “is” and “should” with respect to how science works is essential for understanding Feyerabend’s oeuvre and why he is referred to as an “enfant terrible”: Feyerabend (1975a) … has been referred to as the enfant terrible in philosophy of science. A brilliant student of Popper, he both rebelled against and embraced his mentor’s writings. Insisting that good science has always involved a lot of faith, chaos, play, and downright irrationality, he, like Popper, believes in the proliferation of theories. He says, however, that the winning theories could not have arrived at their position without conceit, passion, and prejudice (Loving, 1991, p. 830, italics added). (Classified as Level V).

This is an interesting presentation as it first draws attention to Feyerabend’s relationship as a student of Popper and later the two drew apart. The reference to “downright irrationality” is important as it sums up the continuous debate and dialogue in “For and against method” between Lakatos and Feyerabend (cf. Motterlini, 1999). Although, the idea of proliferation of theories is also implicit in the writings of Popper and Lakatos, in the case of Feyerabend this is attributed to violation of some rule or method that is generally considered to be part of the scientific method.

4.2.7  Scientific Method Lederman, Abd-El-Khalick, Bell and Schwartz (2002) traced the origin of the scientific method to the seventeenth century work of Francis Bacon. Furthermore, these authors consider that inductivism and other related epistemological stances, such as Bayesianism, falsificationism and hypothetico-deductivism have facilitated the formation of a myth (especially in science textbooks) with respect to the scientific method:

4.2  Results and Discussion

77

The myth of the scientific method is regularly manifested in the belief that there is a recipe like stepwise procedure that all scientists follow when they do science. This notion was explicitly debunked: There is no single scientific method that would guarantee the development of infallible knowledge (AAAS, 1993b; Bauer, 1994; Feyerabend, 1993; NRC, 1996; Shapin, 1996). It is true that scientists observe, compare, measure, test, speculate, hypothesize, create ideas and conceptual tools, and construct theories and explanations. However, there is no single sequence of activities (prescribed or otherwise) that will unerringly lead them to functional or valid solutions or answers, let alone certain or true knowledge (pp. 501–502). (Classified as Level II).

Abd-El-Khalick, Waters, and Le (2008) evaluated the representations of nature of science in high school chemistry textbooks published over the last four decades in the US and found with respect to scientific method that it seems to thrive and is championed despite longstanding and continuous debunking as a myth. Furthermore, the authors consider that the myth continues despite criticisms (e.g., AAAS, 1993b; Bauer, 1994; Feyerabend, 1993; and Shapin, 1996). (Classified as Level II). In this context, Windschitl (2004) considers that some of the ideas of science teachers appear consistent with a “folk theory” of an atheoretical scientific method, “… that is promoted subtly, but pervasively, in textbooks, through the media, and by members of the science education community themselves” (p. 481). In order to facilitate students’ understanding of science by doing science, O’Neill and Polman (2004) have compared the philosophies of Popper and Feyerabend: Popper (1959) argued for a strict logic of theory building and testing, which relied on the verifiability and refutability of theories. In his system, empirical investigations that refute theories will always be possible and should often be pursued, because knowledge develops through the debate that ensues. Arguing from a radically different and less logical perspective on scientific method, Feyerabend (1993) characterized theories and observations as also competing with one another for ascendancy, albeit in a different manner than Popper claimed (p. 238).

Comparing the philosophical work of Popper and Feyerabend is instructive for science educators, as despite the common features, the scientific enterprise goes far beyond the Popperian refutability of theories as, “It involves uncertainty, wasted effort, and critical encounters with peers. The traditional lecture, lab, and demo approach to science education, which is streamlined to expose students to as many scientific findings, principles, and methods as possible in the shortest time, systematically obscures these elements of the scientific enterprise” (O’Neill & Polman, 2004, p. 239). (Classified as Level IV).

4.2.8  Teacher Demonstrations Teacher demonstrations are an important part of science education in most parts of the world. It is assumed that such experiences can provide students far more elements to understand a topic than oral presentations. Although such demonstrations may be well intentioned, Roth, McRobbie, Lucas and Boutonné (1997) have raised an important problem with respect to the epistemological stance of the teacher

78

4  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

c­ onducting the demonstration (e.g., transmission view of learning) that may make it difficult for students to grasp essential elements of the topic under discussion (e.g., conservation of angular momentum). In this particular study (based on Australian high school students) the authors found that students did not understand the demonstration, due to (among other reasons) lack of a theoretical framework to separate signals from noise and opportunities for students to test their descriptions and explanations. Based on these considerations, the authors concluded: Science teachers often employ demonstrations to show scientific principles in action. However, there may be some problems with this practice. It is widely accepted that all observation is interpretation (Feyerabend, 1975a; Hanson, 1958; Hodson, 1992; Rorty, 1989). Because interpretation arises from the interplay of existing understandings (prior experience) and the world, what one observes depends on what one already knows. This means that students who do not yet know the relevant scientific principles will be unlikely to see just what the demonstration is to show, for the very principles that are to be exhibited are prerequisite to seeing the intended phenomenon (p. 512, italics added). (Classified as Level II).

Given the empiricist stance of most science educators and curricula it is difficult for students to understand that all observations are theory laden. This presentation raises some important issues faced by many teachers in different parts of the world and in the context of constructivist practice in the classroom, constitute a dilemma. The issue of relevant scientific principles, as pointed out by the critics of constructivism, is of utmost importance for classroom practice (cf. Matthews, 2015). For example, how would students construct the concept of angular momentum in a constructivist class, without prior exposure to relevant scientific principles and the degree to which the teacher would have to resort to a traditional teaching format. In this context this issue is important as Roth generally considers constructivism to be a more mature form of knowing (cf. Roth & Roychoudhury, 1994, p. 7).

4.2.9  Worldviews An example of the traditional view of Feyerabend’s contribution to understanding philosophy of science is provided by Duschl and Wright (1989): “Thesis III represents the Weltanschaunngen (worldview) analysis science in which science is considered to be a relativistic enterprise. It is exemplified by the writings of Kuhn (1970), Hanson (1969), and Feyerabend (1970a)” (p. 472). (Classified as Level I). These authors consider that such philosophers endorse Thesis III, viz., theories dictate the meaning of observations, and thus espouse a relativistic interpretation (see Fig.  2, p.  471). Chapter 2 presents a more nuanced interpretation of Feyerabend’s views. In an editorial, titled “Slippery slopes of postmodernism” Ronald G.  Good (1993) has criticized the National Research Council for having endorsed postmodernism in their National Science Education Standards. Furthermore, Good is right in pointing out that both logical positivism and postmodernism constitute two

4.2  Results and Discussion

79

extreme views of the nature of science and not conducive to a robust science education practice. However, Good also considers that Michael Foucault and Paul Feyerabend are two gurus of postmodernism. When this editorial was published in 1993, there were sufficient writings of Feyerabend to sustain his critical stance towards postmodernism (for details see Chap. 2). This chapter provides examples of research reported in the Journal of Research in Science Teaching that facilitate a wide range of perspectives with respect to Feyerabend’s epistemological anarchism. Conclusions based on these findings will be integrated with those from other chapters and presented in Chap. 8.

Chapter 5

Understanding Epistemological Anarchism (Feyerabend) in Research Reported in the Journal Interchange (Springer)

5.1  Method Interchange: A Quarterly Review of Education, is published by Springer in collaboration with the University of Calgary, Canada. It started publishing in 1970 and embraces educational theory, research, analysis, history, philosophy, policy, practices and a particular interest in science education. In November 2017, I made an online search on the website of Interchange with the key words “Epistemological anarchism” and “Feyerabend” (www.springer.com/10780). This gave a total of 15 articles published since 1982. All articles were evaluated on the same criteria (Levels I–V) as in a previous study (see Chap. 3). Following the guidelines based on Charmaz (2005), presented in Chap. 3, and in order to facilitate credibility, transferability, dependability, and confirmability (cf. Denzin & Lincoln, 2005) of the results, I adopted the following procedure: (a) All the 15 articles from Interchange, were downloaded and after evaluation were classified in one of the five levels, I–V (for levels see Chap. 3); (b) After a period of approximately three months all the articles were evaluated again and there was agreement of 95% between the first and the second evaluation. It is important to note that all the articles evaluated in this study referred to epistemological anarchism in some context, which may not have been the primary or major subject dealt with by the authors. Detailed examples from different levels are presented in the next section. A complete list of all the 15 articles from Interchange that were evaluated is presented in Appendix 5. Distribution of all the articles according to author’s area of research, context of the study and level (classification) is presented in Appendix 6.

© Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_5

81

82

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

5.2  Results and Discussion Based on the treatment of the subject by the authors 10 categories were developed to report and discuss the results. These categories are presented in alphabetical order. It is important to note that the idea behind creating the 10 categories is to facilitate the reader to find the subject of her/his interest. Given the wide range of subjects discussed by the authors over a period of more than 30 years, it is difficult to create the semblance of a continuous storyline. Similarly, due to limitations of space it is not possible to present a detailed critical analysis of every article. The following are the 10 categories (presented in alphabetical order) that were created to present and discuss the results: 1. African and modern medicine 2. Alternative approaches to growth of knowledge 3. Constructive alternativism 4. Diversity of rival theories 5. Genius in science 6. History of science 7. Objectivity versus subjectivity 8. Presuppositions of science teachers 9. Rationalism 10. Scientific method

5.2.1  African and Modern Medicine In most African and developing countries women’s education is neglected and even if they overcome the odds and graduate they are discriminated against in their professional careers. Woodhouse and Ndongko (1993) interviewed ten women from Cameroon in order to find out how they had managed to become scientists and science educators. Three of the participants had a Ph.D., and worked at a university in Cameroon, one was a medical doctor and six (with a Bachelor’s or Master’s degree) were high school science teachers. It is plausible to suggest that the sample of the study is inclined towards those who would support Western style education. The objectives of the study were to find: (a) about the kinds of support they had been given by their families during their educational career; (b) how science was taught in schools both in the past and at present; and (c) whether or not they thought it possible to integrate science and African traditional thought in schools and universities. Responses of the participants showed the difficulties involved in integrating medical science and traditional medicine of Cameroon: They were uncompromising in their belief that it was traditional doctors who needed to change their methods of investigation and dissemination of information if they wished to work as partners with medical doctors. Moreover, they suggested that the secretive methods of communication that typified the practice of traditional medicine should be remodelled

5.2  Results and Discussion

83

along the lines of the methods of communication used by the medical profession. All of these suggestions presuppose that the medical profession is “open” in its methods of clinical investigation, diagnosis, and prognosis, while traditional medicine remains “closed” in all of these practices. Yet this presupposition that modern Western scientific practices and beliefs are open in character while traditional African practices and beliefs are closed is denied by the often authoritarian nature of Western scientific communities and their resistance to change … At the same time, there are various examples of open methods of discovery and learning among different African traditional beliefs systems (Feyerabend, 1975a …). (Woodhouse & Ndongko, 1993, p. 154). (Classified as Level III).

It is important to note that the defense of these participants of medical doctors could perhaps be attributed to the fact that despite difficulties they had still accomplished their goal of joining the profession. Nevertheless, it is essential that the practices of multinational medical corporations be subjected to close scrutiny.

5.2.2  Alternative Approaches to Growth of Knowledge Providing in-service teachers an opportunity to become familiar with the controversial nature of progress in science (growth of knowledge) is an important objective of most innovative educational systems and curricula. Based on a history and philosophy of science perspective (Feyerabend, Kuhn, Lakatos, and Popper), Niaz (2004) has reported a study based on 41 Venezuelan in-service teachers who had registered for a nine-week course on Methodology of Investigation in Education as part of their Master’s degree program. Teachers were drawn from the following areas of work: English, Spanish, Chemistry, Mathematics and Physics. Based on 20 readings, course activities included: written reports, classroom discussions based on participants’ oral presentations, and written exams. In one of the exam questions teachers were asked to select a topic from their area of work and explain the growth of knowledge through any of the following conceptualizations: Kuhn (1970), Lakatos (1970), Campbell (1988a, 1988b) and Erickson (1986). Three responses from teachers working in different areas of work are reproduced here (for complete details see Niaz, 2004). A mathematics teacher selected the topic of geometry and following is the response based on the work of Erickson: The development of geometry conforms to the thesis of Erickson, namely, the coexistence of Euclidean geometry with the new paradigm (non-Euclidean) based on the geometry of Riemann, Lobatchevsky, and Gauss. The old paradigm held its sway for almost 1500 years until the appearance of the new paradigm in the middle of the 19th century. The new paradigm refuted some of the axioms of Euclidean geometry … Nevertheless, some of the old postulates of Euclidean geometry survived, which we still use in our daily lives … This shows that the new and the old paradigms can coexist without entering into conflict (reproduced in Niaz, 2004, p. 162).

A chemistry teacher selected the topic of atomic models and following is the response based on the work of Lakatos: Evolution of the different atomic models from Dalton to the quantum mechanical model of the atom can be interpreted as competing and rival research programs as conceptualized by

84

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported… Lakatos … Dalton’s model of the atom was based on the hard core idea that the atom was indivisible … Thomson’s experiments led to the postulation of a model of the atom that was divisible. Thus Thomson’s model of the atom had to compete with that of Dalton … Rutherford based on his alpha-particle experiments was led to the nuclear model, which had to compete with Thomson’s model of the atom … Bohr’s model of the atom explained the stability of Rutherford’s nuclear atom and thus the two had to compete. Bohr’s model of the atom could explain only the spectrum of atoms with one electron. Quantum mechanics presented a rival model of the atom that could explain the spectra of atoms with more than one electron (reproduced in Niaz, 2004, pp. 162–163).

An English teacher selected the topic of English as a second language and following is the response based on the work of Kuhn: The development of different methodologies in the teaching of English as a second language adapts perfectly to the Kuhnian conceptualization. At first we had the grammatical perspective … By the end of the 19th century we had a radical change, when this was displaced by the method based on the translation of written texts … In the 1930s, under the influence of Skinnerian behaviourism, this was displaced by the oral approach. This approach had its days of glory during the II world war when it was used by the US government for teaching German and Japanese … During this period the previous approach was thrown in the garbage dump of history—to use a Marxist terminology. The oral approach or the Army method as it came to be known later, reigned absolutely till the 1970s. At this stage as behaviourism was being bashed, the oral approach was criticized for being savagely repetitious, and it agonized until it was displaced by the functional approach with its emphasis on interactions (reproduced in Niaz, 2004, p. 163).

Similar responses were also presented by other teachers. It is important to note that the differences between the four conceptualizations (Campbell, Erickson, Kuhn and Lakatos) are rather subtle and can even be interchangeable. For example, a Kuhnian conceptualization is characterized by abrupt changes that leads to the displacement of one paradigm by another. A Lakatosian conceptualization is characterized by competition between rival research programs instead of paradigms. A conceptualization based on the work of Erickson is characterized by the coexistence of the old and the new paradigms. It is difficult to replace an old paradigm by falsification. A conceptualization based on the work of Campbell is characterized by rivalry between different hypotheses that appear plausible for explaining a phenomenon. It is important to note that these in-service teachers were neither historians of science nor philosophers of science, and had very little previous experience in understanding growth of knowledge critically. However, their teaching experience and a critical perspective helped most of them to understand the nuanced differences between the different conceptualizations. For example, Kuhn emphasizes displacements, Lakatos emphasizes rivalries and Erickson emphasizes coexistence. Interestingly, participants also provided competing interpretations of the same developments in their area of work. For example, a response very similar to the one presented as Kuhnian (teaching of English as a second language) was considered as Ericksonian by another English teacher. Other examples of multiple interpretations of the same developments were also found, and was the subject of considerable discussion in class. Actually, this is how science works (Feyerabend, 1970a)—alternative approaches to understanding growth of knowledge (Classified as Level III).

5.2  Results and Discussion

85

Winchester (1993) has expressed some skepticism with respect to such developments, as they might lead the participants to think that there is no “truth” in growth of knowledge. This will be the subject of discussion in the final Chap. 8 (Conclusions section).

5.2.3  Constructive Alternativism In his personal construct psychology George Kelly (1955) emphasizes the internal world of a person. Kelly was offering a constructivist alternative to the behaviorist framework that was dominant in the 1950s, and for many years his theory was neglected in psychology. According to Kelly, each person constructs for her/him a representational model of the world. The model is subject to change over time, since constructions of reality are constantly tested—quite similar to scientific theorizing. Just as the scientist designs experiments around rival hypotheses, students can be encouraged to do the same. A parallel can be drawn between Kelly and Feyerabend as both considered that any event is open to as many reconstructions of it as our imagination will allow (hence Feyerabend’s advice “anything goes”). Again, both deplored the traditional teaching methods based upon the cultural transmission approach that emphasizes the student’s role as the passive receiver rather than the active participant. Such an approach precisely represents the positivist, empiricist/ inductivist conception of science represented by the naïve realist who looks for “absolute truths” in the scientific endeavor. Based on these considerations, Pope (1982) concluded: Kelly pointed out that all theories are hypotheses created by people and that, although they may fit all the known facts at any particular time, they may eventually be found wanting and eventually be replaced by a “better theory.” An example from physics is the re-appraisal of Newton’s theory by Einstein. However, Einstein’s theory is not the ultimate truth—Einstein himself regarded his theory as defective and spent much of his life trying to find a better one. This view of theory, science, and knowledge is echoed in the writings of Popper (1963a), Kuhn (1970), Lakatos (1970), and Feyerabend (1975a), all of whom argue for a more relativistic picture of knowledge (p. 6, italics added). Classified as Level V.

It is plausible to suggest that “relativistic” in the present context means replacement of one theory by a “better” theory. Interestingly, the role of Newton and Einstein comes up again with respect to “ultimate truth” (see a later section on this subject). If knowledge can be seen as being produced by transactions between a scientist/student and the environment, leading to an active and creative participation, it is not difficult to see how all theories must eventually change—Feyerabend was particularly enthusiastic about such changes, and perhaps overstated his advice that “anything goes.” Generally, such aspects in Feyerabend’s philosophy have been ignored and it is precisely for such reasons that he alluded to them in his famous “How to defend society against science” Feyerabend (1980). He argues cogently that science education should strengthen the minds of the young against any easy acceptance of comprehensive views and to be receptive to alternative

86

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

c­ ounter-­suggestive views (including counterinduction, see Chap. 7). Furthermore, science needs to be taught as an historical phenomenon and not as a set of absolute facts. It is important to note that good grades in school science do not necessarily facilitate an inventive attitude.

5.2.4  Diversity of Rival Theories Brownian motion was first discovered by Thomas Brown in 1827, who observed irregular motion of small suspended particles in various fluids. Since then various attempts were made to understand this phenomenon. Finally, Perrin (1923) demonstrated that Einstein-Smoluchowski theory provided the evidence in favor of the molecular-kinetic theory of gases and against its rival classical thermodynamics. Anti-atomists like E. Mach and W. Ostwald attacked the molecular-kinetic theory by arguing that a phenomenological description such as thermodynamics has enough information and avoids various problems faced by the atomic theory. Einstein (1926) responded in the following terms: If the movement [Brownian] discussed here can actually be observed (together with the laws relating to it that one would expect to find), then classical thermodynamics can no longer be looked upon as applicable with precision to bodies even of dimensions distinguishable in a microscope: an exact determination of actual atomic dimensions is then possible. On the other hand, had the prediction of this movement proved to be incorrect, a weighty argument would be provided against the molecular-kinetic conception of heat (pp. 1–2).

Einstein’s argument was crucial. Mayo (1988) has clarified further: If Brownian motion was caused by something outside the liquid medium or something within the particles themselves then it would coincide with classical thermodynamics. However, as the Brownian motion was caused by a molecular motion in the liquid medium it supports molecular-kinetic theory. Based on these considerations, Brown (1997) concluded: Not only are theories evaluated by means of the evidence relative to their rivals, but what counts as evidence may depend heavily on what rival theories are being considered. Perhaps the best illustration of this is Paul Feyerabend’s famous example of Brownian motion (1962/1981). Early in the 19th Century, a Scottish botanist, Thomas Brown, noticed a remarkable phenomenon. Tiny bits of pollen moved randomly around in a fluid. Are they alive? What’s their source of energy? Does polarized light cause this bizarre motion? (p. 384). (Classified as Level IV).

At the turn of the century, Einstein, Perrin, and others offered an explanation of Brownian motion based on the kinetic theory—a rival of classical thermodynamics. The success of this explanation for the rival kinetic theory put classical thermodynamics in a rival position. Thus Brownian motion became relevant evidence for classical thermodynamics, only because a rival had explained it, which shows the role of diversity of rival theories (see Chap. 7 for further details on Brownian motion).

5.2  Results and Discussion

87

5.2.5  Genius in Science The role played by geniuses in both art and science continues to be controversial. In the nineteenth century it was generally accepted that science was analytical and rational, whereas art was, holistic, intuitive and novel. Hattiangadi (1985) recognizes that in the twentieth century physical theories of Newton have been overthrown by the beautiful, speculative cosmological ideas of a genius like Einstein, and it seems that now even imagination and excitement are also important ingredients in the scientific enterprise. Based on these considerations, Hattiangadi (1985) has pointed out: In this century, there has been a gradual encroachment of the 19th century conception of the artist into interpretations of science as an activity. Arthur Koestler in The Act of Creation (1964) and Michael Polanyi in “Genius in Science” (1972) have explicitly raised the issue of genius in science. But it arises less explicitly in the writings of Paul Feyerabend (1975a) and in the “punctuated equilibrium” model of scientific development that has been suggested by Thomas Kuhn (1962). I think we should all recognize the temptation to find genius in science now that science has become more like the 19th-century conception of art (p. 40). (Classified as Level III).

In a footnote, Hattiangadi elaborates his position further by noting that according to Agassi (1975) as compared to the democratic view of the Enlightenment, that of Polanyi (1972) is elitist and romantic. In the same footnote, it is reported that in a review (Hattiangadi, 1977), he had described Feyerabend (1975a) as an “egalitarian romantic,” a description to which Feyerabend did not object in his reply, notwithstanding the fact that there was an implicit criticism in it. At this stage it is important to note that Polanyi’s (1964) views on genius in science are based on a historical perspective and much more nuanced: The power to expand hitherto accepted beliefs far beyond the scope of hitherto explored implications is itself a pre-eminent force of change in science. It is this kind of force which sent Columbus in search of the Indies across the Atlantic. His genius lay in taking it literally and as a guide to practical action that the earth was round, which his contemporaries held vaguely and as a mere matter for speculation. The ideas which Newton elaborated in his Principia were also widely current in his time; his work did not shock any strong beliefs held by scientists, at any rate in his own country. But again, his genius was manifested in his power of casting these vaguely held beliefs into a concrete and binding form … [same can be said of] diffraction of X-rays by crystals (in 1912) … by a mathematician Max von Laue … These advances were no less bold and hazardous than were the innovations of Copernicus, Planck or Einstein (p. 277, italics added).

This clearly shows that Polanyi considered a genius (Columbus, Copernicus, Einstein, Max von Laue, Newton, Planck and others) to go far beyond simply holding speculative cosmological ideas. Finally, Hattiangadi (1985) concluded that a closer scrutiny of any intellectual development, in both science and art reveals that there is no place for genius, individual creativity, or novelty—this of course, does not coincide with Feyerabend and Polanyi.

88

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

5.2.6  History of Science Positivist philosopher of science and physical chemist Wilhelm Ostwald (1953–1932) was perhaps the first in the twentieth century to emphasize the importance of history and philosophy of science for science education. In the case of chemistry, Smith (1925), co-founder of the American Chemical Society’s Division of the History of Chemistry was perhaps the first to introduce a course on the history of chemistry. A major problem with such courses was that they merely asserted the conclusions and did not emphasize how they were reached and what alternatives were plausibly advocated that led to controversies. Later based on the modern historiography of science, Harvard Project Physics (Holton, 1978b) had considerably more success and acceptance. More recently, Holton (2014a) while recalling why he decided that Harvard Project Physics be based on a humanistic approach stated, “I based my decision in part on the hunch that more beginning students would come to take this course, to learn not only that F is equal to ma, but also that science is a fascinating part of human culture” (p. 1876). Indeed, to recognize that science is not culture free is a humbling experience for scientists. Based on these considerations Brush (1989) has suggested that: … what one learns from history is that one side was victorious for reasons that can be specified, and thus earned the privilege of having its position adopted, not as immutable truth but as the best available working hypothesis on which to base further research. We want to make the point that scientists are often most successful when they act as if some law, theory, or even philosophical principle has been firmly established even though they recognize that it may be drastically revised sometime in the future… At the other extreme, some sociologists and philosophers claim that all knowledge, including scientific facts and mathematical theorems, is “socially constructed.” From this viewpoint science has no claim to a preferred status over astrology, religion, mysticism, creationism, and other doctrines (see Bloor, 1976, 1978; Collins, 1981; Feyerabend, 1975a) … An important task of science education is to strike a balance between these two extremes—to show how science can acquire valid and useful knowledge that is nevertheless a product of human thought, subject to change in the light of new evidence and reasoning (see Holton, 1974). (p. 64, italics added). (Classified as Level IV).

Brush is trying to argue that traditional science teaching looks for objective facts instead of recognizing that scientific research does not provide immutable truths but rather working hypotheses useful for future research. Furthermore, although objective facts are valid and useful knowledge they are subject to change and revision namely the tentative nature of scientific knowledge. In Feyerabend’s framework counterinduction provides alternative working hypotheses that can help to strike a balance between the two extremes. Such considerations can make science education more meaningful for both students and teachers. In most parts of the world the benefits of science are recognized and still its history is ignored, primarily because textbooks are supposed to provide the necessary background. Research in science education has shown that most high school and university science textbooks are particularly misleading and even perhaps distort how science really works (McDonald & Abd-El-Khalick, 2017). Textbooks

5.2  Results and Discussion

89

g­ enerally ignore the difficulties involved in the interpretation of data and theory construction. Based on these considerations, and writing for a special issue of Interchange, related to history of science and science teaching, Winchester (1989) has concluded: This is connected with another highly neglected aspect of science in science teaching, namely the question of the difficulty of framing the concepts useful in science. Two centuries of empiricism have tended to suggest that scientists simply “find” things which are subsequently named. But this bears no relation to what actually goes on (as many of the papers in this volume will argue and as authors such as Popper, Whitehead, Kuhn, Feyerabend, and a host of others have argued earlier) and consequently is deeply misleading to the young student of science. It is mainly through the history of science that one can see how difficult it is to frame concepts which have the necessary generality and suggestiveness to get on with the investigations in science (p. ii, italics added). (Classified as Level III).

The reference to two centuries of empiricism and its relationship to what actually goes on can easily be considered as the influence of the Newtonian method (cf. section in Chap. 3) on scientific practice (cf. Lakatos, 1978 for a similar statement; also Duhem, 1914). Furthermore, Winchester goes on to show that the history of science is important not only for science education but for science itself, by citing two examples: Galileo and Darwin on the one hand, and even Newton and Einstein were fully aware of the work of their predecessors and the controversies surrounding earlier work in their respective fields of expertise. Again, history of science is also necessary for the formation of a democratic citizenry.

5.2.7  Objectivity Versus Subjectivity Bailin (1990) has analyzed the role of creativity in science in the context of the work of Kuhn and Feyerabend and concluded: Another argument proposed by both Kuhn and Feyerabend which may seem to challenge the idea of objective standards is that, in actual scientific practice, subjective factors do often enter into the process of theory choice? This is a claim for which there is ample evidence and one which would, in general, not be disputed by philosophers and historians of science. Science, whatever else it may be, is a social enterprise conducted by human beings of particular genders, races, and classes in particular social and political contexts. But even if we grant the existence of psychological and social factors in scientific practice, this does not mean that science reduces to a totally subjective practice nor is it an argument against the existence of some objective standards … And a crucial task is to distinguish the objective aspects of scientific theory and practice from the subjective (p.  39, italics added). Classified as Level III.

This is an interesting presentation that recognizes the essential tension between objectivity and subjectivity during scientific progress (see the part in italics). Whatever the philosophical stance adopted by a researcher (both in science and science education) it is important to differentiate between the objective and subjective aspects of the scientific endeavor. Recent scholarship has recognized the intricate

90

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

relationship between objectivity and subjectivity (both define each other) and that there is no objectivity without subjectivity to suppress and vice versa (Daston & Galison, 2007). More recently, Niaz (2018) has explored the evolving nature of objectivity in the history of science and its implications for science education. With this perspective it would be interesting to reconsider and evaluate Feyerabend’s thinking on the subject. This will be the subject of discussion in the final Chap. 8 (Conclusions section) of this book.

5.2.8  Presuppositions of Science Teachers After teaching history and philosophy of science (HPS) to science teachers, for many years, in Canada, Ian Winchester (1993), Editor of Interchange, found that at the beginning of the course most of his students had the same standard picture of science based on the following (among others) presuppositions (quite similar to Karl Pearson’s The Grammar of Science): (1) All human problems should be tackled using scientific method, which constitutes the five or six steps found in most school science textbooks and reproduced in the format of all laboratory reports; (2) Science is value-free; (3) Closer a science approximates to physics, the better a science it is; (4) Science is concerned with producing laws and a scientific law is a general truth about the world; (5) Given scientific laws, explanations of what happens can be given by reference to the laws; (6) Science is an enterprise in which numbers and measurement are crucial, so qualitative descriptions should be avoided; (7) Science is entirely empirical and one does not make up stories in advance of observation; (8) Science is essentially ahistorical, since what matters are its truths, not how one got there; and (9) A scientist is neutral in the face of facts, uncommitted in advance, and open-minded. It is interesting to note that despite all the reform efforts over the last twenty-five years, most students, science teachers and science textbook authors in most parts of the world hold similar beliefs (Grammar of Science) with respect to nature of science. Next, Winchester (1993) summarizes his experience of teaching HPS for twenty years in the following terms: The standard fare of Kuhn, Popper, and Feyerabend (not to mention the up-to-date debate in the journals, which is largely a debate on their themes used by those of us who teach science teachers doing advanced work) challenges each and every one of these presuppositions. So naturally I have considered myself successful as a philosopher of science if I have convinced my students to be doubtful about all of the above things, which they took for granted before they came in contact with me (pp. 194–195, italics added). (Classified as Level III).

Given the considerable debate, controversy and differences among Kuhn, Popper and Feyerabend, it is significant (and even sound educational practice) that Winchester included these three philosophers in the standard fare. In other words, despite the differences students can find common elements among these philosophers that also provided alternative frameworks. Interestingly (and even perhaps

5.2  Results and Discussion

91

unexpectedly), however, Winchester is skeptical of his success in teaching students to question their presuppositions based on twentieth century HPS, and he expresses his concern in cogent terms: Teaching students that there is no truth in science may lead them to consider the pursuit of science as futile and may even “kill” science itself. Based on these considerations Winchester (1993) raises the following dilemma: “Why should we expect them to believe, if we teach them that there is, in the end, no truth in science but that there is truth in the history of science or the philosophy of science (which surely rest on shakier ground than, say, the bulk of physics or chemistry or mathematics)?” (p. 196). Some readers consider that the inclusion of Feyerabend’s perspective in science education may lead to an “erosion of trust in science.” To a certain extent this apprehension is justified. Nevertheless, it is also important to ask: trust in which “science?” Let us suppose that the answer is: Karl Pearson’s The Grammar of Science (cf. Winchester’s 1993, experience mentioned earlier). Most reform documents, for example the Next Generation Science Standards, NGSS in the U.S. (NGSS Lead States, 2013) would not only question Pearson’s Grammar of Science, but instead emphasize the nature of science (NOS) based on a history and philosophy of science perspective, namely the epistemological dimensions of scientific knowledge (including Feyerabend, among other philosophers of science). Recent research in science education has precisely emphasized the importance of NOS (cf. Matthews, 2014a). Despite this emphasis recent research has also reported that, “The textbooks devote miniscule attention (as measured in pages) to epistemological dimensions of scientific knowledge and the workings of the scientific enterprise. It is very hard to imagine that students would engage in any meaningful thinking about NOS themes, which have been consistently touted as central to precollege science education reform efforts of the past 60 years …” (Abd-El-Khalick, Belarmino et al., 2017, p. 53). This clearly shows the difficulties involved, if we want science to be more meaningful for the students and consequently science education research has to explore new and alternative approaches. It is precisely in this context that Feyerabend’s epistemological perspective provides an alternative to traditional approaches, such as Pearson’s Grammar of Science. Furthermore, a major objective of the journal Science & Education (Springer) is based on the premise that science education needs alternative epistemological approaches. This book promotes a new approach to teaching science that facilitates an understanding of various issues beyond that found in traditional science textbooks, such as levels of validity in science, robustness of scientific theories and verisimilitude. Some physicists do consider the laws of physics as true principles and Weinberg (2001) a Nobel Laureate in physics represents such thinking: “What drives us onward in the work of science is precisely the sense that there are truths out there to be discovered, truths that once discovered will form a permanent part of human knowledge” (p. 126, emphasis added). If science looks for truths, then Newton’s laws should have been the prime example. However, at the beginning of the twentieth century, Einstein’s theory of special and general relativity and later quantum mechanics questioned Newton’s perspective. Does this mean that Newton’s laws were false or even that perhaps he was not being objective? Giere (2006a), a

92

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

p­ hilosopher of science, has questioned such philosophical positions as “objectivist realism” (p. 5) and explained cogently: Weinberg should not need reminding that, at the end of the nineteenth century, physicists were as justified as they could possibly be in thinking that classical mechanics was objectively true. That confidence was shattered by the eventual success of relativity theory and quantum mechanics later (p. 118).

At this stage it is important to note that Giere (2006a) is skeptical and questions “true” theories in both science and history and philosophy of science (HPS), and instead endorses “perspectival realism”: … I wish to reject objective realism but still maintain a kind of realism, a perspectival realism, which I think better characterizes realism in science. For a perspectival realist, the strongest claims a scientist can legitimately make are of a qualified, conditional form: “According to this highly confirmed theory (or reliable instrument), the world seems to be roughly such and such.” There is no way legitimately to take the further objectivist step and declare unconditionally: “This theory (or instrument) provides us with a complete and literally correct picture of the world itself (Giere, 2006a, pp. 5–6).

In the light of this discussion, it is plausible to suggest that Winchester’s (1993) qualms with respect to truth in science and HPS are not warranted or at least misplaced. Instead, as science teachers we should endeavor to facilitate an understanding of “how science works”, and that there is no truth either in science or in HPS. In order to explore this subject further, Niaz (2016, see Chap. 3, for complete details) has designed a study in which in-service science teachers enrolled in a doctoral program participated in various activities based on a history and philosophy of science perspective (Popper, Kuhn, Lakatos, Feyerabend, Laudan, Cartwright and Giere). After the course all participants responded to a 3-item questionnaire, of which the following was one item: Many scientists, science textbook authors and professors believe that science is ‘objective.’ If we accept this perspective, Newton’s laws constitute the best example of objectivity in science. Nevertheless, at the beginning of the 20th century, Einstein’s theories of relativity (special and general) questioned Newton’s laws. Accordingly, do you think that Newton’s laws are false and consequently that he was not ‘objective’? (Reproduced in Niaz, 2016, p. 60).

Background to this item is provided by Giere’s (2006a) critique of those scientists and philosophers of science who consider that what drives scientists onwards is that there are truths out there to be discovered, and that such philosophical positions can be considered as “objectivist realism.” Of the 12 participants, 10 stated that Newton’s laws were not false and that he was “objective” in the formulation of his laws, and following is one example: First it is important to recognize that Newton molded his vision of the material world based on the law of universal gravitation, thanks to the work of scientists such as T.  Brahe, N. Copernicus, J. Kepler, and G. Galilei. Was Newton objective in the formulation of his theory? He thought that he was and many believed that his vision was the last word with respect to this problem. However, Einstein demonstrated with his theory of relativity that Newton was not sufficiently objective as his theory could not explain certain phenomena that the theory of relativity could. But thanks to Newton, Einstein could see beyond Newton.

5.2  Results and Discussion Are Newton’s laws false? In physics it is known that these laws are not fulfilled in the context of Einstein’s physics and consequently are not objective in this context. Nevertheless, these days Newton’s laws continue to be applied, and consequently, I think that in a certain sense these laws have ‘some degree of truth’ in their natural context of application. Was Einstein objective? Until now history tells us that he was. For how long? We still do not know (Participant #2, reproduced in Niaz, 2016, p. 65). Most philosophers of science (including Duhem, Giere, Kuhn, Lakatos, Laudan and Feyerabend) would agree that if a scientific theory is replaced by another with greater explanatory power, it does not mean that the previous theory was either false or that its author was not being ‘objective.’ This is the dilemma faced by the participants in this item. In other words, Newton’s laws when first proposed in the 17th century were ‘true’ for that epoch (actually for more than 200 years) and he was as ‘objective’ as one could possibly expect a scientist to be. Consequently, the solution to the dilemma lies in recognizing that both Newton and Einstein were being ‘objective’ and provided theories that varied in their explanatory power in certain domains (e.g., Einstein explained better the behavior of particles approaching the velocity of light). With this background it is easier to understand the responses provided by the participants of this study. It seems that a majority (10 out of 12) of the participants had a fairly good understanding of the role of ‘truth’ of a theory and consequently the ‘objectivity’ of the scientist. Following Giere (1999, 2006a, 2006b) scientific theories are not ‘true’ or ‘false’ and similarly the role of the scientist is more perspectival rather than ‘objective’. At this stage it would be interesting to have a closer look at the response provided by Participant #2, who tries to understand Newton’s contribution in a historical context by recognizing the work of Brahe, Copernicus, Kepler and Gelilei, which is a sound approach. However, this participant is clearly struggling to understand the dilemma, as she/he asks, ‘Was Newton objective in the formulation of his theory?’ and again responds in a historical context by pointing out that, ‘many believed that his vision was the last word with respect to this problem.’ Next this participant reminds us that ‘But thanks to Newton, Einstein could see beyond Newton’, and this helped to respond to the question, ‘Are Newton’s laws false?’ Finally, this participant raises a thought provoking question, ‘Was Einstein objective?’ and responds laconically, ‘For how long?’ In my opinion, this line of reasoning approximates to a historical reconstruction of the Newton-Einstein debate. The underlying issue that this participant is trying to understand is the following: Is Newton’s theory still “true” after Einstein’s theory of relativity? Furthermore, is Einstein’s theory here to stay and “true” forever? Worrall (2010), a philosopher of science has referred to the same problem in the following terms: “On what grounds, then, could the realist deny the possibility that Einstein’s theory might itself eventually be replaced by a theory bearing the same relation to it as it does to Newton’s …” (p. 288). The similarity between the thinking of Participant #2 and Worrall is striking, indeed. In other words, it is plausible to suggest that the course on HPS provided the impetus and background for developing such ideas, namely understanding “how science works.” I am sure Winchester would agree that cultivating such ideas and thinking is what science education is all about. Consequently, contrary to what Winchester suggested, questioning the presuppositions of our students would not kill science but rather facilitate an understanding of how science works. Interestingly, even in his early work Feyerabend (1968) questioned methodological rules and instead emphasized the practice of science based on thoroughgoing pluralism (e.g., Newton, Einstein, what next) in order to check dogmatism. Indeed, history and philosophy of science can help science educators to facilitate an understanding of the practice of science.

93

94

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

At this stage it would be interesting to consider what Einstein himself had to say with respect to Newton: Newton, forgive me; you found the only way which in your age was just about possible for a man with the highest powers of thought and creativity. The concepts which you created are guiding our thinking in physics even today, although we now know that they will have to be replaced by others farther removed from the sphere of immediate experience, if we aim at a profounder understanding of relationships (Einstein, 1949, p. 684, italics added).

This is a very interesting statement, as on the one hand it recognizes the debt that modern science owes to Newton and at the same time calls for a change in our thinking. Furthermore, it is important to note that Einstein refers to “highest powers of thought and creativity” and not the “Newtonian method” which has been the subject of criticism in the history and philosophy of science literature (cf. Duhem, 1914). Once again, the similarity between Einstein’s thinking, Worrall (2010) and Participant #2 (referred to above) is striking, namely despite the merits of our existing theories, new theories will always replace them.

5.2.9  Rationalism J. Agassi’s (1996) review of Feyerabend’s autobiography Killing Time reveals some interesting aspects of the latter’s life, work and interpersonal relations. This review is important as both of them had a close friendly relationship for many years. Overall, although Agassi considers Feyerabend to be an enfant terrible, but still successful as a philosopher of science, and following are some of the important features of the review: (a) Feyerabend tried to replace abstract terms like “truth” and “objectivity” with “democracy”, however, to no avail; (b) While working in the Nazi army Feyerabend witnessed some atrocities and reacted by emotional impassiveness— perhaps it had something to do with his lower middle class background of his childhood; (c) Feyerabend had the chance to meet Martin Heidegger (the famous Nazi philosopher) but refused; (d) Feyerabend translated Popper’s Open Society. Popper, however, did not like the result and had to rework the whole translation; (e) Feyerabend’s explanation of incommensurability. It is a technical term introduced by Thomas S. Kuhn and Feyerabend in the same year, initially to denote an idea concerning relative truth. However, the idea was first suggested by Pierre Duhem that is two different theories of different periods, cannot be contrasted even if they deal with the same facts. Thus we cannot say that since quantum mechanics is true, Newtonian mechanics is untrue; and (f) Feyerabend rightly attributes the origin of the claim that “there is no scientific method” to Popper. Finally, Agassi (1996) concludes the review by stating that: After all, he did say such things as, Farewell to Reason, and here he repeats it, but while denying that he ever argued simply against reason, rationality, and even rationalism: he argued only against that kind of rationalism – rigid, pompous, harsh (p. 88). (Classified as Level III).

5.2  Results and Discussion

95

Although, Agassi is skeptical of Feyerabend’s thinking with respect to rationalism, actually he repeated such ideas in a particular context (rigid, pompous and harsh) at various places in his writings (see Chaps. 1 and 2). For further details on the origin of Feyerabend’s thinking on rationalism see Shaw (2019).

5.2.10  Scientific Method Traditional reporting of science is primarily preoccupied with the “context of justification,” and almost completely ignores the “context of discovery,” and thereby presents a mistaken conception, and perhaps even a travesty, of the nature of scientific thought (Medawar, 1967). This formal style of writing consistently obscures the role of intuitive, inspirational insights, tentative attempts and ambiguous formulations, prior intellectual commitments, and social interactions with other researchers. On the contrary the positivist myth has led some scholars to believe that conformity with the rules of the scientific method will yield scientific knowledge that is “true” and perhaps even beyond criticism. Based on these considerations Rampal (1992) has criticized the role played by the scientific method: This indeed has been the naïve yet popular conception of the process of scientific creation, and has, surprisingly, continued to hold sway, despite major philosophical shifts caused by the work of Kuhn, Popper, Lakatos, and Feyerabend, among others (p. 310, italics added). (Classified as Level III).

It is important to note that despite their differences, Feyerabend is included along with Kuhn, Popper and Lakatos as responsible for major philosophical shifts. The relationship between the scientific method, induction and plurality of methods in the context of the philosophical writings of Dewey and Popper has been explored by Swartz (1985). In this context it is important to consider: Should inductive procedures be taught to students as a significant part of the method that scientists use to discover and test hypotheses about the observable world? This is a controversial question and according to Swartz (1985, p. 45) eventually may have to be rejected as it is based on the erroneous assumption that inductive procedures are a significant part of the method scientists use to test ideas. The controversy resides in the fact that Dewey responded affirmatively to this curriculum question and Popper in the negative. Actually, differences with respect to the inductive method are widespread and difficult to resolve. As an illustration, Swartz (p. 44) considers that Dewey could argue that theories were pragmatic truths that had evolved from the truths discovered by Newton, whereas for Popper, Einstein’s ideas were more satisfactory than Newton’s because they had survived the tests that had falsified Newton’s theories. Of course, the falsificationist approach is of prime importance in Popper’s oeuvre. Finally, Swartz (1985) concluded: When educational programs use the liberal idea of the toleration of diversity as the basis for studying scientific method, schools become social institutions which accept the idea that methodological pluralism is both reasonable and desirable, [footnote, Feyerabend, 1981,

96

5  Understanding Epistemological Anarchism (Feyerabend) in Research Reported… pp.  65–79]. Although Dewey and Popper endorse many of the ideals associated with Western liberalism, neither argues that the toleration of diversity should be used as a guiding principle for organizing learning situations at all levels of schooling. But if methodological pluralism becomes an intricate part of school life, it will then be possible for all school members to study the ideas of Dewey, Popper, or anyone else who might shed light on how best to advance learning (p. 47, italics added). (Classified as Level V).

On the one hand, Swartz associates the liberal idea of toleration of diversity with Feyerabend’s (1981) endorsement of methodological pluralism, as it is reasonable and desirable. Secondly, both Dewey and Popper lack this perspective of plurality of methods. In other words if Popper questions induction—this is not enough, unless accompanied by alternative/rival methodologies. Despite this neglect on the part of two of the major philosophers, Swartz recommends that methodological pluralism needs to form an intricate part of educational practice. This clearly shows the importance of Feyerabend’s approach of counterinduction. This chapter provides examples of research reported in the journal Interchange that provide a series of different perspectives with respect to Feyerabend’s epistemological anarchism. Conclusions based on these findings will be integrated and synthesized with those from other chapters and presented in Chap. 8.

Chapter 6

Understanding Epistemological Anarchism (Feyerabend) in Research Reported in Reference Work

6.1  Method This chapter evaluates research reported in the International Handbook of Research in History, Philosophy and Science Teaching (HPST). HPST is the first handbook (www.springer.com 978-94-007-7653-1) devoted to the field of historical and philosophical research in science and mathematics education. The handbook has 76 chapters written by 125 authors from 30 countries, which makes it truly an international endeavor. More than 300 reviewers from the disciplines of history, philosophy, education, psychology, mathematics, and natural science contributed with their expertise to its elaboration. In order to understand the rationale of the handbook it is important to consider the following invitation that was sent to the prospective authors of the different chapters: The guiding principle for the Handbook chapters is to review and document HPS [History and philosophy of science]-influenced scholarship in the specific field, to indicate any strength and weaknesses in the tradition of research, to draw some lessons from the history of this research tradition, and to suggest fruitful ways forward … The expectation is that the handbook will demonstrate that HPS contributes significantly to the understanding and resolution of numerous theoretical, curricular and pedagogical questions and problems that arise in science and mathematics education (Matthews, 2014b, p. 7).

This clearly shows the wide ranging and multiple objectives of the Handbook that can provide guidance for future research as well as curricular and pedagogical feedback to those working in the educational field. Based on the subject index of the handbook, I found 10 chapters that referred to “Feyerabend” or “epistemological anarchism.” However, 4 chapters made only a simple mention with no elaboration and thus were not included. Six chapters discussed some aspect of epistemological anarchism or Feyerabend. Following the guidelines based on Charmaz (2005), presented in Chap. 3, and in order to facilitate credibility, transferability, dependability and confirmability (cf. Denzin & Lincoln, 2005), of the results I adopted the following procedure: (a) All the 6 chapters from the International Handbook of Research © Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_6

97

98

6  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

in History, Philosophy and Science Teaching (HPST) were evaluated and classified in one of the five levels (I to V, see Chap. 3 for levels); and (b) After a period of approximately 3  months all the articles were evaluated again and there was an agreement of 95% between the first and the second evaluation. It is important to note that the authors of these chapters were not necessarily writing about epistemological anarchism, but rather referred to it in the context of their selected topic. Appendix 7 provides a complete reference to each of these 6 chapters that can provide readers with an overview of the topic of interest. Distribution of all the articles according to author’s area of research, context of the study and level (classification) is presented in Appendix 8.

6.2  Results and Discussion Each of the 6 articles from the Handbook was evaluated (Levels I–V) with respect to the context in which they referred to epistemological anarchism. Based on the treatment of the subject by the authors six categories were developed to report and discuss the results (cf. guidelines presented in Chap. 3, from Charmaz, 2005). These categories along with examples are presented in alphabetical order. It is important to note that some of the articles could easily be placed in more than one category. The idea behind the creation of six categories is to facilitate the reader to find the subject of her/his interest. The Handbook has a readership and contributors that include science educators, historians, philosophers of science and sociologists that cover many areas of expertise in the science curriculum. Given the wide range of subjects discussed by the authors, it is difficult to create the semblance of a continuous storyline. Complete information about each article, its evaluation and the author is provided in Appendices (7 and 8), which can be consulted by the interested readers. Following categories (n = 6) with detailed examples of different levels are presented (in alphabetical order) in the next section: 1 . Historical-investigative approach to science 2. Kuhn and normal science 3. Nature of science 4. Postmodernism 5. School science and curriculum 6. Science as cultural tyranny

6.2.1  Historical-Investigative Approach to Science Recent research in the history of science has stressed the idea that scientific experimentation is a multifaceted activity with many possible relationships between observation, inference and theory development. Based on this perspective a ­scientific

6.2  Results and Discussion

99

experiment is an act of intervention (cf. Hacking, 1983) in which material and theoretical entities interact within a cultural and societal context (for further details on Hacking and intervention see Niaz, 2018, chap. 6). With this background Heering and Höttecke (2014) have explored implications for science education and concluded: “… science should be taught as an exemplar of how knowledge is generally acquired in the empirical sciences. While the first idea is still accepted today, the idea of a clear-cut, single epistemic methodology appears to be obsolete. This does not mean that science is not driven by a limited number of methodological rules, like, for instance, the use of controlled experiments …” (pp. 1479–1480). (Classified as Level III). In other words the authors criticize the scientific method on the one hand but still recognize the importance of some methodological rules, and draw support from Feyerabend. Furthermore, these authors raise an important issue by pointing out that manual procedures cannot be explicitly communicated because of their tacit nature (p. 1479). Interestingly, this difficulty is attributed to Collins (1985) and Polanyi (1966). A possible relationship between the ideas of Feyerabend and Polanyi has been suggested by Preston (1997).

6.2.2  Kuhn and Normal Science Kuhn’s normal science has been the subject of considerable discussion and controversy in both science education and the history and philosophy of science. According to Lakatos (1970, p. 155), normal science is nothing but a research program that has achieved monopoly and that happens very rarely in the history of science. Similarly, Popper (1970) voiced a concern with respect to teaching normal science at both the secondary and university level. Siegel (1979), a philosopher of science with considerable interest in science education is particularly critical of Kuhn, “Put concisely Kuhn’s view is that science education does, and should, distort the history of science” (p. 111) and textbooks in most parts of the world are designed to perpetuate normal science. Collins (2000) has attributed the difficulties involved in understanding science to the “falsified history” found in most textbooks. Kuhn (1970, p. 165) himself suggested that a student of physics did not have to study the works of Newton, Faraday, Einstein or Schrӧdinger, as these are already recapitulated in our present day textbooks. Kuhn ignores the point that in order to understand the “how” and “why” of scientific progress it is important to emphasize the role of presuppositions, contradictions, controversies and speculations, which contrasts with normal science. Machamer, Pera and Baltas (2000) have recognized that while nobody would deny that “science in the making” has been replete with controversies, and still it is depicted as the “rational human endeavor par excellence” (p. 3). Recent research in science education has recognized the importance of historical episodes in order to go beyond a “rhetoric of conclusions” (Schwab, 1974) and facilitate conceptual understanding (Matthews, 2014a, 2014b). With this background it is important to note that Taber (2014) considers that many controversies concern issues that are not linked to core epistemological

100

6  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

c­ ommitments within a research tradition and so can be accommodated in Kuhn’s normal science and consequently: “From this view, criticism of Kuhnian normal science as a description of most scientific activity would not undermine what Kuhn has to say about the research training of individual new scientists, …” (p.  1848, original italics). This defense of normal science contrasts somewhat sharply with what most philosophers of science and science educators consider to be Kuhn’s perspective on textbooks and training of new scientists. Furthermore, it is not clear if Kuhn himself would have subscribed to this interpretation of his views. Next, Taber’s views seem to be incongruent to what Matthews (2014a), Editor of the Handbook, considered to be its guiding principle (see quotation at the beginning of Chap. 6). In other words, if a criticism of Kuhnian normal science is not necessary then what was the Handbook supposed to accomplish, especially with respect to science textbooks, training of scientists and science education in general. Despite this incongruence, finally, Taber (2014) concluded: Indeed, Feyerabend (1988) countered the notion of normal science by claiming that the history of science suggested that there was no standard method or set of preferred approaches in science, but rather that scientists were much more pragmatic, adapting and inventing method to meet the needs of the problem at hand (p. 1847). (Classified as Level III).

This leaves the reader wondering with respect to what was Taber’s (2014) purpose in first “defending” Kuhn’s normal science and later invoking Feyerabend to counter the very conception of normal science.

6.2.3  Nature of Science In a section entitled, “Philosophy of education and the nature of science”, Schulz (2014) has asked a very pertinent question: who defines science for science educators? (p. 1280). Next he goes on to suggest some of the possible candidates: scientific experts, philosophers of science, historians, sociologists, scholars working within cultural and women’s studies, postmodern thinkers or students and teachers themselves. Given the lack of consensus among philosophers of science and other experts, science educators are faced with a difficult problem: These polarized camps have made the business of science education a messy and complicated affair—it has become increasingly difficult to navigate a pedagogical course between competing views “from diehard realism to radical constructivism” (Rudolph, 2000, p. 404). At best consensus can be found that several common classroom myths must be exposed, including talk of “scientific method” (Bauer, 1994; Feyerabend, 1975a; Hodson, 1998; Jenkins, 2007). Teachers clearly require substantial philosophical background to familiarize themselves with the issues, but even if consensus could be achieved (which seems unlikely), the question cannot be solely confined and determined on HPSS [History, philosophy & sociology of science] grounds (Schulz, 2014, p. 1281, italics in the original). (Classified as Level III).

Indeed, “scientific method” could have been a good candidate for exploring alternative approaches to teaching science, so that students and teachers (why not?) could

6.2  Results and Discussion

101

approximate to what Hoffmann (2014), referred to as, “By analyzing exactly how scientists approach scientific literature, I hope to reveal the humanity of the scientific method” (p. 323). Hoffmann is a Nobel Laureate in chemistry with considerable interest and expertise in the history and philosophy of science. Nevertheless a recent study (Niaz, 2018, chap. 6) found that of the 60 general chemistry textbooks (published in U.S.A., see Appendix 9) analyzed only 18% presented a satisfactory account of scientific method, beyond that of a myth. In other words, despite the consensus, school science still persists in holding some myths. This clearly shows the difficulties involved in changing classroom practice. Schulz (2014, p. 1282) has suggested that the classroom teacher needs the collaboration of the historian (for correcting pseudo-history in textbooks) and the philosopher of science (for correcting misleading epistemology). This seems to be a sound strategy and Niaz (2016) has referred to it as the integration of domain-specific and domain-general aspects of the curriculum for introducing nature of science (NOS), and designed a study based on in-service science teachers in which historical episodes were discussed, in the context of scientific method. Following is a response from one of the participating teachers: Scientific method is an implement used by every investigator to obtain information related to a problem in the natural and social sciences. However, in the history of the natural sciences it has been found that the scientific method as understood in the scientific community has not been followed in a strict and rigorous manner. One example is the discovery of charge of the electron (1.601 × 10−19C), based on the “oil drop experiment.” There is evidence that Millikan discarded data obtained in his experiment, which means that he did not follow or respect the scientific method rigorously and still his findings are to this day accepted by the scientific community (Reproduced in Niaz, 2016, p. 61, italics added).

In a sense this response approximates to what Hoffmann (2014) has referred to as “humanizing the scientific method.” First, this teacher recognizes the importance of scientific method for both the natural and social sciences. Of course, this forms part of her/his classroom environment and culture. Next, she/he incorporates new elements in the pseudo-history found in most textbooks and discussed in class. Most important, however, is the recognition that although Millikan did not follow the scientific method (in the strict sense) and still his findings are accepted by the scientific community (this required changes in the traditional epistemology found in textbooks). The oil drop experiment forms an important part of the determination of the elementary electrical charge in science curricula of most countries (domain-specific aspect). However, the pseudo-history taught in most textbooks does not refer to the difficulties involved (as the experiment was not simple and straightforward, as suggested by the textbooks, cf. Niaz, 2015), nor that a bitter controversy ensued between Millikan and Ehrenhaft that lasted for many years. In his published scientific papers Millikan did not refer to his discarding data. Consequently, the participating teacher’s conclusion that he did not follow or respect the scientific method rigorously, clearly shows the domain-general aspect. In other words, it is the integration of domain-specific and domain-general aspects that facilitates an understanding of NOS.

102

6  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

6.2.4  Postmodernism Postmodernism is a controversial subject for science educators. Mackenzie, Good and Brown (2014) have presented an appraisal of postmodernism from three different perspectives (one by each author). Ron Good considers the “wispy world of postmodernism” to be based on “slippery slopes” that have very little insight to offer to science education. Next he refers to Holton’s (1993), Science and Anti-­ Science, which considers that those who espouse postmodernism form a loose consortium advocating nothing less than the end of science itself. Good considers that this loose consortium in science education consists of radical constructivism, queerism, variants of multiculturalism, some versions of feminism and the new journal “Cultural Studies of Science Education” CSSE (with Ken Tobin and Wolff-­ Michael Roth as founding editors). Those espousing CSSE consider the method of decomposing unitary systems into sets of variables as insidious. On the contrary, Good considers that the work of Galileo on falling bodies (dissolving into variables on an inclined plane) and of Newton on planetary motion (moon, earth, sun orbits), precisely shows that “without abstraction and idealization [controlling variables], science goes nowhere” (p. 1060). Jim Mackenzie refers to the postmodern dogma that science is a cultural product of Western societies (p. 1068). To counter this belief he recounts the experience of the renowned biologist Ernst Mayr, who lived with a tribe of Papuans in the mountains of New Guinea. He found that the local woodsmen coincided on 136 names for the 137 species of birds distinguished by him (confusing only two species). Contrary to Feyerabend and some postmodernists, this led Mackenzie to conclude that often Western and indigenous science agree to an extraordinary extent. James Brown emphasizes the role of scientific reason, observation and objectivity in scientific progress and how it has been ignored by postmodernism. According to J. Brown: Though Duhem and Popper challenge some aspects of the standard picture of science, they do not quarrel with those features that are most central, namely, the idea that reason and observation play a dominant role in theory evaluation. The postmodern challenge is really quite different. The very idea of scientific reason and objectivity is at issue. Consequently, when we talk about the standard picture of science, we will include Duhem and Popper and almost every other major philosopher of science as embracing that picture. Of course, they differ significantly in detail, but they all hold that reason and observation are at least in principle objective and play a dominant role in science (Reproduced in Mackenzie, Good & Brown, 2014, p. 1070, italics added). (Classified as Level III).

In their standard picture of science, these authors include most prominent philosophers of science as holding some version of it, such as: Whewell, Mill, Mach, Poincare, Pierce, Duhem, Russell, Carnap, Neurath, Popper, Quine, Lakatos, Putnam, van Fraassen and many others. However, they were somewhat hesitant to include Kuhn and Feyerabend as they are often seen as postmodern. Furthermore, the role of objectivity, reason and observation has been the subject of considerable debate and controversy in the history and philosophy of science literature and has led to changes in the standard view of science. For example, what

6.2  Results and Discussion

103

was the reason behind the experimental observations of Robert Millikan and Felix Ehrenhaft in the determination of the elementary electrical charge, in the early twentieth century? Both had very similar experimental data and still their interpretations (reason) were entirely different, and the controversy lasted for at least 10 years (for details see Holton, 1978a, 1978b; Niaz, 2005, 2015). A similar controversy ensued after E. Rutherford and J.J. Thomson did alpha particle experiments in the early twentieth century and postulated entirely different atomic models (for details see Heilbron, 1981a, 1981b; Wilson, 1983; Niaz, 2009). Cooper (1970) a practicing physicist, Nobel Laureate, has explained the reasons of Rutherford and Thomson that led to the controversy, “Rutherford calculated that from the large Thomson positive charge distribution [alpha] particles should never be deflected more than 0.03 degrees in a single collision; in undergoing multiple collisions they should have about an equal chance of being deflected one way as another. Therefore, large angle deflections as a result of many single deflections in the same direction were very improbable” (p. 321). Actually, Rutherford reasoned that based on the Thomson model a total deflection greater than 90° in traversing the metal foil would have only one chance in 103500 of occurring. Did Thomson accept Rutherford’s reason for understanding alpha particle experiments and consequently the nuclear model of the atom? For anyone familiar with the history of science, the answer was in the negative and the controversy dragged on for many years. Thus, reasons are important, but scientists are generally quite adamant in following their own reasons, based on their presuppositions. Again undoubtedly, observations are important, but the history of science shows that they are generally theory-laden, and hence lead to rival theories and thus the importance of pluralism for Feyerabend. Actually, Feyerabend would go beyond by postulating counterinduction, namely proposing theories for which there is little experimental evidence. Interestingly, as early as 1861, Charles Darwin had seen this problem with much acumen by stating: About thirty years ago there was much talk that geologists ought only to observe and not theorize; and I well remember someone saying that at this rate a man might as well go into a gravel-pit and count the pebbles and describe the colors. How odd it is that anyone should not see that all observation must be for or against some view if it is to be of any service! (Letter written to Henry Fawcett, September 18, 1861, in Charles Darwin, Collected correspondence, 21 volumes. Cambridge University Press, Vol. 9, p. 269, italics added).

Again, understanding of objectivity in the standard view of science has also undergone changes and the following statement from Daston and Galison (2007) is quite thought provoking: To grant objectivity a history is also to historicize the framework within which much philosophy, sociology, and history of science has been cast in recent decades. The opposition between science as a set of rules and algorithms rigidly followed versus science as tacit knowledge (Michael Polanyi with a heavy dose of the later Ludwig Wittgenstein) no longer looks like the confrontation between an official ideology of scientists as supported by logical positivist philosophers versus the facts about how science is actually done as discovered by sociologists and historians. Instead, both sides of the opposition emerge as ideals and practices with their own histories—what we have called mechanical objectivity and trained judgment (p. 377, italics added).

104

6  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

This historicized version provides a more nuanced vision of objectivity and its evolution in the history of science. Interestingly, Daston and Galison (2007, p.  478) consider the resolution of the controversy with respect to the determination of the elementary electrical charge (mentioned above) between Millikan and Ehrenhaft as an example of trained judgment and not just objectivity, or for that matter mechanical objectivity. For further details and educational implications of this historical episode, and the evolving nature of objectivity see Niaz (2018). Readers must have also noted that Mackenzie, Good and Brown (2014) do not include Polanyi and Wittgenstein in their list of philosophers of science that represent their standard view of science. Interestingly, J. Brown had included Feyerabend as a postmodern philosopher of science. Despite some obvious affinities of Feyerabend with postmodernism, eccentric ideas such as traditional medicine and astrology, he promoted pluralism and took delight in diversity. Recent philosophy of science has reevaluated Feyerabend’s oeuvre and found that it was imbued with a humanitarian vision of science and thus was more modern (in the Enlightenment tradition) than postmodern (for details see the section “Feyerabend and recent philosophy of science” in Chap. 2).

6.2.5  School Science Curriculum In order to suggest guidelines for school science curriculum, Hodson (2014) compares the philosophical stances of Harvey Siegel and Israel Scheffler. According to Siegel (1991, p. 45), contemporary research has revealed that a scientist is driven by prior convictions and commitments and thus frequently she/he is quite unable to recognize evidence for what it is. In stark contrast, Scheffler (1967, p. v) has argued that scientific change is a product of evidential appraisal and logical judgment. Based on these considerations Hodson (2014) has raised the following questions for science educators, in the context of teaching nature of science (NOS) in the classroom: In building a school science curriculum, are we faced with a stark choice between the traditional and the postmodern? Are we required to choose between the image of a scientist as a cool, detached seeker-after truth patiently collecting data from which conclusions will eventually be drawn, when all the evidence is in hand, and that of an “agile opportunist who will switch research tactics, and perhaps even her entire agenda, as the situation requires” (Fuller, 1992, p. 401). Which view is the more authentic? Equally important, what should we tell students? What is in their interests? Some years ago, Stephen Brush (1974) posed the question: “should the history of science be rated X?” The question is just as pertinent to the philosophy of science and the sociology of science. Should we expose students to the anarchistic epistemology of Paul Feyerabend? Should we lift the lid off the Pandora’s Box that is the sociology of science? … Can the curriculum achieve a balance that is acceptable to most stakeholders? (p. 919, italics added). (Classified as Level III).

After this presentation, Hodson (2014) presents the following thought provoking questions: (i) Would students be harmed by too early an exposure to these views? (ii) When we seek to question (and possibly reject) the certainties of the traditional

6.2  Results and Discussion

105

view of science, are we left with no firm guidance, no standards and no shared meaning? (iii) Does recognition of the sociocultural baggage of science entail regarding science as just one cultural artifact among many others, with no particular claim on our allegiance? (iv) Is any kind of compromise possible between these extremes and among this diversity? and (v) Can we retain what is still good and useful about the old view of science (such as conceptual clarity and stringent testing) while embracing what is good and useful in the new (such as sensitivity to sociocultural dynamics and awareness of the possibility of error, bias, fraud and the misuse of science)? It is important to note that Derek Hodson is a prominent science educator/researcher with considerable expertise in history and philosophy of science. Despite this some of his views expressed here need a critical appraisal and I will comment on them in the same order as they appear in the quotation: (a) “In building a school science curriculum, are we faced with a stark choice between the traditional and the postmodern?” Most science teachers and even science education researchers would agree that we do not have to make a “stark choice” as most (including textbook authors) would follow some form of traditional empiricist epistemology and very few would consider themselves as postmodern or concur with the scenario presented by Fuller, as suggested by Hodson. (b) “Some years ago, Stephen Brush (1974) posed the question: ‘should the history of science be rated X?’” Actually, this reference to Brush is dated and might even be considered as misleading. Brush (1978) has clarified his position in the following terms: “I believe also that the new style of history of science, which emphasizes the dynamics of scientific change and its relation to the philosophical, technological, and social background, is much more suitable for educational purposes than the older tradition that stressed the accumulation of facts and the assignment of credit for discoveries” (p. 289, italics added). Later in the same article, Brush (1978) added: “Of course, as soon as you start to look at how chemical theories developed and how they were related to experiments, you discover that the conventional wisdom about the empirical nature of chemistry is wrong. The history of chemistry cannot be used to indoctrinate students in Baconian methods ... If we really believe that Lavoisier and Dalton were great chemists, we should be able to live with a more accurate account of how they made their discoveries rather than censor such an account because it does not conform to obsolete doctrines about scientific methodology” (p. 290, italics added). It seems that Brush has not only clarified his position with respect to the history of science (e.g., Lavoisier & Dalton, and many others) but also drawn attention to the need for abandoning the traditional Baconian approach to science based on conventional wisdom that emphasizes its empirical nature. (c) “Should we expose students to the anarchistic epistemology of Paul Feyerabend?” This is the most surprising part of Hodson’s presentation. Before referring to anarchistic epistemology there is no mention of Feyerabend. However, later on page 921, there is a reference to Feyerabend’s (1962) views related to observation statements. At best Hodson’s (2014) position with respect

106

6  Understanding Epistemological Anarchism (Feyerabend) in Research Reported…

to ­anarchistic epistemology remains ambiguous. If he is endorsing this epistemology for the classroom then it would be necessary to provide a framework based on plurality of theories, and some elaboration of Feyerabend’s philosophy of science (methodological pluralism). However, I have included some details of Hodson’s (2009) views in Chap. 1. (d) “Can the curriculum achieve a balance that is acceptable to most stakeholders?” This is the most difficult question. What exactly do we mean by “balance.” If this means some tacit acceptance of “Baconian methods”, as found in most science textbooks and curricula, then there is no room for compromise. Furthermore, this would mean that the whole edifice built by the research community (starting with Harvard Project Physics, the journal Science & Education, even this Handbook) would “crumble.” At this stage it is interesting to refer to an exchange that took place between S. Brush (1979) and H. Siegel (1979) with respect to the distortion of the history of science in science education. Despite their differences (Siegel being more critical of Kuhn with respect to distortion), both agree that the Harvard Project Physics (1975) is a successful example of a historical account for science education (especially the unit on the Copernican Revolution). For a critical appraisal of Kuhn’s views on the distortion of history of science in science education see Niaz (2011a, 2011b, chap. 2, pp. 17–33).

6.2.6  Science as Cultural Tyranny In a section entitled, “Science as cultural tyranny: Feyerabend”, McCarthy (2014) has drawn on Feyerabend’s (1974/1988), How to defend society against science, by reproducing the following: … my criticism of modern science is that it inhibits freedom of thought. If the reason is that it [modern science] has found the truth and now follows it, then I would say that there are better things than first finding, and then following such a monster (Feyerabend, 1974/1988, p. 37, italics added).

Based on this quotation, McCarthy concluded that if “true belief limits freedom” then this leads to a “wholesale rejection of modern science” (p. 1933). At this stage it would be interesting to contrast Feyerabend’s and McCarthy’s views on modern science with those of Giere (2006a): … at the end of the nineteenth century, physicists were as justified as they could possibly be in thinking that classical mechanics was objectively true. That confidence was shattered by the eventual success of relativity theory and quantum mechanics a generation later … For a perspectival realist, the strongest claims a scientist can legitimately make are of a qualified, conditional form: “According to this highly confirmed theory (or reliable instrument), the world seems to be roughly such and such.” There is no way legitimately to take the further objectivist step and declare unconditionally: “This theory (or instrument) provides us with a complete and literally correct picture of the world itself.” (pp. 5–6, italics added).

6.2  Results and Discussion

107

Now, let us compare the statements (reproduced above) from Feyerabend and Giere. What are the differences or similarities? Both are skeptical and question the quest for finding the “truth.” Furthermore, Feyerabend’s “inhibits freedom of thought” is quite similar to Giere’s “confidence was shattered”. What is, however, perhaps objectionable, is the use of the word “monster” for referring to “truth” by Feyerabend—some scholars would consider this as part of his rhetoric. Nobody, would consider Giere to be a postmodernist or relativist as they would consider Feyerabend to be. Interestingly, Giere (2016) considers not only himself but also Feyerabend to be a perspectival realist (see Chap. 2 for details). In order to illustrate the educational consequences of Feyerabend’s philosophy, McCarthy (2014) reproduced the following text: Three cheers to the fundamentalists in California who succeeded in having a dogmatic formulation of the theory of evolution removed from the text books, and an account of Genesis included (Feyerabend, 1974/1988, p. 41).

Indeed, here McCarthy is correct and this is surprising as according to Feyerabend’s own philosophical prescription of plurality of theories, both accounts (evolution and Genesis) could have been retained in the textbooks, and thus even represent counterinduction. Next, McCarthy refers to Feyerabend’s thesis of “the disunity of science” in which he refers to how classical physicists distinguished between the objective world of scientific laws and the subjective world of our experiences, and “They ascribed reality to the former and regarded the latter as an illusion” (Feyerabend, 1996, p.  39). McCarthy considered Feyerabend’s criticism to be well grounded. Interestingly, recent scholarship has gone beyond by recognizing a dualism between objectivity and subjectivity leading to a conflict in the evolving nature of objectivity (Daston & Galison, 2007; Niaz, 2018). Finally, McCarthy (2014) refers to The worst enemy of science? (Preston, Munévar, & Lamb, 2000) which provides insight with respect to Feyerabend’s scholarly intentions from philosophers of science who knew him personally. These scholars provide testimony to the effect that Feyerabend was intentionally provocative and extreme in his assertions about science. This chapter provides examples of research reported in the Handbook (HPST) that provide a number of different perspectives with respect to Feyerabend’s epistemological anarchism. Conclusions based on these findings will be integrated and synthesized with those from other chapters and presented in Chap. 8.

Chapter 7

Feyerabend’s Counterinduction and Science Textbooks

According to Feyerabend (1970a) most scientific theories are not consistent with all known facts (p. 43). Consequently, if we tell students to accept only those theories which are consistent with the available and accepted facts, we shall be left without any theory. To solve this dilemma, Feyerabend suggested a change in methodology by admitting counterinduction, namely accepting unsupported hypotheses. According to Kalman (2019b): “Counterinduction is the process by which one theory or idea is used to effect change in its rival.” Furthermore, all scientists working in a field of knowledge do not necessarily agree with respect to all “observations”, “experimental results” and “theories.” Such views may be considered as “apocalyptic” by some and Feyerabend himself as the worst enemy of science, who still paid the following tribute to science, “Science gives us theories of high beauty and sophistication. Modern science has developed mathematical structures which exceed anything that has existed so far in coherence and generality” (Feyerabend, 1970a, p. 42). With this background the objective of this chapter is to explore the following historical episodes as examples of counterinduction and draw implications for science textbooks and possible teaching strategies: 1. Brownian motion 2. Kinetic theory of gases 3. Michelson-Morley experiment 4. The oil-drop experiment 5. Alpha particle scattering experiment 6. Bohr’s incorporation of “quantum of action” to classical electrodynamics 7. Photoelectric effect 8. Wave-particle duality 9. Mendeleev’s periodic table of chemical elements 10. Lewis’s postulation of the covalent bond

© Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_7

109

110

7  Feyerabend’s Counterinduction and Science Textbooks

1 1. Discovery of the planet Neptune 12. Discovery of elementary particle Neutrino 13. Discovery of the Tau Lepton

7.1  Brownian Motion In a letter written to Lakatos on 20 January 1972, Feyerabend recounts how on a visit to Berlin for a seminar, “von Weizsäcker [became] responsible for my jump into anarchism … My ‘spiritual development’ always depends on theatrical episodes, never—thank Behemoth—on argument. The situation was as follows. I gave my Brownian motion routine, trying to show that it is good, given a highly confirmed theory, to invent an alternative and to work on it … So, if I had not met von W., there would not have been any AM [Against Method]” (Reproduced in Motterlini, 1999, p. 272). Indeed, this incident is recounted by Feyerabend in his particular style with rhetoric and a hyperbolic flourish. Brownian motion refers to the irregular motion of small suspended particles in various fluids which keeps them from sinking due to gravitation, and was discovered by the Scottish botanist Robert Brown in 1827. Attempts to explain the phenomenon have been associated with controversies since the late nineteenth to the early twentieth century. Some of the debates for understanding the cause of Brownian motion were based on different guiding assumptions (cf. Laudan, Laudan & Donovan, 1988), such as the following: kinetic-molecular versus energeticist ontologies, mechanical versus phenomenological explanation, atomic versus continuous metaphysics, statistical versus non-statistical models, hypotheses versus positivistic methodologies, realism versus instrumentalism, among others (cf. Mayo, 1988, p. 220). Of these conflicting appraisals, the ones by Einstein (kinetic-­ molecular), Duhem-Ostwald (energeticist) and thermodynamics Clausius-Kelvin-­ Born (thermodynamics) were more influential. Furthermore, the degree to which these programs had empirical support varied and hence the need for many alternate explanations in accord with Feyerabend’s thesis of counterinduction. Finally, after the work of Perrin (1923) the Einstein-Smoluchowski theory of Brownian motion became pre-eminent. In the late nineteenth and early twentieth century, atomic theory was still being questioned by some leading physicists (e.g., P. Duhem & W. Ostwald) and consequently Brownian motion could not be explained by the kinetic-molecular theory, or in other words support could be accepted provided we considered it as an example of counterinduction. Following is an example of a general chemistry textbook (see Appendix 9 for references to general chemistry textbooks) that established an explicit relationship between Brownian movement and confirmation of kinetic-­ molecular theory: When a colloidal system against a black background in a darkened space is viewed with a microscope at right angles to an intense beam of light, the colloidal particles look like tiny bright flashes of light in irregular rapid, dancing motion—This motion is called Brownian

7.2  Kinetic Theory of Gases

111

movement after the botanist Robert Brown, who first observed it in 1828. He could not explain it, but we now know that the colloidal particles are so small that bombardment by molecules of the dispersion medium makes them move irregularly. This movement explains why dispersed particles in colloidal systems do not settle, even though they are denser than the dispersion medium. The kinetic-molecular theory received one of its earliest confirmations as a result of studies of Brownian movement (Holtzclaw & Robinson, 1988, p. 349, emphasis in the original).

This textbook clearly recognizes that Brown could not explain the “dancing motion” and hence at first the movement of the particles remained a counterinduction. It was only later (early twentieth century) that support for atomic theory and then kinetic theory, explained the Brownian movement. Of course, this presentation could be improved by including a reference to the work of Einstein and Perrin, and hence was classified as “Mention.” At this stage, it is interesting to consider the following presentation that was classified as an example of “counterinduction”: If you look at a ray of sunlight entering a window, you can see little pieces of dust in it moving suddenly as gas molecules hit them. This random movement, which was first observed in 1827 by the Scotch botanist Robert Brown, is called Brownian motion. Early in the twentieth century, theoretical work on Brownian motion by Albert Einstein and quantitative studies of it by French physical chemist Jean-Baptiste Perrin put an end to the last doubts about the existence of atoms and molecules (Umland & Bellama, 1999, p. 176, emphasis in the original).

This presentation explicitly refers to the role played by Einstein’s work on the kinetic-molecular theory and experimental support by the work of Perrin, which led to an explanation of Brownian motion and was classified as an example of counterinduction in the context of Feyerabend’s epistemological anarchism.

7.2  Kinetic Theory of Gases Clausius’ (1857) is considered to be the first full-fledged kinetic theory of gases and his following simplifying (basic) assumptions can be considered as a prelude to those of later work: (1) The space actually filled by the molecules of the gas must be infinitesimal in comparison to the whole space occupied by the gas; (2) Duration of the impact (i.e., change of direction) of the molecules must be infinitesimal compared with the time interval between the collisions; and (3) Influence of the molecular forces between the molecules must be infinitesimal. The starting point of James Clerk Maxwell’s (1860) work on the kinetic theory of gases was his reading of the paper by Clausius (1857), entitled: “On the nature of the motion which we call heat” (cf. Garber et al., 1986, p. xix). Similarly, Maxwell recognized the work of early kinetic theorists, such as Bernouilli, Herapath, Joule, and Krönig. Maxwell (1860) includes the following simplifying (basic) assumptions of his theory (and thus extends those of Clausius): 1 . Gases are composed of minute particles in rapid motion. 2. Particles are perfectly elastic spheres.

112

7  Feyerabend’s Counterinduction and Science Textbooks

3 . Particles act on each other only during impact. 4. Motion of the particles is subject to mechanical principles of Newtonian mechanics. 5. Velocity of the particles increases with the temperature of the gas. 6. Particles move with uniform velocity in straight lines striking against the sides of the container, producing pressure. 7. Derivation of the distribution law assumes that the x-, y-, and z-components of velocity are independent. According to Achinstein (1987), a philosopher of science: “How did Maxwell arrive at them [simplifying assumptions]? They are highly speculative, involving as they do the postulation of unobserved particles exhibiting unobserved motion” (p. 410, italics added). Did Maxwell have an independent warrant (i.e., plausibility of the hypotheses) for his simplifying assumptions? It is plausible to suggest that Maxwell’s simplifying assumptions are precisely the ceteris paribus clauses, which helped him to progress from simple to complex models of the gases. Taking our cue from Feyerabend’s counterinduction and Galilean idealizations, it is plausible to interpret Maxwell’s basic assumptions in the following terms: “The move from the complexity of nature to the specially contrived order of the experiment is a form of idealization. The diversity of causes found in Nature is reduced and made manageable. The influence of impediments, i.e., causal factors which affect the process under study in ways not at present of interest, is eliminated or lessened sufficiently that it may be ignored” (McMullin, 1985, p. 265). This research methodology of idealization, that is, building of simple to complex models, is an important characteristic of modern non-Aristotelian science (for details, see Kitchener, 1993; Kitcher, 1993; Matthews, 1987; Niaz, 1993). Lakatos (1970) has endorsed this position in the following terms: “Moreover, one can easily argue that ceteris paribus clauses are not exceptions, but the rule in science” (p. 102, original italics). Furthermore, Maxwell’s research program is yet another example of a program progressing on inconsistent foundations (similar to Bohr, cf., Lakatos, 1970, p. 142). Among other assumptions, Maxwell’s (1860) paper was based on ‘strict mechanical principles’ derived from Newtonian mechanics and yet at least two of Maxwell’s simplifying assumptions (referring to the movement of particles and the consequent generation of pressure) were in contradiction with Newton’s hypothesis explaining the gas laws based on repulsive forces between particles. Newton provided one of the first explanations of Boyle’s law in his Principia (1687) in the following terms: “If a gas is composed of particles that exert repulsive forces on their neighbors, the magnitude of force being inversely as the distance, then the pressure will be inversely as the volume” (Brush, 1976, p. 13). Apparently, due to Newton’s vast authority, Maxwell even in his 1875 paper, ‘On the dynamical evidence of the molecular constitution of bodies’ reiterated that Newtonian principles were applicable to unobservable parts of bodies (cf. Achinstein, 1987, p. 418). Brush (1976) has pointed out the contradiction explicitly: “… Newton’s laws of mechanics were ultimately the basis of the kinetic theory of gases, though this theory had to compete with the repulsive theory attributed to Newton” (p. 14).

7.2  Kinetic Theory of Gases

113

Based on Lakatos’ (1970) philosophy of science, Clark (1976) considers the 7 simplifying assumptions (mentioned above) as the hard core (negative heuristic) of Maxwell’s research program and summarizes it in the following terms: “… the behaviour and nature of substances is the aggregate of an enormously large number of very small and constantly moving elementary individuals subject to the laws of mechanics” (p. 45, original italics). Similarly, Clark (1976, p. 45) considers the following methodological directives (among others) as the positive heuristic (cf. Lakatos, 1970) of Maxwell’s research program: (a) Try to weaken or if possible eliminate the simplifying assumptions, so as to simulate, as far as possible, conditions obtaining in a ‘real’ gas; and (b) Use the specific assumptions introduced to investigate the internal properties of gases (e.g., viscosity) while the macroscopic (hydrodynamic) and equilibrium properties should be derivable as limiting cases. Indeed, even Feyerabend (1970a, 1970b, 1970c, 1970d) would endorse such a strategy. In subsequent work, Maxwell’s major contribution (and was the first theory to do so) was to make predictions beyond the hydrodynamical laws (Boyle, Charles, Gay-Lussac, etc.), referring to transport properties of gases, by modifying their original simplifying assumptions. Niaz (2000a, 2000b) has evaluated general chemistry textbooks (published in U.S.A., between 1970s to 1990s, see Appendix 9) to examine the role of Maxwell’s simplifying assumptions. Of the 22 textbooks only three made a satisfactory presentation and following is an example: At this point we want to build a model (theory) to explain why a gas behaves as it does … laws do not tell us why nature behaves the way it does. Scientists try to answer this question by constructing theories (building models). The models in chemistry are speculations about how individual atoms or molecules (microscopic particles) cause the behavior of macroscopic systems (collections of atoms and molecules in large enough numbers so that we can observe them). A model is considered successful if it explains known behavior and predicts correctly the results of future experiments. But a model can never be proved absolutely true. In fact, by its very nature any model is an approximation and is doomed to fail at least in part. Models range from the simple (to predict approximate behavior) to the extraordinarily complex (to account precisely for observed behavior) … A relatively simple model that attempts to explain the behavior of an ideal gas is the kinetic molecular theory. This model is based on speculations about the behavior of the individual particles (atoms or molecules) in a gas. The assumptions (postulates) of the kinetic molecular theory can be stated as follows: … (Zumdahl, 1990, pp. 434–435, original italics).

Zumdahl’s (1990) presentation of Maxwell’s simplifying assumptions emphasizes their following aspects: speculative, models, approximate, and that models develop (tentativeness) in order to explain the behavior of gases. Of course, it can be argued that such presentations are not based on an overt understanding of history and philosophy of science. However, it is particularly important to note the following assertion, “a model can never be proved absolutely true.” Interestingly, not only Feyerabend but also recent philosophy of science would endorse such a thesis (cf. Giere, 2006a, 2006b). Nevertheless, most teachers would agree, that it constitutes a convincing and helpful preamble before presenting the postulates of the kinetic theory. Such an approach could facilitate a better understanding of the importance of the kinetic theory, as a source for conceptual understanding.

114

7  Feyerabend’s Counterinduction and Science Textbooks

7.3  Michelson-Morley Experiment The Michelson-Morley experiment was first conducted in 1887 and provided a “null” result with respect to the ether-drift hypothesis, namely, that there was no observable velocity of the earth with respect to the ether (Michelson & Morley, 1887). According to Kalman (2019b), “Michelson did not regard his experiment as a ‘null’ result with respect to the ether-drift hypothesis. He was examining two rival ether-drift theories [G.  Stokes and A.  Fresnel]” (original italics). H.A.  Lorentz (1895), a leading European physicist, presented an auxiliary hypothesis to explain the null findings of the Michelson-Morley experiment. According to the ether theory the round-trip velocity of light should be lower in the direction of the earth’s motion through the ether than in the direction perpendicular to this motion. In contrast, results of the Michelson-Morley experiment had shown that the round-trip velocity of light is the same in all directions on the earth’s surface. The Lorentz contraction (auxiliary) hypothesis suggested that all bodies on the earth undergo a minute contraction in the direction of the earth’s motion through the surrounding ether. Even Popper (1959) considered that Lorentz had successfully explained the null results of the experiment. Dayton Clarence Miller (a colleague of Michelson at Case) continued to work to provide experimental evidence for the ether-drift hypothesis. It was argued that the original Michelson-Morley experiment was conducted in the basement of a laboratory and hence the result. Miller instead decided to perform the experiment on a hilltop (Mount Wilson Observatory), between April 8 and 21, 1921. Einstein had arrived at Princeton to give lectures on the relativity theory. While still at Princeton, Einstein was informed of Miller’s result (a nonzero ether drift). It is precisely on hearing these experimental findings, that Einstein stated, “Subtle is the Lord, but malicious He is not” (Pais, 1982, p. 113). Interestingly, although Einstein was skeptical of Miller’s results, still on May 25, 1921 he went to see Miller in Cleveland. Even the news correspondent of a journal like Science reported to his readers that, “Professor Miller’s results knock out the relativity theory radically” (reproduced in Lakatos, 1970, p. 165). For his part, Michelson (1927) continued to have doubts with respect to relativity, especially how the propagation of light waves without a medium can be explained (for details see Shankland, 1963, 1964). On April 28, 1925, Miller read a paper before the National Academy of Sciences, Washington, D.C., and reported that an ether drift had definitely been established (Miller, 1925). Miller continued to publish on the subject (Miller, 1933) and was one of the organizers of the international conference on the Michelson-Morley experiment held in 1927 at the Mount Wilson Observatory, in which among others Michelson and H.A. Lorentz participated. According to Lakatos (1970) despite considerable experimental evidence (both for and against), starting in 1905, it took almost 25 years for this hypothesis to be refuted and recognized as the “greatest negative experiment in the history of science” (p.  162). Interestingly, Feyerabend (1970a) presents a somewhat different interpretation of the events: “The special theory of relativity was retained, despite

7.3  Michelson-Morley Experiment

115

D.C. Miller’s decisive refutation. (I call this refutation “decisive” because the experiment was, from the point of view of contemporary evidence, at least as well performed as the earlier experiment of Michelson and Morley” (p. 37). As an end note Feyerabend provides further information with respect to the fact that H.A. Lorentz studied Miller’s work for many years and could not find any problems. Finally, it was in 1955, 25 years after Miller had finished his experiments that a satisfactory account of the results was found (for details see Shankland, 1963). At this stage it is interesting to compare the interpretations of Feyerabend and Lakatos: the former highlights counterinduction, namely, Einstein’s special theory of relativity was retained despite empirical evidence (Miller’s) to the contrary, and the latter emphasized that it took 25 years for the ether-drift hypothesis to be refuted. Many leading scientists and science textbooks have generally attributed the origin of Einstein’s special theory of relativity in 1905 to the Michelson-Morley experiment. Despite some ambivalence, Einstein provided his interpretation in the following terms: In my own development Michelson’s result has not had a considerable influence. I even do not remember if I knew of it at all when I wrote my first paper on the subject (1905). The explanation is that I was, for general reasons, firmly convinced that there does not exist absolute motion and my problem was only how this could be reconciled with our knowledge of electro-dynamics. One can therefore understand why in my personal struggle Michelson’s experiment played no role or at least no decisive role. (Einstein’s letter to Davenport, 9 February 1954, reproduced in Holton, 1969a, b, p. 969).

Despite this statement from Einstein, a recent appraisal has shown that Einstein, “appeared undecided as to what the true story was, at times denying that he knew about it [Michelson-Morley experiment] and at times giving it a decisive role” (Arabatzis & Gavroglu, 2015, p. 152). Consequently, it is understandable, if textbooks continue to emphasize the “myth” that Einstein was led to his special theory of relativity due to the null result of the Michelson-Morley experiment. Textbooks published in different countries even today generally emphasize that it was the Michelson-Morley (MM) experiment that led Einstein to postulate his special theory of relativity (STR). Brush (2000) has analyzed 26 physics textbooks (published in U.S.A., see Appendix 10) with respect to the relationship between the experiment and Einstein’s STR. Only nine textbooks still attributed the STR to the negative result of the MM experiment. Interestingly, however, these nine textbooks included some of the most well-known and widely used textbooks, (not only in the U.S.A., but almost all over the world), such as Serway (1996) and Sears et  al. (1991). Similarly, Aarriassecq and Greca (2007) have reported that high school physics textbooks published in Argentina also suggest that the starting point for the STR was the null result of the MM experiment. With this background it would be interesting to reconsider the two interpretations presented above: (a) Lakatos, it took 25 years for the ether-drift hypothesis to be refuted; and (b) Feyerabend, Einstein’s special theory of relativity was retained despite empirical evidence (Michelson, Miller) and including Lorentz contraction, to the contrary. Although, both interpretations have common elements (empirical evidence) there are some important differences. Lakatos implies that empirical

116

7  Feyerabend’s Counterinduction and Science Textbooks

e­ vidence was necessary for refuting the ether-drift hypothesis, whereas Feyerabend would imply that despite empirical evidence to the contrary, STR was not refuted. Does this have educational implications (especially for writing textbooks)? In other words, a review of the literature at present shows that researchers were interested in finding: Michelson-Morley (MM) experiment led Einstein to postulate his special theory of relativity (STR). With Feyerabend’s perspective of counterinduction it would be interesting to find if science textbooks explore: Despite empirical evidence to the contrary, STR was not refuted.

7.4  The Oil-Drop Experiment The determination of the elementary electrical charge (e) aroused considerable interest in the scientific community and both Robert Millikan (1868–1953), University of Chicago and Felix Ehrenhat (1879–1952), University of Vienna, became deeply involved in its measurement. Both scientists used very similar experimental method, except that Millikan used oil drops, whereas Ehrenhaft used metal drops. The experiment is generally considered to be simple, beautiful and straightforward, that unambiguously led to the determination of the elementary electrical charge. According to Crease (2002), in a poll conducted for Physics World, its readers considered the oil drop experiment to be one of the ten ‘most beautiful’ ones of all times. Both scientists obtained experimental data that were quite similar. Yet, Millikan postulated the existence of a universal charged particle (the electron) whereas Ehrenhaft postulated the existence of sub-electrons based on fractional charges. The controversy started when Ehrenhaft recalculated Millikan’s data of oil drops and found a large spread of values of the electrical charge, quite similar to his own data. Ehrenhaft showed how Millikan’s method led to paradoxical situations. For example, two oil drops having very similar charges were considered by Millikan to have different numbers of electrons. The scientific community wondered: How to explain these differences? The controversy between the two was intense and lasted for many years (from around 1910–1923, when Millikan was awarded the Nobel Prize). At one stage Millikan (1916) asked the reader to suppose that Ehrenhaft’s data were free of all the possible errors that he had discussed previously, and posed the following dilemma: That these same ions have one sort of charge when captured by a big drop and another sort when captured by a little drop is obviously absurd. If they are not the same ions which are caught in the two cases, then, in order to reconcile the results with the existence of the exact multiple relationship …, it would be necessary to assume that there exist in the air an infinite number of different kinds of ionic charges corresponding to the infinite number of possible radii of drops, and that, when a powerful electric field drives all of these ions toward a given drop, this drop selects in each instance just the charge which corresponds to its particular radius. Such an assumption is not only too grotesque for serious consideration but is directly contradicted by my experiments … (p. 617, original italics).

7.4  The Oil-Drop Experiment

117

This passage is indeed revealing, and all the more so as it was published in a leading physical science journal. In summary, Millikan is telling the reader that experimental observations are important, but there is something even more important, viz. the guiding assumptions, and any data that go against them would appear to be “absurd” and “grotesque,” and hence subelectrons could not exist. This story, however, does not end here. Almost 55 years later, in 1978, a new dimension to the controversy was added by Holton’s (1978a) discovery of Millikan’s two laboratory notebooks in his Archives at the California Institute of Technology, Pasadena. The Millikan notebooks have data (175 pages with data from experiments conducted between October 28, 1911 and April 16, 1912) that were published in his Physical Review article (Millikan, 1913). In these notebooks Holton found data from 140 drops, but the published article reported data from 58 drops. What happened to the other 82 drops? It seems that Millikan made a rough calculation for the value of e (elementary electrical charge), as soon as the data for the times of descent/ascent of the oil drops started coming in. Apparently, Millikan ignored any experiment that did not give the value of e that he expected and he went on to do another experiment. This leads to the question: What was the warrant under which Millikan discarded more than half of his observations? Millikan’s guiding assumption, based on the atomic nature of electricity (dating back to Benjamin Franklin) and the value suggested by the previous experiments of Ernest Rutherford (1871–1937) at the University of Manchester, was a constant source of guidance. Indeed, Millikan would have liked to warn Ehrenhaft that all the data could not be used as their experiments were constantly faced with difficulties such as evaporation, sphericity, radius, and change in density of drops and variation in experimental conditions (battery voltages, stopwatch errors, temperature, pressure and convection). It is important to note that, similar to Millikan, Ehrenhaft also obtained data that he interpreted as integral multiple of the elementary electrical charge (e). Nevertheless, Ehrenhaft’s argument was precisely that there were many drops that did not lead to an integral multiple of e. According to Holton (1978a), Ehrenhaft used data from all the drops that he studied and this led to the impasse: It appeared that the same observational record could be used to demonstrate the plausibility of two diametrically opposite theories, held with great conviction by two well-equipped proponents and their collaborators (pp. 199–200, italics added).

With this historical reconstruction it is plausible to suggest that: (a) Ehrenhaft allowed his theory (subelectrons) to be dictated by experimental data and hence in a sense followed the scientific method (accepted all the data); (b) Millikan discarded data that did not support his guiding assumption—this coincides with Feyerabend’s (1975a) claim that scientific theories are not consistent with all the experimental data (p. 43); and (c) Millikan supported a theory that was not supported by at least 59% of his data, and hence accepted a theory that was at least partially unsupported, and Feyerabend would consider this as counterinduction. At this stage it would be interesting to consider how science textbooks present the oil-drop experiment.

118

7  Feyerabend’s Counterinduction and Science Textbooks

One general chemistry textbook presented Millikan’s contribution in the following terms: According to his report of his work (Millikan, R.A. Phys. Rev. 1913, 2(2), 109–143), Millikan included all observations except those he knew to be wrong because some mistake was made in carrying out the experiment. He also said that he did not do any calculations until all of the observations had been made so that his observations could not possibly be influenced by his idea of what the result should be (Umland & Bellama, 1999, p.  227, underline added).

This is an interesting presentation that raises various issues. First, the inclusion of an original reference (Millikan, 1913) is very helpful and sound practice for understanding the context in which the experiment was conducted. Second, both Millikan and Ehrenhaft had prior guiding assumptions and hence they were “influenced by their ideas of what the result should be” (cf. Holton, 1978a, 1978b). Third, there is enough historical evidence to show that Millikan discarded at least half (or more) of his observations. Fourth, there is evidence to show that Millikan did on the spot calculations and took decisions as the data started pouring in, “It appears likely that after almost every run some rough calculations of e [elementary electric charge] were privately made on the spot, and often a summary judgment appended” (Holton, 1978b, p. 70). In other words, Millikan was fully aware of what the result of his experiments should be and took steps to achieve that end. More recently, Holton (2014a, b) has clarified his position further on this issue: “So even if Millikan had included all drops and yet come out with the same result, the error bar of Millikan’s final result would not have been remarkably small, but large—the very thing Millikan did not like” (Italics in the original). A physics textbook presented the oil-drop experiment and referred to Millikan’s guiding assumption in the following terms: By observing the motion of the hundreds of droplets with different charges on them, Millikan uncovered the pattern he expected: the charges were multiples of the smallest charge he measured (Olenick, et al., 1985, p. 241, italics added).

“The pattern he expected” can be considered as a reference to his guiding assumption. The same textbook reproduced the following quote from Millikan’s laboratory notebook (dated 15 March, 1912; see Holton, 1978a, 1978b for Millikan’s lab. notebooks):“One of the best ever [data] ... almost exactly right. Beauty—publish” (p.244, original italics). After reproducing the quote, the textbook authors asked a very thought provoking question: “What’s going on here? How can it be right if he’s supposed to be measuring something he doesn’t know? One might expect him to publish everything!” (Olenick, et al., 1985, p. 244). Millikan’s hand-written notebooks were first studied by Holton (1978a, 1978b), almost 55 years after the controversy with Ehrenhaft, and I am, not aware if Feyerabend knew of them. Nevertheless, the question, “How can it be right if he’s supposed to be measuring something he doesn’t know?” can easily be considered as an example of “counterinduction,” namely, accepting or postulating unsupported hypotheses. According to Kalman (2019b): “In my opinion this is not counterinduction. Counterinduction involves the consideration of two incommensurable theories. Kuhn used the term

7.4  The Oil-Drop Experiment

119

‘­ incommensurable’ to characterize the holistic nature of the changes that take place in a scientific revolution. Feyerabend’s incommensurability corresponds to questions that have meaning only in a particular theory. Thus, I am using incommensurability in the sense of Feyerabend. In analysing Galileo in Against Method, Feyerabend’s point is that if you only consider one theory, you will not come to another theory even if your theory is not entirely supported by experimental evidence. You use counter induction to propose another theory incommensurable with the accepted theory that is also not entirely supported by experimental evidence. By contrasting the two theories you make progress.” Actually, I agree entirely with Kalman that in order to have counterinduction a scientist must have two incommensurable theories. In the oil drop experiment, both Millikan and Ehrenhaft had two theories: (a) Existence of a universal charged particle (electron); and (b) Existence of fractional charges (subelectrons). Although, Ehrenhaft had data for both theories, he supported the second theory. Millikan’s case is more complex. Having data for both theories he denied having data for the second theory (cf. Holton, 1978a for Millikan’s handling of the data). Thus, Millikan had data that only provided partial support to both theories—counterinduction. To continue their treatment, Olenick et  al., 1985 went to considerable length (about 5 pages) to present Millikan’s research methodology and point out the dilemmas and contradictions in the handling of the data: “Now, you shouldn’t conclude that Robert Millikan was a bad scientist ... What we see instead is something about how real science [cutting-edge] is done in the real world. What Millikan was doing was not cheating. He was applying scientific judgement ... But experiments must be done in that way. Without that kind of judgement, the journals would be full of mistakes, and we’d never get anywhere. So, then, what protects us from being misled by somebody whose ‘judgement’ leads to wrong results? Mainly, it’s the fact that someone else with a different prejudice can make another measurement ... Dispassionate, unbiased observation is supposed to be the hallmark of the scientific method. Don’t believe everything you read. Science is a difficult and subtle business, and there is no method that assures success” (Olenick et al., 1985, p. 244). Despite, a very good presentation, this textbook does not mention Holton (1978a, 1978b) and thus ignores research in the history of physics, as Holton’s account of the Millikan–Ehrenhaft controversy had first appeared in 1978. Interestingly, after having read a preliminary version of our article (Rodríguez & Niaz, 2004) Holton (2000) recommended: “… introduction of the history and methodology of physics into the physics classroom, not least in terms of important controversies – is completely congenial to me… I agree one should teach research methodology in introductory physics and Millikan’s case is certainly a well documented case that would lend itself to this purpose”. Following the advice of Holton (2000), Niaz and Rivas (2016) designed a teaching strategy to facilitate high school students’ understanding of the oil drop experiment. Atomic structure forms an important part of high school chemistry courses in almost all parts of the world. Among other aspects, this topic deals with the atomic models of J.J. Thomson (based on cathode ray experiments), E. Rutherford (based on alpha particle experiments), N.  Bohr (based on quantum theory) and the

120

7  Feyerabend’s Counterinduction and Science Textbooks

e­ lementary electrical charge (based on Millikan’s oil drop experiment). The objective of this study (Niaz & Rivas, 2016) was to facilitate Venezuelan high school students’ understanding of research methodology based on alternative interpretations of data, reasoning strategies, role of controversies based on arguments used by scientists, creativity and the scientific method, in the context of the oil drop experiment. These aspects form an important part of the nature of science (NOS). This study is based on a reflective, explicit and activity-based approach to teaching nature of science (NOS). In this respect, the oil drop experiment has been of particular interest to science educators for facilitating students’ understanding of research methodology and the dynamics of scientific progress, i.e., NOS (cf. Holton, 1978a, 1978b; Niaz, 2005, 2015). The study reported here is based on three groups of high school students (tenth grade, 15–18 year olds) enrolled at a public school in Venezuela. One group (n = 33) was randomly designated as the Control and the other two as Experimental group A (n = 33) and Experimental group B (n = 38), respectively. All three groups were taught by the same instructor and participated in the following activities: First week: Instruction in the traditional expository manner on the following aspects of atomic structure: Thomson, Rutherford, Bohr models of the atom and the Millikan oil drop experiment for determining the elementary electrical charge. At the end of the week students were asked to draw a concept map (Novak, 1990) based on how they perceived the development of scientific knowledge. Students were familiar with drawing concept maps from a previous course. Second week: All three groups responded to a 3-item Pretest. Experimental groups A and B were provided a Study Guide based on the scientific method and the Millikan-Ehrenhaft controversy with respect to the determination of the elementary electrical charge, based on the oil drop experiment. Even today chemistry textbooks continue to perpetuate the myth that this was a simple, beautiful and straightforward experiment (cf. Niaz, 2015). Students were asked to read the Study Guide over the weekend and prepare for discussing it the next week. Third week: Experimental group students (A and B) were sub-divided into small groups and asked to present and discuss what they considered to be the principal ideas in the Study Guide. The instructor acted as a moderator and clarified issues, as the students argued with respect to various aspects of the experiment. Study Guide generated considerable discussion, both in and outside the classroom. After this interactive session students were asked to draw another concept map based on what they considered to be the most important aspects of scientific development. Fourth week: Both control and experimental group (A and B) students responded to a 5-item Posttest. During the next month, 17 students from the experimental groups (A and B) and 11 from the control group were selected randomly for a semi-structured interview. Results obtained show that the difference in the performance (conceptual responses) of the control and experimental group (A & B) students on the three items of the Pretest is statistically not significant. However, on the five items of the Posttest experimental groups performed better than the control group and the difference in the performance on conceptual responses is statistically significant. After the experimental treatment most students changed their perspective based on arguments and drew concept maps in which they emphasized the

7.4  The Oil-Drop Experiment

121

creative, accumulative, controversial nature of science and the scientific method. Interviews with students provided a good opportunity to observe how students’ reasoning changed after the experimental treatment. Multiple data sources were an important feature of this study. It is concluded that a teaching strategy based on a reflective, explicit and activity-based approach in the context of the oil drop experiment can facilitate high school students’ understanding of how scientists elaborate theoretical frameworks, design experiments, report data that leads to controversies and arguments and finally with the collaboration of the scientific community a consensus is reached. Most students in this study (Niaz & Rivas, 2016) were quite enthusiastic about the exploration of concept maps and dedicated considerable time and effort in expressing what they considered to be the underlying issues (especially the oil drop experiment). Concept map drawn by one of the students after the experimental treatment is a good example of the student’s interest, engagement with the topic, and creativity. This concept map has the structure of a running dialogue between a student (person on the left) and a teacher (person on the right). Actually, it could also be a dialogue between two students or between a teacher and a student (the concept map drew pictures of what resembled a teacher and a student). In order to understand better we present the concept map (for details see Fig. 4, reproduced in Niaz & Rivas, 2016, p. 34) as three segments involving questions (person on the left) and answers (person on the right): First segment: Question: Who helps in the development of science? Answer: The scientists Question: How do the scientists develop science? Answer: Through investigations, experiments, many studies and then publication. Second segment: Question: How is the relationship between the scientists? Answer: Through controversies, differences, conflicts. Question: How do the scientists analyze the data? Answer: Through their own knowledge and creativity. Third segment: Question: Is science a reality forever? Answer: No …! Question: Do the scientists manage the reality completely? Answer: No, science is tentative. Question: Are they (scientists) creative? Answer: Yes, creativity is very important. These seven questions sub-divided into three segments, in a sense represents the Socratic approach to learning and knowledge. This immediately raises the question, how could a high school student with almost no prior experience in educational research or nature of science could elaborate such a dialogue. Interestingly, the

122

7  Feyerabend’s Counterinduction and Science Textbooks

c­ oncept map drawn by this student before the experimental treatment (Figure  3, p. 33, Niaz & Rivas, 2016) provides no indication of the understanding manifested in Figure 4. Actually, the structures of the two concept maps (Figures 3 and 4) are entirely different. It is plausible to suggest that the classroom treatment provided this student a new perspective that was previously not available. The three questions (Figure 4) in segment three are thought-provoking indeed and represent some ideas of research methodology (also nature of science) in a very lucid manner. The question: “Is science a reality forever?” is impregnated with an answer and this did not require an extensive response, but instead a laconic “No …!” sufficed. Without being overly optimistic, it is plausible to suggest that the experimental treatment played a crucial role in facilitating this students’ understanding of the underlying issues. Concept map drawn by this student in Figure 4, could be considered as representing the Socratic approach to learning and education. Nola (1997) a philosopher of science, in the context of recent trends in constructivism has referred to the Socratic approach in the following terms: In Socrates’ view, students do not acquire knowledge through picking up bits of (true) information didactically conveyed to them. Even being led through a question-answer session does not provide, by itself, knowledge; at best the process can only lead pupils to the correct belief. Only when they can go through the steps of reasoning by themselves and thereby make fully explicit to themselves the reasons for the correct answer will they have knowledge (p. 59).

We wonder, how would Nola characterize this concept map (Figure 4, Niaz & Rivas, 2016, p. 34)? Interestingly, this student is perhaps playing the role of Socrates and the student at the same time. All the questions and answers in the three segments were this student’s own creation. At one stage of the concept map (third segment), the student asked a very thought provoking question: “Is science a reality forever?” It is important to note that this particular question was not discussed in class. Indeed, this is one of the most difficult ideas related to nature of science and the dynamics of scientific progress, and represents the very ethos of Feyerabend’s epistemological anarchism, namely science changes reality forever and hence the need for counterinduction. Actually, Feyerabend (2011) goes even beyond by suggesting that the current view of science may soon be voted out of office (p. 125). It is interesting to consider how this student arrived at this question? Was it the experimental treatment, the classroom discussions or his/her own curiosity and the ability to go beyond what was discussed in class? Niaz et al. (2003) have argued that progress in science and educational theory (constructivism) is characterized by continual critical appraisals. Physicist-philosopher of science, Gerald Holton (1986) has expressed this idea in cogent terms: … the scientists chief duty … [is] … not the production of a flawlessly carved block, one more in the construction of the final Temple of science. Rather, it is more like participating in a building project that has no central planning authority, where no proposal is guaranteed to last very long before being modified or overtaken, and where one’s best contribution may be one that furnishes a plausible base and useful material for the next stage of development (p. 173, italics added).

7.5  Alpha Particle Scattering Experiment

123

Indeed, these ideas characterize the research methodology used by scientists in the context of the dynamics of scientific progress. This, of course leads to the crucial question, how often do we convey this message and facilitate such reasoning in our educational practice. On the contrary, most science educators and textbook authors prepare students as if they were going to enter the “final Temple of science.”

7.5  Alpha Particle Scattering Experiment In the very first paragraph of his famous article in the Philosophical Magazine Rutherford (1911) starts on a controversial note: “It has generally been supposed that the scattering of a pencil of alpha or beta rays in passing through a thin plate of matter is the result of a multitude of small scatterings by the atoms of matter traversed” (p.  669). This of course referred to the experimental work of Crowther (1910), a colleague of Thomson. Rutherford (1911) explicitly points out that Crowther’s experimental results provided support for Thomson’s hypothesis of compound scattering (which led to the “plum pudding” model): The theory of Sir J.J. Thomson is based on the assumption that the scattering due to a single atomic encounter is small, and the particular structure assumed for the atom does not admit of a very large deflexion of an alpha particle in traversing a single atom, unless it be supposed that the diameter of the sphere of positive electricity is minute compared with the diameter of the influence of the atom (p. 670).

This served as a preamble for Rutherford (1911) to present his side of the story in the following terms: The observations, however, of Geiger and Marsden (1909) on the scattering of alpha rays indicate that some of the alpha particles must suffer a deflexion of more than a right angle at a single encounter. They found, for example, that a small fraction of the incident alpha particles, about 1 in 20,000 were turned through an average angle of 90o in passing through a layer of gold-foil … A simple calculation based on the theory of probability shows that the chance of an alpha particle being deflected through 90o is vanishingly small … A simple calculation shows that the atom must be a seat of an intense electric field in order to produce such a large deflexion at a single encounter (p. 669, italics added).

It is interesting to note that Rutherford had the experimental data as early as June 1909 (Geiger and Marsden, 1909), to postulate his model of the nuclear atom, and yet he did not do so until March 1911. What happened between June 1909 and March 1911 is important not only for historians and philosophers of science, but also for science teachers. It is generally ignored that soon after Geiger and Marsden (1909) published their results, Thomson and colleagues started working on the scattering of alpha particles in their own laboratory (Cavendish). Although, results from both laboratories were similar, interpretations of Thomson and Rutherford were entirely different. Thomson propounded the hypothesis of compound scattering, according to which a large angle deflection of an alpha particle resulted from successive collisions between the alpha particle and the positive charges distributed throughout the atom. Rutherford, in contrast, propounded the hypothesis of single

124

7  Feyerabend’s Counterinduction and Science Textbooks

scattering, according to which a large angle deflection resulted from a single collision between the alpha particle and the massive positive charge in the nucleus. The rivalry between Rutherford’s hypothesis of single scattering based on a single encounter and Thomson’s hypothesis of compound scattering, led to a bitter dispute between the proponents of the two hypotheses. At one stage, Rutherford even charged Crowther (1910), a colleague of Thomson, to have “fudged” the data in order to provide support for Thomson’s model of the atom (Wilson, 1983). Heilbron (1981b) has provided a similar account of the rivalry between Thomson and Rutherford. In a letter written to Schuster (Secretary of the Royal Society), on February 2, 1914, Rutherford expressed his opinion forcefully: … I have promulgated views on which J.J. [Thomson] is, or pretends to be, skeptical. At the same time I think that if he had not put forward a theoretical atom himself, he would have come round long ago, for the evidence is very strongly against him. If he has a proper scientific spirit I do not see why he should hold aloof and the mere fact that he was in opposition would liven up the meeting (Reproduced in Wilson, 1983, p. 338).

It is important to note that J.J. Thomson at that time was considered a world master in the design of atomic models (Heilbron & Kuhn, 1969). Given, Thomson’s credentials most scientists could argue that Rutherford’s hypothesis of single scattering, perhaps constituted Feyerabend’s counterinduction. Historical events, however, turned out to be otherwise and the scientific community eventually accepted Rutherford’s arguments and supported his hypothesis of single scattering. One general physics textbook (see Appendix 10) presented Rutherford’s arguments in cogent terms to show that Thomson’s hypothesis was difficult to sustain: The Thomson model for scattering fails when we examine the probability for scattering at large angles. If each individual scattering deflects the projectile through an angle of around 0.01o, then to observe projectiles scattered through a total angle greater than 90o, we must have about 104 successive scatterings, all of which push the projectile toward larger angles. Since the probabilities of individual scatterings toward either larger or smaller angles are equal, the probability of having 104 successive scatterings toward larger angles, like the probability of finding104 successive heads in tossing a coin, is about (1/2)10.000 = 10−3000 (Krane, 1996, p. 178).

Another general physics textbook went even beyond and provided the following arguments: Rutherford calculated that from the large Thomson positive charge distribution particles should never be deflected more than 0.03 degrees in a single collision; in undergoing multiple collisions they should have about an equal chance of being deflected one way as another. Therefore, large deflections as a result of many single deflections in the same direction were very improbable. (It had been calculated on the basis of the Thomson model that a total deflection greater than 90o in traversing the gold foil would have only one chance in 103500 of occurring.) (Cooper, 1970, p. 321).

It is important to note that Cooper is a theoretical physicist and Nobel Laureate with considerable interest in the history of science. At this stage, it would be interesting to ask, as to how did Krane and Cooper come to have arguments based on the probability of the deflections. As is generally the case, history of science is a rich source of information provided we look for the evidence. In the present case, Rutherford

7.6  Bohr’s Incorporation of “quantum of action” to Classical Electrodynamics

125

(1911) himself in his seminal article provided the arguments based on probability (p. 669). The next question that comes to mind is: What role did counterinduction play in this historical episode? It is plausible to suggest that Thomson’s hypothesis of compound scattering provided counterinduction and thus helped Rutherford to strengthen his arguments.

7.6  B  ohr’s Incorporation of “quantum of action” to Classical Electrodynamics In the fourth paragraph of his famous trilogy, Bohr (1913) acknowledged the inadequacy of Maxwell’s classical electrodynamics in describing the behavior of systems of atomic size and thus suggested to include, “Planck’s constant, or as it often is called the elementary quantum of action” (p. 2). Bohr (1913), in general, had a fairly adverse reception in the scientific community. Otto Stern told a friend: “If that nonsense is correct which Bohr has just published, then I will give up being a physicist” (reproduced in Holton, 1986, p. 145). Lord Rayleigh was categorical: “It does not suit me” (reproduced in Holton, 1993, p. 79). H. A. Lorentz objected: “… the individual existence of quanta in the aether is impossible …” (reproduced in Holton, 1993, p.  79). J.  J. Thomson, whom Bohr considered as the “world master in the design of atomic models” objected to Bohr’s conception in most of his writings from 1913 to 1936 (cf. Holton, 1993, p. 79). Rutherford, himself although supportive of Bohr’s model had similar misgivings, “… the mixture of Planck’s ideas with the old mechanics makes it very difficult to form a physical idea of what is the basis of it all … How does the electron decide what frequency it is going to vibrate at when it passes from one stationary state to another? (Letter written by Rutherford to Bohr, dated 20 March 1913, reproduced in Holton, 1993, p.  79). More recently, however, philosophers of science have been more understanding of Bohr’s model of the atom and following are some examples: … it is understandable that, in the excitement over its success, men overlooked a malformation in the theory’s architecture; for Bohr’s atom sat like a baroque tower upon the Gothic base of classical electrodynamics. (Margenau, 1950, p. 311). Thus Bohr’s atom of 1913 was really a kind of mermaid—the improbable grafting together of disparate parts, rather than a new creation incorporating quantum theory at its core. (Holton, 1986, p. 145).

What exactly was Bohr doing? The ideas of “baroque tower, Gothic base and mermaid” are indeed quite picturesque. In contrast, Lakatos (1970) has argued that Bohr used a methodology employed frequently by scientists in the past and perfectly valid for the advancement of science: … some of the most important research programmes in the history of science were grafted on to older programmes with which they were blatantly inconsistent. For instance, Copernican astronomy was “grafted” on to Aristotelian physics, Bohr’s programme on to Maxwell’s. Such “grafts” are irrational for the justificationist and for the naive falsificationist, neither of whom can countenance growth on inconsistent ­foundations … As the

126

7  Feyerabend’s Counterinduction and Science Textbooks

young grafted programme strengthens, the peaceful co-­existence comes to an end, the symbiosis becomes competitive and the champions of the new programme try to replace the old programme altogether. (Lakatos, 1970, p.  142, italics in original, underline added).

Growth on inconsistent foundations as suggested by Lakatos can easily be construed as counterinduction as postulated by Feyerabend, namely postulating hypotheses that are not entirely supported by experimental evidence. In contrast, Popper (1963a) a falsificationist had recommended that Bohr’s (1913) article should not have been published then as it was based on inconsistent foundations. Most science textbooks include the following postulates as part of Bohr’s model of the atom: (a) The electron in an atom has only certain definite stationary states of motion allowed to it and each of these stationary states has a definite, fixed energy; (b) When an atom is in one of these states, it does not radiate energy. However, energy is emitted or absorbed depending on the difference between the energies of two states; (c) In the different energy states, the electron moves in a circular orbit around the nucleus; and (d) Only those energy states are allowed to the electrons in which the angular momentum is an integral multiple of h/2π. According to the modern atomic theory these postulates need corrections and further elaboration. Crux of the issue is the following: Did Bohr have experimental evidence for his postulates? Of course, some experimental evidence was available, but for the most part Bohr’s model was based on inconsistent foundations and hence the use of counterinduction. Most critics of Bohr’s model (such as Lorentz, Rayleigh, Rutherford, & Stern) would implicitly or explicitly reject counterinduction. The one who came closest to Feyerabend’s hypothesis was Lakatos. At this stage it is interesting to consider the evaluation of Bohr’s oeuvre by a modern scholar: The first assumption [Bohr’s postulate] is the existence of stationary states, the second is the frequency rule. Bohr regarded them as the unshakable pillars of his theory. They were indeed more directly related to experiments than other assumptions of his theory. Until at least 1925, they remained the two basic postulates of the quantum theory, despite the vicissitudes of most other assumptions (Darrigol, 2009, p. 154).

This shows that lack of experimental evidence has continued to worry even recent scholars. Most science textbooks generally also follow a similar line of criticism of Bohr’s model. Following is an example of a general chemistry textbook (see Appendix 9) that tries to understand Bohr’s predicament and approximates to Feyerabend’s perspective: There are two ways of proposing a new theory in science, and Bohr’s work illustrates the less obvious one. One way is to amass such an amount of data that the new theory becomes obvious and self-evident to any observer. The theory then is almost a summary of the data. The other way is to make a bold new assertion that initially does not seem to follow from the data, and then to demonstrate that the consequences of this assertion, when worked out, explain many observations. With this method, a theorist says, “You may not see why, yet,

7.6  Bohr’s Incorporation of “quantum of action” to Classical Electrodynamics

127

but please suspend judgment on my hypothesis until I show you what I can do with it.” Bohr’s theory is of this type. Bohr said to classical physicists: “You have been misled by your physics to expect that the electron would radiate energy and spiral into the nucleus. Let us assume that it does not, and see if we can account for more observations than by assuming that it does (Dickerson, Gray, Darensbourg, & Darensbourg, 1984, p.  264, italics added).

This is a very interesting presentation and shows that sometimes even textbook authors can facilitate understanding of historical events in science that are elusive even for historians. After reading a preliminary version of this chapter, Kalman (2019b) considered the part in italics as, “This is certainly the thrust of counterinduction.” Furthermore, this illustrates how textbooks can enrich their presentations. Now, let us consider the following sequence of arguments, as presented by the textbook: two ways of proposing a new theory in science ➔ a bold new assertion that initially does not seem to follow from the data ➔ account for more observations. It is plausible to suggest that this sequence comes quite close to Feyerabend’s counterinduction, precisely as Bohr accepted postulates for which he had no experimental evidence. In the present case Bohr followed a different way of understanding Rutherford’s nuclear atom. Introduction of Planck’s quantum of action would constitute a bold new hypothesis and finally more observations can be accounted for, such as stability of Rutherford nuclear atom and the hydrogen line spectra (Balmer, Paschen etc.). Indeed, another general chemistry textbook expressed a similar idea in picturesque terms: “Niels Bohr had tied the unseen (the interior of the atom) with the seen (the observable lines in the hydrogen spectrum)—a fantastic achievement” (Joesten, Johnston, Netterville, & Wood, 1991, p. 78). A general physics textbook (see Appendix 10) presented a different perspective by emphasizing the relationship between the consistency of a theory and its success, in the following terms: In 1913 Niels Bohr proposed his famous theory of the hydrogen atom. One cannot say that he resolved the problems raised by Rutherford. In a sense he crystallized the dilemma in an even more dramatic form. Focussing his attention entirely on the construction of a nuclear atom, Bohr took what principles of classical physics he needed and added several nonclassical hypotheses almost without precedent; the mélange was not consistent. But they formed a remarkably successful theory of the hydrogen atom. It would be years before it could be said that one had a consistent theory again” (Cooper, 1970, p. 325).

In other words, consistency was not an immediate objective of Bohr’s theory provided it successfully explained stability of Rutherford’s nuclear atom. At this stage it is interesting to compare the previous presentations (Dickerson et al., 1984; Joesten et al., 1991; Cooper, 1970) with that of Sandin (1989), a general physics textbook (see Appendix 10) that considered Bohr to be a cautious “revolutionary” as he accepted the unsupported hypothesis of quantization (counterinduction) but still maintained the previous conceptual framework: “Bohr was a careful revolutionary. He did not ‘throw out the baby with the bath water’ as many impetuous revolutionaries do. He quantized the angular momentum, but did not do away with Coulomb’s law, centripetal acceleration, or Newton’s laws” (p. 135).

128

7  Feyerabend’s Counterinduction and Science Textbooks

7.7  Photoelectric Effect During the second half of the nineteenth century, light was considered to be a wave propagating in an all-pervading medium. Properties such as diffraction, interference, and polarization convinced physicists that visible monochromatic light is a periodic transverse oscillation. The photoelectric effect is generally considered to be a byproduct of Hertz’s (1887) experimental demonstration of electromagnetic waves. Hertz (1883) had previously shown that cathode rays were a type of wave in the ether similar to light. Later, Thomson (1897) conclusively showed that cathode rays consisted of charged particles (electrons). By 1889, the photoelectric effect had assumed a special significance for many scientists in different parts of Europe and consisted primarily of the following observation: illuminating a metal plate with ultraviolet light initiates a flow of negatively charged particles from the plate. The nature of the photoelectric current was not clear and led to considerable controversy. Einstein (1905) proposed that ordinary light behaves as though it consists of a stream of independent localized units of energy that he called lightquanta. His suggestion of this hypothesis arose from the close analogy he perceived between the behavior of radiation and the behavior of a gas (Wheaton, 1983, p. 106). According to Einstein, if light consists of localized quanta of energy, an electron in an atom will receive energy from only one lightquantum at a time. Monochromatic light of frequency ν can, therefore, grant electrons only energy, hv, where h is Planck’s constant. If some small part p of that energy must be used to release the electron from the metal itself, all electrons of charge e so released will be stopped by a decelerating potential P, following the relation:

½ mu 2 = Pe = hn – p

(where ½ mυ2 is the maximum kinetic energy of the ejected electrons). Einstein (1905) predicted that the stopping potential would be a linear function of the frequency of the incident light, and its slope (h) should be independent of the nature of the substance investigated (cf. Jammer, 1966, p. 35). This became the cornerstone of Robert Millikan’s research program. Interestingly, according to Wheaton (1983), “Einstein’s hypothesis of lightquanta was not taken seriously by mathematically adept physicists for just over fifteen years. The reasons are clear. It seemed to be an unnecessary rejection of the highly verified classical theory of radiation … How lightquanta could possibly explain interference phenomena was always the central objection” (p. 109). Millikan (1916) devoted considerable effort to the experimental determination of Planck’s constant h based on Einstein’s photoelectric equation (presented above), and his results were never questioned. However, incredible as it may sound, Millikan in the same publication accepted Einstein’s photoelectric equation, but questioned the hypothetical lightquantum, the theoretical base of the photoelectric theory: It was in 1905 that Einstein made the first coupling of photo effects … with any form of quantum theory by bringing forward the bold, not to say the reckless, hypothesis of an

7.7  Photoelectric Effect

129

electro-magnetic light corpuscle of energy hν, which was transferred upon absorption to an electron. This hypothesis may well be called reckless first because an electromagnetic disturbance which remains localized in space seems a violation of the very conception of an electromagnetic disturbance, and second because it flies in the face of the thoroughly established facts of interference. The hypothesis was apparently made solely because it furnished a ready explanation of one of the most remarkable facts brought to light by recent investigations, viz., that the energy with which an electron is thrown out of a metal by ultraviolet light or X-rays is independent of the intensity of the light while it depends on its frequency. This fact alone seems to demand some modification of classical theory or, at any rate, it has not yet been interpreted satisfactorily in terms of classical theory (Millikan, 1916, p. 355, italics added).

This passage describes succinctly the photoelectric effect, its explanation by Einstein, experimental determination of h based on Einstein’s photoelectric equation, the controversy surrounding the photoelectric effect and the classical wave theory of light. Interestingly, many scientists both in the U.S.A., and Europe shared Millikan’s thoughts on the subject, which is the fact that Millikan referred to— energy of the ejected electrons in photoelectric effect depends on the frequency of light (Einstein’s hypothesis) and not on the intensity of light (classical wave theory). Despite this acknowledgment, based on his presupposition, Millikan considered Einstein’s hypothesis as reckless—a clear example of counterinduction, viz., accepting an unsupported hypothesis (wave theory), as Einstein’s hypothesis did not explain interference, diffraction and other phenomena. Of course, it can be argued that the wave theory explained some phenomena (e.g., interference, diffraction) that Einstein’s lightquantum could not. Millikan’s adherence to the wave theory of light as a presupposition has been corroborated by Holton (1999) in the following terms: What we now refer to as the photon was, in Millikan’s view, a ‘bold, not to say the reckless, hypothesis’—reckless both because it was contrary to such classical concepts as light being a wave propagation phenomenon, and because of the ‘facts of interference’ … In the background we glimpse the presence of Michelson, the ‘Artist of Light,’ who was Millikan’s admired patron and colleague at the Ryerson Laboratory, the 1907 Nobelist, famous for his interferometers, the work carried out with their aid—and for his adherence to ether physics to his death in 1931 (p. 232).

For textbook authors and classroom teachers, Holton (1999) has raised a very pertinent question: “So Millikan’s (1916) paper is not at all , as we might now naturally consider it to be, an experimental proof of the quantum theory of light” (p. 232). Interestingly, Niaz, Klassen, McMillan and Metz (2010) have reported that of 103 general physics textbooks (Published in U.S.A., between 1950s to 2000s, see Appendix 10) evaluated, almost 87% considered Millikan’s determination of h, as an experimental proof of the quantum theory of light and following is one example: Einstein’s first paper on light-quanta, “On a Heuristic View Concerning the Generation and Conversion of Light,” was published in 1905 (before Relativity). Heuristic means something that serves as a guide in the solution of a problem but is otherwise itself unproved. In that spirit, he postulated that every electromagnetic wave of frequency f is actually a stream of energy quanta, each with energy:



E = hf = hc / l

130

7  Feyerabend’s Counterinduction and Science Textbooks

… A photon colliding with an electron in a metal can vanish, imparting essentially all of its energy to the electron. (Hecht, 2003, emphasis in original, p. 1004) … A photon’s energy goes into freeing the electron, and whatever is left appears as KE [kinetic energy].When the electron is at the surface, the liberating energy is at a minimum, and the electron takes on a KE given by:



hf = KE max + j This wonderfully simple expression is known as Einstein’s Photoelectric Equation, and it explains every aspect of the effect. Since KEmax = hf − φ, increasing the intensity (i.e., irradiance) of the light leaves the maximum kinetic energy unchanged. Only by changing f is the KEmax, or equivalently, the stopping potential, changed for a given metal. (Hecht, 2003, emphasis in original, underline added, p. 1005).

These results (Niaz, et al., 2010), show that Holton (1999) was right that although at present we may consider Millikan’s (1916) paper an experimental proof of the quantum theory of light, it was not considered as such by Millikan himself. Another general physics textbook went beyond by raising the issue of novel ideas and men of genius: Although Millikan showed that this equation [Einstein’s] agreed with his experiment in every detail he himself remained unconvinced that Einstein’s light particles were real. He wrote of Einstein’s “… bold, not to say reckless, hypothesis …” and wrote further that Einstein’s photon concept “… seems at present to be wholly untenable.” Planck, the very originator of the constant, h, did not at once accept Einstein’s photons either. In recommending Einstein for membership in the Royal Prussian Academy of Sciences in 1913 he wrote: “that he may sometimes have missed the target in his speculations, as for example in his theory of light quanta, cannot really be held against him.” Far from being unusual it is almost commonplace that truly novel ideas are accepted only slowly, even by such men of genius as Millikan and Planck. (Halliday & Resnick, 1981, p. 783).

If we accept the thesis that Millikan’s rejection of lightquanta in Einstein’s Photoelectric Equation can be considered as an example of rejecting counterinduction (i.e., accepting inductivism), then it seems plausible to suggest that counterinduction being a novel idea may even be rejected by men of genius (Planck, Millikan and others as examples in the case of the photoelectric effect). At this stage it is important to note that most scientists in the early twentieth century believed in inductivism (including Millikan) and counterinduction was thus difficult to accept. To pursue the subject further, it was decided to explore the presentation of the photoelectric effect in general chemistry textbooks published (between 1950s to 2000s, see Appendix 9) in U.S.A. Of the 118 textbooks evaluated almost 52% concluded that Einstein’s postulate of light quanta explained the photoelectric effect and following is an example (for details see Ospina, 2010; Niaz, 2016, chap. 7): All these observations [related to the photoelectric effect], could be explained by combining Planck’s idea of energy quanta with the notion that light could be described not only as having wave-like properties but also as having particle-like properties. Einstein assumed that these massless “particles,” now called photons, carry the energy given by Planck’s hypothesis; that is, the energy of each photon in a stream of photons is proportional to the frequency of its wave.

7.8  Wave-Particle Duality



131

E photon = hn



With Einstein’s revolutionary postulate, the photoelectric effect can be explained. It is easy to imagine that a high-energy particle would have to bump into an atom to cause the atom to lose an electron (Kotz & Treichel, 1999, p. 300, emphasis in original, underline added). Interestingly, some of the general chemistry textbooks do not explicitly refer to Millikan’s experimental determination of h. However, the reference to “all these observations” (see above Kotz and Treichel, 1999), implicitly does include Millikan’s experimental data. Furthermore, it is important to note that 87% of general physics textbooks referred to the issue involved (Holton’s thesis discussed above), as compared to 52% of general chemistry textbooks. This could perhaps be attributed to the fact that determination of a physical constant (Planck’s h) perhaps belongs more to the physics curriculum. It is plausible to suggest that the inclusion in science textbooks of the following aspects related to the photoelectric effect can facilitate a better understanding of the dynamics of scientific progress: (a) Millikan considered Einstein’s hypothesis as reckless—in other words accepting Einstein’s photoelectric equation constituted a clear example of counterinduction, viz., accepting unsupported hypothesis; (b) Einstein’s hypothesis was not accepted by the scientific community, including Planck, the ‘originator’ of the quantum hypothesis, for many years; (c) Millikan presented experimental evidence to support Einstein’s photoelectric equation and still rejected his quantum hypothesis; (d) scientific theories are underdetermined by experimental evidence, that is, no amount of experimental evidence can provide conclusive proof for a theory (these aspects can be included in the textbooks by presenting Millikan’s experimental determination and at the same time pointing out that this was not considered as sufficient evidence for Einstein’s theory); (e) scientists customarily have prior theoretical beliefs or presuppositions before doing an experiment, and they resist any change in those epistemological beliefs. Inclusion of these aspects could also help to facilitate an integrated view of the nature of science (NOS) and to introduce the historical aspects of the photoelectric effect as an unfolding story.

7.8  Wave-Particle Duality The origin of wave-particle duality can be traced to Einstein’s (1905) hypothesis of the light quantum to explain the photoelectric effect and this was followed by important theoretical formulations by Einstein (1916) and later De Broglie (1922, 1923a, 1923b, 1924). De Broglie (1922) was the first to attempt study black-body radiation in the context of the light quantum hypothesis. Later, this interest in the properties of quanta motivated De Broglie’s search for a theory that would unify the wave and particle aspects (Medicus, 1974). In 1923, he published three short articles in

132

7  Feyerabend’s Counterinduction and Science Textbooks

Comptes Rendus in which he generalized this wave-particle duality to include material corpuscles (De Broglie 1923a). This was followed by a short note to Nature (De Broglie 1923b) and then the complete article in Philosophical Magazine (De Broglie 1924). De Broglie (1923a) applied the wave-particle duality hypothesis to existing problems in physics and among others he referred to the following important issues. Application of his hypothesis to electron orbits in an atom required that the wave be in phase with itself and that the circumference be an integral multiple of the wavelength and this led De Broglie to conclude: “We believe that this is the first physically plausible explanation for the Bohr-Sommerfeld stability rules” (Reproduced in Medicus 1974, p. 40). The famous formula λ = h/mv is found in this explicit form for the first time in the chapter on statistical mechanics. For De Broglie it is not the wavelength of the particle but the frequency that is more important and that there was no essential difference between photons and particles. One of the most important aspects of this article was that De Broglie even suggested a possible experimental confirmation of his hypothesis in the following terms: “A beam of electrons passing through a very small opening could present diffraction phenomena. This is perhaps the direction in which one may search for an experimental confirmation of our ideas” (Reproduced in Medicus 1974, p.  40). Despite Einstein’s prestige and authority, duality remained a controversial hypothesis, until conclusive experimental evidence was presented by Davisson and Germer (1927), based on their work with nickel crystals, which led them to conclude: Because of these similarities between the scattering of electrons by the crystal and the scattering of waves by three- and two- dimensional gratings a description of the occurrence and behavior of the electron diffraction beams in terms of the scattering of an equivalent wave radiation by the atoms of the crystal, and its subsequent interference, is not only possible, but most simple and natural. This involves the association of a wave-length with the incident electron beam, and this wave-length turns out to be in acceptable agreement with the value h/mv of the undulatory mechanics… (Davisson and Germer 1927, p. 707).

λ  =  h/mv, of course refers to the famous formula first presented by De Broglie in 1923. Most general chemistry textbooks (see Appendix 9) present atomic structure by referring to the work of J.  J. Thomson, R.  Millikan, E.  Rutherford and N.  Bohr. Following this, Einstein’s interpretation of the photoelectric effect is presented as an application of quantum theory. Bohr’s model of the atom was the first to depart from the classical wave theory of light by introducing the ‘quantum of action’. Next, in order to introduce the wave mechanical model of the atom (E.  Schrödinger), De Broglie’s contribution is mentioned by posing the question: if light can have both wave and particle properties then why particles of matter (for example, electrons) cannot also have both properties. Furthermore, experimental work of C. Davisson and L. H. Germer is reported based on diffraction of electron beams by metal foils. Based on a historical reconstruction of the wave-particle duality, Niaz and Marcano (2012) elaborated six criteria for evaluating 128 general chemistry ­textbooks published (1954 to 2011, see Appendix 9) in U.S.A. In one of the criterion, textbooks were evaluated to find if they referred to the following aspects: (a) Origin of wave-particle duality can be attributed to Einstein and de Broglie, and

7.8  Wave-Particle Duality

133

(b) Experimental evidence was presented later by Davisson and Germer (1927). Of the 128 textbooks evaluated, 26 (20%) had a satisfactory presentation and following are two examples: Energy, prior to 1900, was not considered to consist of particles. It was noncorpuscular in nature, and therefore continuous. It was this distinction between matter and energy that had been abandoned by Planck in 1900, by Einstein in 1905, and again by Bohr in 1913 … The French physicist Louis de Broglie proposed in 1924 that not only light but all matter has a dual nature and possesses both wave and corpuscular properties. He reasoned that there should be symmetry in nature: If a radiant corpuscle—that is, a photon—has a frequency and a wavelength and therefore has wave properties, why should not a material particle also have wave properties? (p. 429, original italics) …When de Broglie first published his wave theory of matter, there was no experimental evidence to support his bold hypothesis. Within three years, however, two different experiments had been performed that demonstrated the diffraction of a beam of electrons. Clinton J.  Davisson, assisted by L.H.  Germer, … observed the diffraction of electrons when a beam of electrons was directed at a nickel crystal (Segal, 1989, p. 431, original italics, underlined added).

Interestingly, the presentation by Segal (1989) is almost a historical reconstruction of wave-particle duality, starting with the contributions of Planck in 1900 to Einstein in 1905, Bohr in 1913, de Broglie in 1924, and finally 3 years later Davisson and Germer (1927). The underlined part is very important as very few textbooks refer to it, namely De Broglie’s conceptualization of wave-particle duality preceded its experimental determination by Davisson and Germer. The argument of symmetry in nature can be particularly helpful for students. Furthermore, it is plausible to suggest that the reference to bold hypothesis can be considered as an example of counterinduction. The following presentation from another general chemistry textbook was also considered as satisfactory: Einstein used the photoelectric effect to demonstrate that light, which is usually thought of as having wave properties, can also be thought about in terms of particles or massless photons. This fact was pondered by Louis Victor de Broglie (1892–1987). If light can be considered as sometimes having wave properties and other times having particle properties, he asked why doesn’t matter behave similarly? That is, could a tiny object such as an electron, which we have so far considered a particle, also exhibit wave properties in some experiments? … This idea was revolutionary, since it linked the particle properties of the electron (m and ʋ) with possible wave properties (λ). Experimental proof was soon produced. Davisson and Germer, … found that a beam of electrons was diffracted like light waves by the atoms of a thin sheet of metal foil and that de Broglie’s relation was followed quantitatively … After de Broglie’s suggestion that an electron can be described as having wave properties, a great debate raged in physics. How can an electron be described as both a particle and a wave? … One can only conclude that the electron has dual properties. The result of a given experiment can be described either by the physics of waves or by the physics of particles; there is no single experiment that can be done to show that the electron behaves simultaneously as a wave and a particle! (Kotz & Purcell, 1991, pp.  295–297, original italics, underlined added).

Again, the presentation by Kotz and Purcell (1991) is very much in accord with the historical record and the underlined part adds a new dimension with respect to the “debate raged in physics” and how scientists went about resolving this dilemma. Indeed, Feyerabend would endorse that science involves not only “hypothesizing”

134

7  Feyerabend’s Counterinduction and Science Textbooks

but also “pondering” that leads to “counterinduction.” On comparing different textbooks, a thoughtful student may wonder if these two textbooks (Segal, 1989; Kotz & Purcell, 1991) are presenting chemistry or history of chemistry. Indeed, this is the dilemma faced by most chemistry teachers and textbook authors (cf. Niaz, 2017, for a recent appraisal). However, a critical appraisal of most of our current textbooks would show that if we want to understand chemistry, its history cannot be ignored. In other words, the history of chemistry is ‘inside’ chemistry (Niaz & Rodriguez, 2001), and according to Feyerabend (1993, p. 21) history of science becomes an inseparable part of the science itself.

7.9  Mendeleev’s Periodic Table of Chemical Elements Most chemistry teachers consider the periodic table to be an important concept, both in principle and practice. It facilitates a succinct organization and understanding of the fundamental building blocks of chemistry, the chemical elements. Mendeleev’s periodic table formed part of his textbook (Principles of Chemistry, first written between 1868 and 1870), in which he endeavored to facilitate students’ understanding of methods of observation, experimental facts, laws of chemistry, and the “… unchangeable substratum underlying the varying forms of matter” (author’s preface to the sixth Russian edition, reproduced in Mendeleev, 1897, p. vii). Most historians consider the international congress held in Karlsruhe (September 3–5, 1860) as crucial in the development of chemistry. A circular (dated July 10, 1860) sent by the organizers of the congress to most outstanding chemists of Europe outlined its objective as the need to reach a consensus on “More precise definitions of the concepts of atom, molecule, equivalent, atomicity, alkalinity, etc.; discussion on the true equivalents of bodies and their formulas; initiation of a plan for a rational nomenclature (reproduced in de Milt, 1951, p. 421). Mendeleev (then 26 years of age) attended the congress and was greatly impressed by Cannizaro’s contribution and, in a letter dated September 7, 1860, summarized an important achievement of the congress: It is decided to take a different understanding of molecules and atoms, considering as a molecule the amount of a substance entering a reaction and determining physical properties, and considering as an atom the smallest amount of a substance included in a molecule. Further, it reached an understanding about equivalents, considered as empirical, not depending on the understanding about atoms and molecules (reproduced in de Milt, 1951, p. 422).

Mendeleev (1869) enunciated the first form of his periodic law and later elaborated in the following terms: “The properties of simple bodies, the constitution of their compounds, as well as the properties of these last, are periodic functions of the atomic weights of elements (Mendeleev, 1879, p. 267). It is important to note that the elucidation of the concept of atomic weight by Stanislao Cannizaro at Karlsruhe

7.9  Mendeleev’s Periodic Table of Chemical Elements

135

was crucial in the discovery of the periodic law. According to van Spronsen (1969), elaboration of the periodic table was difficult and took a long time due to “… lack of a definite conception of atomic weight, which is very closely connected with the definitions of molecules and atoms” (p. 565). Availability of the atomic weights of about 60 elements enabled Mendeleev to accommodate the elements in the table according to various physicochemical properties (density, specific heat, atomic weight, atomic volume, melting point, valence, oxides, chlorides, and sulfides). According to van Spronsen (1969): The actual development of the periodic system seemed to require a catalyst! We think it proper to attribute this catalytic action to Cannizaro’s famous Karlsruhe lecture at the 1860 Congress. He made the distinction between atoms and molecules and defined such concepts as valence; … this initiated the second stage of the discovery and started the history proper of the periodic system of chemical elements. (p. 1, italics added).

Many chemistry students must have wondered as to how Mendeleev and the other co-discoverers could have conceptualized the underlying theoretical rationale of the elements that manifested itself in periodicity. Could “counterinduction” be the answer? It is important to recall that most of the pioneering work of Mendeleev was conducted from 1869 to 1889, before Thomson (1897), Rutherford (1911), Bohr (1913), and Moseley (1913, 1914) laid the foundations of the modern atomic theory. So how could Mendeleev conceptualize periodicity as a function of the atomic theory? An answer to this question will precisely show Mendeleev’s ingenuity, far-­ sightedness, creativity, and the ability to “speculate.” Despite Mendeleev’s own ambivalence and ambiguity, a historical reconstruction does provide a convincing story of this remarkable contribution to our knowledge. Before presenting the reconstruction it is important to note that Mendeleev had the following important sources of information: Dalton’s atomic theory; law of multiple proportions; Cannizaro’s Karlsruhe lecture; fairly reliable atomic weights; atomicity (valence); and various physical and chemical properties of the elements. Following five steps were important in Mendeleev’s reasoning: 1. In his first publication, Mendeleev referred to the relationship, albeit implicitly, between periodicity, atomic weights, and valence: “The arrangement according to atomic weight corresponds to the valence of the element and to a certain extent the difference in chemical behavior, for example Li, Be, B, C, N, O, F (Mendeleev, 1869, p. 405, original italics). 2. After the discovery of gallium and scandium, elements predicted by Mendeleev that were discovered in 1875 and 1879 respectively, he expressed the relationship between atomic weight and atomic theory much more explicitly: It is by studying them [atomic and molecular weights], more than by any other means, that we can conceive the idea of an atom and of a molecule. By this fact alone we are enabled to perceive the great influence that studies carried on in this direction can exercise on the progress of chemistry … The expression atomic weight∗ implies, it is true, the hypothesis of the atomic structure of bodies. (Mendeleev, 1879, p. 243, emphasis added)

136

7  Feyerabend’s Counterinduction and Science Textbooks

The asterisk leads the reader to the following footnote: “By replacing the expression of atomic weight by that of elementary weight, I think we should, in the case of elements, avoid the conception of atoms.” This footnote shows Mendeleev’s ambiguity/ambivalence toward the atomic theory and will be dealt with later. 3. Another example of Mendeleev’s ambivalence can be observed from the following: “I shall not form any hypotheses, either here or further on, to explain the nature of the periodic law; for, first of all, the law itself is too simple∗ …” (Mendeleev, 1879, p. 292). The asterisk leads the reader to the following footnote: “However, I do not ignore that to completely understand a subject we should possess, independently of observations [and experiences] and of laws [as well as systems], the meanings of both one and the other”. This is a clear reference to the understanding of atomic theory and atomic weight, even if experimental observations were lacking—an implicit reference to counterinduction. 4. Although Mendeleev stated in 1879 that he would not formulate an hypothesis, 10 years later in his famous Faraday lecture, Mendeleev (1889) not only attributed the success of the periodic law to Cannizaro’s ideas on the atomic theory (pp. 636–637) but went on to explicitly formulate the following hypothesis: “… the veil which conceals the true conception of mass, it nevertheless indicated that the explanation of that conception must be searched for in the masses of atoms; the more so, as all masses are nothing but aggregations, or additions, of chemical atoms …” (Mendeleev, 1889, p. 640, italics added). Readers at this stage will be reminded of Newton’s public stance of “hypotheses non fingo” (I do not feign hypotheses). Precisely, Newtonian method has been the subject of criticism by Duhem (1914), who considered it to be more of a dream and that in order to formulate his laws Newton inevitably resorted to idealization. Did Mendeleev follow the same “method”? And if the answer is in the affirmative, can it considered as an instance of “counterinduction”, namely, Mendeleev accepted hypotheses (atomic theory & atomic weight) that were at best partially supported by experimental evidence, and some critics even considered that the atomic theory (during the nineteenth century) had no experimental support. 5. Later in the same Faraday Lecture, Mendeleev went beyond and clearly acknowledged the role played by the atomic theory to explain periodicity in the periodic table: … the periodic law has clearly shown that the masses of the atoms increase abruptly, by steps, which are clearly connected in some way with Dalton’s law of multiple proportions; … While connecting by new bonds the theory of the chemical elements with Dalton’s theory of multiple proportions, or atomic structure of bodies, the periodic law opened for natural philosophy a new and wide field for speculation. (Mendeleev, 1889, p.  642, italics added).

Interestingly, despite Mendeleev’s ambivalence (see above) towards the atomic theory, he seems to be considering the law of multiple proportions in chemistry as synonymous with Dalton’s atomic theory (for further details see: Niaz et al., 2004; Brito et al. 2005). Furthermore, in the context of Feyerabend’s hypothesis of “coun-

7.9  Mendeleev’s Periodic Table of Chemical Elements

137

terinduction”, Mendeleev’s reference to a “wide field for speculation” is all the more significant. At this stage it is important to note that there is considerable controversy among philosophers of science with respect to the status of Mendeleev’s oeuvre: ordered domain, empirical law, hypothesis or a theory (Bensaude-Vincent, 1986; Brush, 1996; Shapere, 1977; Wartofsky, 1968; Weisberg, 2007; Ziman, 1978). In my opinion, Wartofsky’s (1968) thesis comes quite close to what Mendeleev was doing: Mendeleev, for example, predicted that the blank space of atomic number 32, which lies between silicon and tin in the vertical column, would contain an element which was grayish-­white, would be unaffected by acids and alkalis, and would give a white oxide when burned in air, and when he predicted also its atomic weight, atomic volume, density and boiling point, he was using the periodic table as a hypothesis from which predictions could be deduced. This was in 1871. (p. 203)

It is important to note Wartofsky’s advice: “This was in 1871”, that is long before the periodic law was changed and came to be based on atomic number (as compared to atomic weight) after the work of Moseley (1913, 1914). In this particular case, Mendeleev’s hypothesis did not have experimental evidence (as was provided starting in 1897 by the work of Thomson, Rutherford and Moseley) and hence it amounted to counterinduction. A general chemistry textbook (see Appendix 9) referred to a problem faced by Mendeleev in the following terms: “Two elements, tellurium (Te) and iodine (I), caused Mendeleev problems. According to the best estimates at that time, the atomic mass of tellurium was greater than that of iodine. Yet if these elements were placed in the table according to their atomic masses, they would not fall into the proper groups required by their properties. Therefore, Mendeleev switched their order and in so doing violated his own periodic law. (Actually, he believed that the atomic mass of tellurium had been incorrectly measured, but this wasn’t so.)” (Brady, Russell, & Holum, 2000, p. 63). In order to solve the problem, Mendeleev reduced the atomic mass of Te from 128 to 125 (present day actual atomic masses: Te = 127,6; I = 126,9). Thus by reducing the atomic mass of Te it preceded that of I, which could then be placed in the same group as F, Cl and Br. Similarly, Mendeleev also changed the following atomic weights: Be from 14 to 9, Uranium from 120 to 240. Could these changes be considered as examples of counterinduction? In my opinion “Yes.” In other words, Mendeleev’s periodic law (ascending order of atomic weights) was not supported by the experimental data (atomic weights of Te and I), and consequently he changed the data (128 to 125), and thus accepted an unsupported hypothesis—counterinduction. Despite Mendeleev’s considerable expertise in predicting new chemical elements, for which he left empty spaces in his periodic system, he was taken by surprise by W. Ramsey’s discovery in 1894 of a new element, later named as argon—the inert one. The new inert gas was announced by J.W.  Strutt (Lord Rayleigh) and Ramsey at a meeting of the Royal Society on January 31, 1895. This discovery led to an intense debate as to the nature of the gas and its place in the periodic table, and was ultimately a crucial test for Mendeleev’s periodic law (cf. van Spronsen, 1969).

138

7  Feyerabend’s Counterinduction and Science Textbooks

Ramsey (1897) recalled the events at a meeting of the British Association in the following terms: The discovery of argon at once raised the curiosity of Lord Rayleigh and myself as to its position in this table [Mendeleev’s]. With a density of nearly 20, if a diatomic gas, like oxygen and nitrogen it would follow fluorine in the periodic table … But when the ratio of its specific heats [Cp/Cv] had, … unmistakably shown that it was molecularly monoatomic, and not diatomic, as at first conjectured, it was necessary to believe that its atomic weight was 40, and not 20, and that it followed chlorine in the periodic table, and not fluorine. But here arises a difficulty. The atomic weight of chlorine is 35.5, and that of potassium, the next element in order in the table, is 39.1; and that of argon, 40, follows, and does not precede, that of potassium, as it might be expected to do (p. 379).

This clearly shows how the placing of an element was not a straightforward question of ordering the elements in the ascending order of their atomic weights, and the placing of argon created serious problems. These difficulties led some chemists, including W.A. Rucker (President of the Royal Society), to entertain the possibility of even rejecting the periodic table itself (cf. Rucker & Kelvin, 1895, p.  62). Mendeleev himself at first refused to accept argon as an element. Instead, at a meeting of the Russian Chemical Society on 2 March 1895, he postulated the hypothesis that argon could be triatomic nitrogen (N3), in analogy to ozone (O3), and that would also explain why it did not react with any other element. This is a clear example of counterinduction as Mendeleev had no empirical evidence to postulate argon as triatomic nitrogen. However, in 1903 changed his mind for among other the following reasons: (a) The finding that argon’s density was barely 19, whereas N3 would have been 21; (b) Discovery of other inert gases, such as helium, neon and krypton; (c) The uniqueness of the spectra of the inert gases. Based on this evidence and arguments, and his constant consultations with Ramsay, he now postulated that the inert gases could be considered as zero-valency 0-group to be placed on the far left of the periodic table (this was later changed and placed on the far right). Mendeleev’s argument seemed to rest on an arrangement from least reactive (the inert gases) to most reactive (the halogens). Interestingly, this even countered an essential aspect of Mendeleev’s periodic law namely periodicity as a function of valency (for further details see Gordin, 2004, p.  211). The placement of argon was finally solved by placing argon and the other noble gases in a new group between the halogens and the alkali metals (Ramsey & Travers, 1901). After his initial reservations, Mendeleev considered the accommodation of argon as a glorious confirmation of the general applicability and validity of the periodic law. These events took place almost 25  years after Mendeleev’s initial work on the periodic table, which shows that progress in science is continually faced with difficulties and challenges, thus requiring new hypotheses (at times with insufficient empirical support—counterinduction). One general chemistry textbook expressed this dilemma in the following terms: The table that Mendeleev developed is in many ways similar to the one we use today. One of the main differences though, is that Mendeleev’s table lacks the column containing the elements helium (He) through radon (Rn). In Mendeleev’s time, none of these elements had yet been found because they are relatively rare and because they have virtually no tendency

7.9  Mendeleev’s Periodic Table of Chemical Elements

139

to undergo chemical reactions. When these elements were finally discovered, beginning in 1894, another problem arose. Two more elements, argon (Ar) and potassium (K), did not fall into the groups required by their properties if they were placed in the table in the order required by their atomic masses. Another switch was necessary and another exception to Mendeleev’s periodic law had been found (Brady, Russell, & Holum, 2000, p. 63, italics added).

This is an interesting presentation despite the fact that it does not refer to counterinduction nor the fact that at first Mendeleev refused to accept argon as an element. Overall, it follows the historical record by pointing out that the placing of argon led to another exception to Mendeleev’s periodic law. In contrast, another general chemistry textbook presented the discovery of the noble gases in the following terms: The periodic table was devised in 1869, but it was not until 1892 that argon, the first of the noble gases, was discovered. Of interest here is the fact that the periodic table was built to accommodate the then known elements, but with the discovery of the noble gases about a quarter of a century after the table was described, a whole family of elements fitted into the table without any necessity of revising it (Lippincott, Garrett & Verhoek, 1968, p. 302, italics in the original, underline added).

It is important to note that W.T. Lippincott, first author of this book, was an Editor (1967–1979) of the Journal of Chemical Education, published by the Educational Division of the American Chemical Society. However, to state that the placing of noble gases was accomplished without any “necessity of revising” the periodic table, is almost a distortion of the historical facts. A historical reconstruction clearly shows that the discovery of argon in 1894, led to: a controversy with respect to its place in the periodic table ➔ Mendeleev was reluctant to accept argon as an element ➔ Mendeleev even presented an alternative hypothesis, namely argon as triatomic nitrogen: counterinduction ➔ Discussions and consultations among members of the scientific community finally led to the placing of argon in a separate group between the halogens and the alkali metals. Inclusion of such details is an essential part of how the periodic table developed and progressed and even shows how the history of science becomes a part of the science itself, as suggested by Feyerabend (1993, p. 21). Brito, Rodríguez and Niaz (2005) have evaluated 57 general chemistry textbooks (published in U.S.A., between 1960s to 2000s, see Appendix 9) and found that none of the textbooks presented satisfactorily the following aspects of Mendeleev’s contribution: (a) How to explain periodicity of elements in the periodic table; and (b) Was Mendeleev’s contribution a theory or an empirical law. However, it is interesting to study and compare some of the following textbook presentations: Following are two examples of an inductive generalization: Mendeleev’s approach to the periodic table was empirical; he based his classification scheme on the observed facts. (Hill & Petrucci, 1999, p. 316, original italics) The periodic table was created by Mendeleev to summarize experimental observations. He had no theory or model to explain why all alkaline earths combine with oxygen in a 1:1 atom ratio—they just do. (Moore et al., 2002, p. 266)

140

7  Feyerabend’s Counterinduction and Science Textbooks

In contrast, McMurry and Fay (2001) provide an example of how Mendeleev’s contribution can be considered a theory: In many ways, the creation of the periodic table by Dmitri Mendeleev in 1869 is an ideal example of how a scientific theory comes into being. At first, there is only random information—a large number of elements and many observations about their properties and behavior. As more and more facts become known, people try to organize the data in ways that make sense, until ultimately a consistent hypothesis emerges (p. 160, italics added).

This presentation, however, ignores the role played by Dalton’s atomic theory in the postulation of Mendeleev’s periodic law. In what follows are two examples of textbooks in which an attempt was made to highlight the role played by atomic theory in the development of the periodic table: Many atomic masses were determined as a direct result of Dalton’s postulates and the work that they stimulated, and scientists attempted to relate the atomic masses of the elements to the elements’ properties. This work culminated in the development of the periodic table by Dmitry Mendeleyev (1834–1907) … (Goldberg, 2001, pp. 87–88). Early in the nineteenth century, when Dalton’s atomic theory was winning general acceptance, the first attempts were made toward classification of the elements into groups or families on the basis of similarities of physical and chemical properties … even in its primitive form as stated in 1869, this [periodic] law clearly pointed to regularities that hinted at an orderly subatomic structure of matter and provided a tremendous stimulus toward seeking to understand the internal structure of atoms, as chemists and physicists sought to construct an atomic model that would explain Mendeleev’s generalization (Sisler et al., 1980, p. 150, italics added).

As we have observed in this study the development of the periodic table is much more complex. Presentations of the first two textbooks are entirely empirical, emphasizing inductive generalization. The next presentation recognizes the role played by “emerging hypotheses” that can facilitate a better understanding of the vicissitudes faced by Mendeleev and others, in their struggle to go beyond the observable entities. The last two presentations explicitly refer to Dalton’s atomic theory and how regularities facilitated an understanding of the internal structure of atoms, without of course, postulating the electronic structure of atoms. These presentations from textbooks clearly show the importance of understanding “science in the making” for understanding the genesis of the periodic table. Before discussing the educational implications of Mendeleev’s contribution it is important to explore his ambiguity/ambivalence toward the atomic theory. Throughout the nineteenth century, positivism was the dominant philosophy, which led all scientific work to be based strictly on experimental observations and all hypothetical propositions were considered speculative and hence nonscientific (Brush, 1976; Gavroglu, 2000; Holton, 1992). Mendeleev was clearly aware of this and on many occasions went out of his way to emphasize that the periodic “… law itself was a legitimate induction from the verified facts” (Mendeleev, 1889, p. 639). In the Faraday lecture, Mendeleev emphasized the inductive aspect of the periodic law in the light of the anti-atomist Marcellin Berthelot’s (1827–1907) criticism, “… the illustrious Berthelot, in his work Les origins de l’ Alchimie, 1885, 313, has sim-

7.9  Mendeleev’s Periodic Table of Chemical Elements

141

ply mixed up the fundamental idea of the law of periodicity with the ideas of Prout, the alchemists, and Democritus about primary matter. But the periodic law, based as it is on the solid and wholesome ground of experimental research, has been evolved independently of any conception as to the nature of the elements; …” (Mendeleev, 1889, p.  644). Apparently, Mendeleev’s dilemma was that, on the one hand, he could rightly claim that the periodic law was based on experimental properties of the elements (an aspiration of scientists in the late nineteenth century), and yet he could not give up the bigger challenge, viz., the possible causes of periodicity, and hence importance of atomic theory. Rutherford (1915, p. 176), himself an experimentalist par excellence, described in eloquent terms the positivist intellectual milieu of the late nineteenth century. The periodic table of chemical elements forms part of almost every high school and introductory university level chemistry textbook (and curriculum) published on this planet. Despite its usefulness as a conceptual tool for organization of the chemical elements, understanding their properties, predicting new elements and a corrective device, most students consider it to be a difficult topic. Niaz and Luiggi (2014) designed a teaching strategy based on the historical aspects in order to facilitate freshman students’ conceptual understanding of the periodic table. While teaching the periodic table the following aspects related to the history and philosophy of science (HPS) constituted the guiding principles of the teaching strategy: 1. How could a simple arrangement of the elements based on atomic mass (atomic weight for Mendeleev) provided such regularities as observed in the periodic table? 2. Many scientists including Mendeleev were continually trying to understand the underlying reason for periodicity. These efforts went through various tentative attempts to understand and classify the elements. On the contrary most textbooks give the impression that for almost 100 years (1820–1920), scientists had no idea or never asked the question as to whether there could be an underlying rationale for explaining periodicity (for further details see Brito, Rodríguez & Niaz, 2005). Furthermore, textbooks in general ignore the tentative nature of scientific knowledge (for details, see Niaz & Maza, 2011). 3. Besides Mendeleev in 1869, following co-discoverers of the periodic table also made important contributions: De Chancourtois in 1862, Odling in 1864, Meyer in 1864, Newlands in 1865, and Hinrichs in 1866. 4. Even before the modern atomic theory (starting 1897) scientists were well aware that periodicity in the periodic table is a function of the atomic theory. 5. Accommodations (based on physico-chemical properties) and predictions of elements provided important evidence for the acceptance of the periodic law and it would be helpful to emphasize both in the classroom. 6. Based on a historical reconstruction the following sequence of heuristic principles can help to facilitate understanding: Accumulation of atomic weights of the elements in the early nineteenth century, Attempts to classify elements starting in 1817, Karlsruhe congress in 1860, Cannizaro’s contributions on atoms and molecules, Mendeleev’s first periodic table in 1869, Corrections of known

142

7  Feyerabend’s Counterinduction and Science Textbooks

atomic weights, Discovery of argon in 1895, and Contribution of Moseley in 1913 that led to the periodic table being based on atomic numbers instead of atomic weights. 7. Implementation of these guiding principles constitutes what in the history of science and science education literature, has been referred to as “science in the making” (for details, see Niaz, 2012), and Feyerabend referred to it as “how science works.” 8. An effective way in which to bridge the gap between how we teach science (periodic table in this case) and what scientists actually do, that is “science in the making” is through the inclusion of humanizing aspects of the history of science in the form of a story (contextual teaching, Klassen, 2006). Niaz and Luiggi (2014) designed a study based on two groups of freshman students enrolled in a Chemistry I course at a major university in Venezuela. In order to avoid possible interactions, control group students received instruction in a semester prior to that of the experimental group. Both groups were asked to look for information on the periodic table from the internet and traditional textbooks found in the university library. Following this, control group students discussed and solved problems found in the textbooks. Experimental group students were then exposed to the following phases of an experimental treatment: (a) Discussion of various aspects related to HPS (as summarized above); (b) Construction of concept maps (Novak, 1990; students were familiar with these maps from a previous course); (c) Evaluation based on Posttest 1 (Items 1–4); (d) Classroom discussions based on students’ arguments on responses to items of Posttest 1; (d) PowerPoint presentation by the instructor based on various HPS aspects; (e) Construction of new concept maps; (f) Discussion and comparison of the two sets of concept maps; (g) Evaluation based on Posttest 2 (Items 5–7); and (h) Five volunteer students participated in semi-­ structured interviews. Control group students were also evaluated on Posttests 1 and 2 and spent the same time in solving traditional problems, as the experimental group used to receive the treatment. Results obtained showed that Experimental group students provided conceptual responses (as opposed to rhetorical responses) on all items. Item 1, dealt with atomic theory as the criterion used by Mendeleev to order the elements, and 19% of the students responded conceptually. Item 2, dealt with the relationship between the periodic table and the early atomic theory, and 47% responded conceptually. Item 3, dealt with the question as to how Mendeleev could elaborate the periodic table before the modern atomic theory, and 28% responded conceptually. Item 4, asked if the idea of ordering the elements originate with Mendeleev and 13% responded conceptually. Item 7 referred to periodicity as a function of the chemical atoms (i.e., atomic theory) and 13% responded conceptually. It was not expected that control group students would respond conceptually. Nevertheless, one student on Item 1, two students on Item 2 and one student on Item 3 responded conceptually. This is an interesting finding and shows that given the opportunity to reason and reflect, even some control group students can go beyond and improve understanding. In order to

7.9  Mendeleev’s Periodic Table of Chemical Elements

143

further appreciate the reasoning and arguments that support a rhetorical and a conceptual response, let us consider the following responses by two experimental group students on Item 1 of the Posttest, which stated: “In your opinion what was the criterion used by Mendeleev to put the elements in the established order in the periodic table”. Following is an example of a response based on rhetorical reasoning, viz., lacks conceptual understanding: Mendeleev (1869) discovers his periodic system, employs chemical and physical properties, also employs atomic weights in order to support his system, and enunciates that elements when placed according to their atomic weights present a clear periodicity in their properties (Experimental group student, reproduced in Niaz & Luiggi, 2014, italics added).

Following is an example of a response based on conceptual understanding: Mendeleev studied the properties of the elements, such as atomic weight, oxides formed by the reaction with oxygen, density, volume—that is chemical and physical properties. Besides this on occasions he used to alter the atomic weights of some elements in order to make them concordant with the established order. Such criteria permitted him to place new elements in his periodic table, such as eka boron [scandium], eka aluminium [gallium] and eka silicon [germanium] (Experimental group student, reproduced in Niaz & Luiggi, 2014, italics added).

Comparing both responses can help us to understand how arguments can facilitate students to go beyond what they learn from the traditional curriculum and textbooks. It is important to note that both responses can be considered as correct. Nevertheless, the conceptual response provides additional information with respect to: (a) Examples of physical and chemical properties (whereas the rhetorical response simply refers to the properties, see the part in italics in both responses); (b) Even after having enunciated his law, Mendeleev used to alter the atomic weights of some elements (counterinduction?) and hence their place in the periodic table. The example of tellurium and iodine is well known; (c) Mendeleev’s table not only classified the known elements but also predicted new ones. It is plausible to suggest that the conceptual response is a better indicator of how scientists when faced with difficulties and anomalous data, accept changes in their enunciated laws, which are basically working hypotheses and hence tentative. It is well known that Mendeleev did not have the necessary experimental evidence for altering the atomic weights. Item 3 of the Posttest 1 in this study (Niaz & Luiggi, 2014) asked the following question: “If the periodic table was elaborated before the modern atomic theory, how could Mendeleev and others construct the periodic table?” Following is an example of a response that shows conceptual understanding: In the year 1817, Döbereiner proposed a system of triads based on three elements, in which the atomic weight of one of the elements was a mean of the other two. In 1860 at the Karlsruhe Congress, the concepts of atomic weight, equivalent weight and the atom were clarified. In 1866, Hinrichs proposed a periodic system based on the shape of the elements. Later Meyer presented his periodic table based on the atomic theory. In 1869 Mendeleev proposed a periodic table based on the physical and chemical properties and the atomic theory. Mendeleev’s periodic system was ingenious as he left empty spaces in his table for predictions of new elements and corrections. For example, a monoatomic gas [argon] was discovered that questioned the very existence of the periodic table, as it was difficult to find

144

7  Feyerabend’s Counterinduction and Science Textbooks

a place for it. Finally, it was proposed that it could be placed in a separate group between the halogens and alkaline metals. Forty years later, the electron was discovered and it was decided that the position of an element in the periodic table was determined by the number of electrons (Experimental group student, emphasis added, reproduced in Niaz & Luiggi, 2014, p. 19).

On reading this response the question that comes to mind is the following: How could a freshman student come up with such reasoning that constitutes a historical reconstruction, which even many general chemistry textbooks lack (cf. Brito et al., 2005). In other words, before 1897 (Thomson’s cathode ray experiments) those who worked on the periodic table found ingenious ways to order the chemical elements based on criteria that kept evolving from 1817 onwards. The placement of argon is particularly important for understanding how the development of the periodic table faced difficulties (even perhaps a crisis) and required the collaboration of the scientific community based on argumentation that supported different interpretations. Mendeleev himself was quite concerned with respect to the accommodation of argon and entertained various hypotheses.

7.10  Lewis’s Postulation of the Covalent Bond G.N. Lewis (1916) is generally considered to have presented the first satisfactory model of the covalent (shared pair) bond based on the cubic atom. However, it is important to note that the genesis of the cubic atom can be traced to an unpublished memorandum written by Lewis in 1902 and recounted by him in the following terms: In the year 1902 (while I was attempting to explain to an elementary class in chemistry some of the ideas involved in the periodic law) becoming interested in the new theory of the electron (Thomson’s discovery of the electron in 1897), and combining this idea with those which are implied in the periodic classification, I formed an idea of the inner structure of the atom (model of the cubic atom) which, although it contained crudities, I have ever since regarded as representing essentially the arrangement of the electrons in the atom (reproduced in Lewis 1923, pp. 29–30, emphasis added).

Lewis (1916, p. 768) reproduced the following postulates of his 1902 theory of the cubical atom at length in his 1916 article: 1. In every atom is an essential kernel which remains unaltered in all ordinary chemical changes and which possesses an excess of positive charges corresponding in number to the ordinal number of the group in the periodic table to which the element belongs; 2. The atom is composed of the kernel and an outer atom or shell, which in the case of the neutral atom, contains negative electrons equal in number to the excess of positive charges of the kernel, but the number of electrons in the shell may vary during change between 0 and 18;

7.10  Lewis’s Postulation of the Covalent Bond

145

3. The atom tends to hold an even number of electrons in the shell, and especially to hold eight electrons which are normally arranged symmetrically at the eight corners of a cube; 4. Two atomic shells are mutually interpenetrable; 5. Electrons may ordinarily pass with readiness from one position in the outer shell to another. Nevertheless they are held in position by more or less rigid ­constraints, and these positions and the magnitude of the constraints are determined by the nature of the atom and of such other atoms as are combined with it; and 6. Electric forces between particles which are very close together do not obey the simple law of inverse squares which holds at greater distances. Postulate 3 was the most striking and at the same time controversial feature of Lewis’s theory which led to the postulation of the “octet rule.” Postulate 4 was particularly important as it further facilitated the idea of sharing of electrons. Postulate 5 showed Lewis’s disagreement with Bohr’s model of the atom, and Postulate 6 coincided with new developments in the structure of the atom. In order to understand Lewis model of the covalent bond in retrospect, Kohler (1971), has presented a detailed account of the origin of Lewis’s ideas: When it was first proposed, Lewis’s theory was completely out of tune with established belief. For nearly 20 years it had been almost universally believed that all bonds were formed by the complete transfer of one electron from one atom to another. The paradigm was the ionic bond of Na+ Cl- , and even the bonds in compounds such as methane or hydrogen were believed to be polar, despite their lack of polar properties. From the standpoint of the polar theory the idea that two negative electrons could attract each other or that two atoms could share electrons was absurd (p. 344, italics added).

Rodebush (1928), a chemist reviewing the origin of the covalent bond in the late 1920s, shared the same concern: “Since according to Coulomb’s law two electrons should exert a repulsion for each other, the pairing of electrons seems at first glance to be a bizarre idea. In order to account for the peculiar behavior Lewis assumed the existence of a magnetic attraction between the electrons” (pp. 513–514). This clearly shows that Lewis’s theory of sharing electrons (covalent bond) had to compete with a rival theory, viz. transfer of electrons (ionic bond). From a philosophy of science perspective the rivalry between competing theories (paradigms/research programs) is an integral part of scientific progress (Lakatos (1970, p. 155). At this stage it is important to note that Thomson’s (1897) discovery of the electron and later publications (Thomson 1907) provided powerful arguments for the polar theory of the ionic bond. According to Thomson (1907): “For each valency bond established between two atoms the transference of one corpuscle from the one atom to the other has taken place …” (p. 138). Although Thomson accepted that overlapping of corpuscles could produce a nonpolar bond in theory, he believed that in reality all bonds were polar bonds (p. 131). According to Kohler (1971), although sharing of electrons to form covalent bonds seemed shocking at first, few chemists have shown interest in the origin of the shared pair bond. In the previous sections it has been shown that the cubic atom was important in the development of Lewis’s theory of the shared pair covalent bond.

146

7  Feyerabend’s Counterinduction and Science Textbooks

Lewis’s cubic atom was first conceived as a teaching device to illustrate the octet rule and can be considered as ‘speculative’. Modern philosophy of science has emphasized the importance of speculation as characteristic of scientific progress (Kuhn 1970, Lakatos 1970). For example, Bohr’s (1913) use of the quantum postulate in his theory is considered to be speculative (see above section on Bohr). Interestingly, Lewis (1919) in a letter written to Robert Millikan complained: “ … I could not find a soul sufficiently interested to hear the theory [1902 memorandum]. There was a great deal of research work being done at the university [Harvard], but, as I see it now, the spirit of research was dead”. First dissenting voices against the polar orthodoxy were raised by Bray and Branch (1913), colleagues of Lewis at MIT. Bray and Branch (1913) objected that the polar theory had been extended far beyond its proper limits and suggested, “… there are two distinct types of union between atoms: polar, in which an electron has passed from one atom to the other, and non-polar, in which there is no motion of an electron (p. 1443). According to Kohler (1971), “By 1913 or so the polar theory completely dominated chemistry, and it did until it was replaced by Lewis’s theory in the early 1920s” (p. 355). Later Thomson (1914) himself changed and accepted that all bonds were not polar bonds after all. The role of Lewis’s cubic atom as a theoretical device for understanding the sharing of electrons has been studied by Niaz (2001). Of the 27 general chemistry textbooks (published in U.S.A., between 1960s to 1990s, see Appendix 9) evaluated, only three had a satisfactory presentation and following is an example: Lewis assumed that the number of electrons in the outermost cube on an atom was equal to the number of electrons lost when the atom formed positive ions … he assumed that each neutral atom had one more electron in the outermost cube than the atom immediately preceding it in the periodic table. Finally, he assumed it took eight electrons—an octet—to complete a cube. Once an atom had an octet of electrons in its outermost cube, this cube became part of the core, or kernel, of electrons about which the next cube was built. (Bodner and Pardue 1989, p. 273, italics added).

With this introduction Bodner and Pardue (1989) explain the formation of the covalent bond in the following terms: By 1916, Lewis had realized that there was another way atoms could combine to achieve an octet of valence electrons—they could share electrons. Two fluorine atoms, for example, could share a pair of electrons and thereby form a stable F2 molecule in which each atom had an octet of valence electrons. (p. 274)

The authors provide pictures of the individual cubes of fluorine coming together to share an edge and thus form the covalent bond in which the eight electrons are oriented towards the corners of a cube. Furthermore, the authors reproduce Lewis’s 1902 memo with hand drawings of the cubic atom that was included by Lewis (1923) in his book on valence. Furthermore, it is important to note that Bodner and Pardue (1989) go beyond by recognizing the importance of Lewis’s contribution within a historical perspective, “The magnitude of this achievement is underlined by the fact that this model was generated only five years after Thomson’s discovery of the electron and nine years before Rutherford proposed the model of the atom …”

7.11  Discovery of the Planet Neptune

147

(p. 273). Clearly, in 1902, Lewis did not have all the empirical data and hence the hypothesis of the cubic atom was at best only partly supported by experimental evidence—a clear example of counterinduction, as suggested by Feyerabend. Interestingly, Bodner and Pardue (1989, p.  273, cited above) used the word “assumed” thrice in order to describe what Lewis did. In the same study another textbook tried to understand the difference between ionic and covalent bonds: Covalent bonds also result in a molecule lower in energy than the isolated atoms, but since in a pure covalent bond there is equal sharing of electrons and a symmetric charge distribution, a simple Coulomb’s law calculation cannot explain the energy lowering as it does for ionic bonding. In order to understand many properties of molecules, such as their geometries and the strengths of the chemical bonds that hold the molecules together, we must investigate the nature of the covalent bond (Segal 1989, p. 496).

Interestingly, Lewis (1916) also referred to the same difficulty in Postulate 6 of his 1902 memorandum (cited above). It is not far-fetched to suggest that if the sharing of electrons was considered to be ‘bizarre’ and ‘absurd’ by the scientists (cf. Kohler 1971, p. 363), it could appear counterintuitive to students as well. The controversial origin of the covalent bond and its rivalry with the ionic bond provides a good opportunity to illustrate how progress in science is based on controversy and how established theories or ways of thinking are difficult to change—hence the importance of counterinduction. Thus besides providing students detailed instructions for writing the Lewis structures (almost all textbooks do that), a brief reference to the historical details can facilitate conceptual understanding of the difference between the two types of bonds.

7.11  Discovery of the Planet Neptune Discovery of this planet is a good example for illustrating ‘science in the making’. Neptune was the first planet to be discovered due to evidence that indicated that it was causing a gravitational effect leading to irregularities in the orbit of another planet, Uranus (discovered in 1781 by Friedrich W. Herschel). Thus, scientists predicted the existence of Neptune before it was observed (Grosser, 1962). In 1845, John C. Adams at St. John’s College, Cambridge, estimated the orbit of the unknown planet to be beyond that of Uranus, and predicted that it could account for the irregularities in its motion. Later, Urbain J.J. Leverrier in France made similar calculations in 1846 and communicated them to the French Academy of Sciences and Johann G. Galle in Berlin, who discovered the planet on September 23, 1846. At the time of its discovery Neptune was only 1° from the place predicted by Leverrier and about 2½o from the place predicted by Adams. Interestingly, Adams had communicated his calculations among others to the English astronomer James Challis at Cambridge. Challis undertook to verify the calculations of Adams and Leverrier, especially with respect to the existence of a new planet (for details, see Smart, 1946). Challis sighted the undiscovered planet

148

7  Feyerabend’s Counterinduction and Science Textbooks

(i.e., Neptune) at least four times during the summer of 1846 (once on August 4), that is before Galle. According to philosopher-physical chemist Michael Polanyi (1964), “… these facts made no impression on him [Challis], for he distrusted altogether the hypothesis which he was testing” (p. 30). This clearly shows how lack of a belief in a presupposition (existence of Neptune) led Challis to ignore relevant experimental data. Now let us see how a physicist-philosopher of science has interpreted the discovery of Neptune based on a conjecture: Leverrier and Adams [must have wondered] “Look here, the planet Uranus is not keeping time properly; the only way we can both acknowledge that fact and also save celestial mechanics is to suppose that there is another object, some ‘dark body,’ which has the following properties, a, b, c … etc.” And they worked out the properties of this ‘in reverse,’ as it were. What would have to be the properties of a planet in order to perturb Uranus as it is perturbed? (Hanson, 1964, pp. 166–167).

This constitutes an interesting example of “science in the making” based on counterinduction, namely accepting the existence of a hypothetical planet, Neptune. Early conjectures of Leverrier and Adams, subsequent discovery of Neptune by Galle and the interpretation by Hanson, are all based on the premise that Newton’s physics and especially the law of gravitation correctly described the orbits of the planets. Hanson (1958) pays tribute to the intellectual efforts of Leverrier in the following terms: How remarkable that this man [Leverrier] should have raised classical mechanics to its highest pinnacle by predicting the unseen Neptune as being responsible for observed aberrations in the orbit of Uranus (pp. 203–204).

Lakatos (1970) goes beyond and provides further insight by presenting an imaginary case of planetary misbehavior that elucidates how scientists do science: A physicist of the pre-Einsteinian era takes Newton’s mechanics and his law of gravitation (N), the accepted initial conditions, I, and calculates, with their help, the path of a newly discovered small planet, p. But the planet deviates from the calculated path. Does our Newtonian physicist consider that the deviation was forbidden by Newton’s theory and therefore that, once established, it refutes the theory N? No. He suggests that there must be a hitherto unknown planet p′ which perturbs the path of p. He calculates the mass, orbit, etc., of this hypothetical planet and then asks an experimental astronomer to test this hypothesis. The planet p´ is so small that even the biggest available telescopes cannot possibly observe it: the experimental astronomer applies for a research grant to build yet a bigger one … Were the unknown planet p´ to be discovered, it would be hailed as a new victory of Newtonian science. But it is not. Does our scientist abandon Newton’s theory and his idea of the perturbing planet? No. He suggests that a cloud of cosmic dust hides the planet from us … But the cloud is not found. Is this regarded as a refutation of Newtonian science? No … [and] yet another ingenious auxiliary hypothesis is proposed … (pp. 100–101).

Motterlini (1999, p. 69) considers that the imaginary story of the planet related by Lakatos is based on many real historical instances including the discovery of the Neptune. Indeed, in a sense Lakatos even goes beyond Feyerabend and suggests that scientists can also resort to repeated counterinductions, namely if experimental evidence is not found for one hypothesis others can be postulated. Of course, Lakatos

7.12  Discovery of the Elementary Particle Neutrino

149

refers to these unsupported hypotheses as “auxiliary hypotheses.” This story encapsulates many aspects of ‘science in the making’ and thus has implications for understanding nature of science, as follows: (a) When confronted with empirical evidence that seems to refute a scientific theory, scientists generally resist such a refutation and look for an alternative hypothesis; (b) The alternative hypothesis requires further experimental evidence (mass, orbit, and other characteristics of an unknown planet, for example the work of Adams and Leverrier in the case of Neptune); (c) The process of finding alternative hypotheses and looking for experimental support can continue for some time; (d) The role of these ‘auxiliary hypotheses’ is to protect the guiding assumptions or hard-core of a theory (Newtonian theory in the present case); (e) Eventually, the hard-core of a theory crumbles and a new theoretical framework assumes the role of theory building (Einstein’s general relativity theory, 1915, in the present case). Once again, Lakatos’s presentation approximates to what Feyerabend would consider as counterinduction. This coincidence between Lakatos (a rationalist) and Feyerabend (an anarchist) is all the more significant (see Chap. 2 for details).

7.12  Discovery of the Elementary Particle Neutrino Before 1930 it was generally believed that, based on Einstein’s equation, E = mC2, mass-energy is conserved in nuclear reactions. Based on this assumption, generally referred to as ‘energy conservation’ whenever there is a change of mass in nuclear reactions, the difference shows up as kinetic energy, as indicated by Einstein’s equation. By the end of the 1920s it was found that energy conservation does not seem to hold for beta decay reactions (changing a neutron into a proton and an electron in radioactivity), as about one-third of the energy seems to disappear. To uphold the law of conservation of energy it was postulated that another particle is emitted that carries off the missing energy. This implied the existence of particles called neutrinos, predicted as early as 1929 by W. Pauli, years before they were actually discovered. In other words, postulation of the neutrino constituted a counterinduction, as there was no experimental evidence of their existence. Although the neutrino could not be detected for many years, it became important after Enrico Fermi presented his theory of beta decay in 1933 in which neutrinos (Italian for ‘small neutral one’) are emitted and by 1940 it was used routinely by nuclear theorists (Kragh, 1999). Fermi’s theory identified the weak nuclear force as being distinct from the strong nuclear force and responsible for beta decay. Interestingly, Fermi’s ground breaking theory of beta decay, which founded the modern theory of weak interactions, was first rejected by Nature. Neutrinos are mass less, charge less, and do not feel the strong nuclear force and interact via the very short ranged weak nuclear force. Recent research based on neutrino oscillations, however, has shown that neutrinos may have mass. Actually physicists believed in the existence of the neutrino even though it had not been detected, and for some it was simply a convenient way of organizing experimental data (counterinduction).

150

7  Feyerabend’s Counterinduction and Science Textbooks

Despite the difficulties and a lack of interest in the experimental detection of the neutrino, in 1951 Frederick Reines and Clyde Cowan at Los Alamos started planning experiments. In 1956, using the Savannah River reactor as a neutron source, Reines and Cowan found signals that were considered signs of neutrino-proton reactions (Cowan et al., 1956). Reines shared with Martin Perl the 1995 Nobel Prize for physics (Cowan had died earlier). In his Nobel Prize acceptance speech, thoughtfully entitled, “The neutrino: From poltergeist to particle”, Reines referred to the original idea of Pauli in the following terms: The neutrino of Wolfgang Pauli was postulated in order to account for an apparent loss of energy-momentum in the process of nuclear beta decay. In his famous 1930 letter to the Tübingen congress, he stated: “I admit that my expedient may seem rather improbable from the first, because if neutrons existed they would have been discovered long since. Nevertheless, nothing ventured nothing gained... We should therefore be seriously discussing every path to salvation.” (Reines, 1997, p. 203. Note: When the neutron was discovered by Chadwick in 1932, Fermi renamed Pauli’s particle the “neutrino”).

In June 1956, Reines and Cowan sent a telegram to the man who started it all (Pauli) informing that they had definitely detected neutrinos from fission fragments by observing inverse beta decay of protons. Pauli’s response was prophetic indeed and shows yet another facet of science in the making, “Everything comes to him who knows how to wait, Pauli” (Reproduced in Reines, 1997, p. 215). According to Hanson (1958), “The neutrino idea, like those of other atomic particles, is a retroductive conceptual construction out of what we observe in the large” (p. 124, emphasis added). Considering the immense efforts required to detect the neutrino, Kuhn (1970) concluded: “… no experiment can be conceived without some sort of theory, the scientist in crisis will constantly try to generate speculative theories that, if successful, may disclose the road to a new paradigm …” (p. 87, italics added). Hanson’s retroductive conceptual construction and Kuhn’s speculative theories, can easily be construed as Feyerabend’s counterinduction—accepting hypotheses (neutrino) for which there is no experimental evidence. In contrast to the traditional textbook science, these two episodes from science in the making (Neptune and Neutrino) clearly show that scientists generally resist the refutation of a theory by putting up alternatives and that besides the experimental apparatus a scientist is almost always accompanied by his presuppositions and counterinduction that provide guidance in the face of difficulties.

7.13  Discovery of the Tau Lepton Martin L. Perl was the recipient of the 1995 Nobel Prize in physics for his discovery of the Tau Lepton, based on a 16 year history (1963–1979), when all experimental measurements agreed with the hypothesis that the Tau was a lepton produced by a known electro-magnetic interaction (Perl et al., 1975; Perl, 1997). Perl’s interest in lepton physics started with his dissatisfaction with the strong-interaction physics of the 1960s and from a desire to discover fundamental facts in particle physics (Perl,

7.13  Discovery of the Tau Lepton

151

2004, p. 405). This led him to do experiments with the electron and the muon, elementary particles that do not participate in strong interactions, but rather in the electromagnetic, weak, and gravitational interactions. There were two puzzles about the properties of the electron (e) and the muon (μ) in the 1960s: (1) Why is muon 208 times heavier than the electron, when their properties with respect to particle interaction are the same? and (2) As compared to the electron, why is the muon unstable (lifetime = 2.2 × 10−6 s)? On arriving at the Stanford Linear Accelerator Center (SLAC) in 1963, Perl proposed to look for the unknown differences between the electron and the muon, and speculated that there might be more leptons similar to the electron and the muon, namely unknown heavier charged particles: “I dreamed that if we could find a new lepton, the properties of the new lepton might teach us the secret of the electron-muon puzzle” (Perl, 2004, p. 407, italics added). Early experiments using photon beams and later inelastic scattering of muons on protons, did not provide the expected results. This led Perl to conclude: Experimental science is a craft and an art, and part of the art is knowing when to end a fruitless experiment. There is a danger of becoming obsessed with an experiment even if it goes nowhere … I avoided obsession and gave up on the scattering experiment. That turned out to be a good decision because modern experiments have shown that these scattering experiments do not illuminate any differences between the electron and the muon beyond the mass difference (Perl, 2004, pp. 408–409, original italics).

After the lack of success of the early experiments, Perl considered the possibility of the existence of charged leptons more massive than the electron and muon, called heavy leptons. By the late 1960s the concept of a heavy lepton (L) along with its decay modes was the subject of speculation in the research literature. This led Perl to conclude: This brings me to the question I raised in the Introduction: is there more and broader speculation these days in particle-physics theory than forty years ago? Judging by the various ideas of forty years ago about possible types of leptons, we were rather timid about speculations. There was a fear of being thought unsound. There was reluctance to stray too far from what was known. Today the only limit to theoretical speculations about particle physics is that the mathematics be correct and that there be no obvious conflict with measured properties of particles and reactions. One example is string theory with all its different forms and extensions. Another example is the recent, tremendous amount of interweaving of particle physics with astrophysics and cosmology. I think this is a good change. It can stimulate the experimenter to go in new directions, but the experimenter must be cautious as to how she or he uses time. If possibly relevant data have already been produced, then it may be relatively fast to test the speculation against the data. But if a new experiment must be built to test the speculation, that is another story (Perl, 2004, pp. 409–410, original italics).

After surmounting difficulties with funding, the construction of SPEAR electron-­ positron collider at SLAC started at the end of the 1960s, along with collaboration with the Lawrence Berkeley Laboratory (LBL) group. The purpose of such detectors (SLAC-LBL) is to detect the type and vector momentum of most of the particles (e.g., electrons, photons, muons, protons, etc.) coming from a reaction taking place at the center of the detector. This led Perl to conclude: These detectors are necessary to obtain the complicated and often subtle data in modern experiments. But large detectors have come at a human cost. It is no longer possible for a

152

7  Feyerabend’s Counterinduction and Science Textbooks

few people to build and operate such detectors. Hence there are often hundreds of experimenters in a typical group and the new very large and complicated detectors require groups with more than a thousand members. Of course such detectors produce tremendous amounts of data (Perl, 2004, pp. 413–414, original italics).

In 1971, Perl and colleagues at SLAC-LBL submitted a research proposal in which the heavy-lepton search (Perl’s major interest), “… was left for the last and allotted just three pages, because to most others it seemed a remote dream. But the three pages did contain the essential idea of searching for heavy leptons using electron-­ muon events” (Perl, 2004, p.  414, italics added). Perl’s colleagues thought that including more about heavy leptons and the electron-muon problem would unbalance the proposal. This did not deter Perl, and he persevered by submitting a 10-page ‘Supplement’ which emphasized that the detector could simultaneously collect relevant data with respect to the following research questions (reproduced in Perl, 2004, p. 415): (1) Are there charged leptons with masses greater than that of the muon? and (2) Are there anomalous interactions between the charged leptons and the hadrons? These questions were, of course based on Perl’s basic strategy of studying muon-proton inelastic scattering, and background provided here is important to understand later developments. By June 1975, Perl and his group had ‘good’ experimental evidence for electron-­ muon events, from the SLAC-LBL detector, which were presented at the Canadian Institute of Particle Physics, McGill University, Montreal. Furthermore, they discussed possible sources of the electron-muon events: heavy leptons, heavy mesons, or intermediate bosons. Of the 126 particle-pair events reported, at least 24 could be attributed to electron-muon events, which was the strongest evidence at that time for the Tau Lepton. Still, Perl was cautious and concluded, “I was still not able to specify the source of the electron-muon events: was the source new leptons, new mesons, or new bosons? But I remember feeling strongly that the source was heavy leptons. It would take two more years to prove that” (Perl, 2004, p. 417). This clearly shows Perl’s strong belief (presupposition, guiding assumption) with respect to the Tau Lepton, in the context of later developments. Based on further experimental evidence and refinement of the procedures, Perl and colleagues finally published in December 1975, their crucial article, which concluded: We conclude that the signature e–μ events cannot be explained either by the production and decay of any presently known particles or as coming from any of the well-understood interactions which can conventionally lead to an e and a μ in the final state. A possible explanation for these events is the production and decay of a pair of new particles, each having a mass in the range of 1.6–2.0 GeV/c2. (Perl, et al., 1975, p. 1492).

Perl and colleagues were, however, not yet prepared to claim that they had found a new charged lepton, and in order to accentuate their uncertainty they denoted the new particle by U for ‘unknown’ in some of their 1975–1977 papers. Despite the uncertainty, and the need to convince the world that the electron-muon events were significant (caused by the decay of a pair of heavy leptons), Perl was still optimistic, “Thus in 1975, twelve years after we began our lepton-physics studies at SLAC,

7.13  Discovery of the Tau Lepton

153

these studies bore fruit” (Perl, 2004, p. 418). Besides his research program, guiding assumptions, new technology (electron-positron collider), Perl attributed the success to: I had smart, resourceful, and patient research companions. I think these are the elements that should be present in speculative experimental work: a broad general plan, specific research methods, new technology, and first-class research companions. Of course, the element of luck will in the end be dominant (Perl, 2004, pp. 418–419, original italics, underline added).

Despite publication of the experimental results (Perl et al., 1975), there was still confusion and uncertainty about validity of the data and its interpretation, namely acceptance of an unsupported hypothesis—counterinduction. One of the impediments to the acceptance of the tau as the third charged lepton (hypothesis) was that no other laboratory had reported similar results. Looking back over those years, Perl (2004) reported: “It was a difficult time. Rumors kept arriving of definitive evidence against the tau: electron-muon events not seen, the expected decay modes not seen, theoretical problems with momentum spectra or angular distribution” (pp. 420–421). Despite the recognition that they were working on a hypothesis that needed further experimental evidence, Perl and colleagues persevered. In 1976, however, things started to change when outside experimental confirmation for anomalous muon events was reported by Cavalli-Sforza et al. (1976). The second confirmation came from Perl’s own group at the SLAC-LBL detector, providing support for the decay of a heavy lepton (Feldman & Perl, 1977). The third confirmation came from an experiment at the DORIS electron-positron storage ring in Hamburg (PLUTO-­ Collaboration, et al., 1977). These events convinced Perl that he was right about the existence of the tau as a sequential heavy lepton. If the tau was indeed a sequential heavy lepton, two substantial semi-leptonic decay modes had to exist. One of the problems in finding these decay modes (years 1977–1979) was the poor efficiency for photon detection in the early detectors. To resolve this problem new detectors (such as Mark II) were put into operation at SLAC-LBL facilities. Given this new efficiency of the detectors, the decay mode was detected and measured by various groups of researchers in different parts of the world, including Perl’s own group (Abrams & Perl, 1979). Finally, after all the difficulties that involved improved experimental equipment and interpretations, the stage was set for Perl (2004) to conclude: By the end of 1979, all confirmed measurements agreed with the hypothesis that the tau is a lepton produced by a known electromagnetic interaction and that, at least in its main modes, it decays through the conventional weak interaction. So ends the sixteen year history, 1963 to 1979, of the discovery of the tau lepton and the verification of that discovery (p. 424).

It is important to note that for almost 16  years, the hypothesis that the Tau is a Lepton remained unsupported (lacked experimental evidence) and still the scientists continued to work on it—a clear example of counterinduction. Interestingly, just as Perl’s discovery of the Tau Lepton solved the puzzles related to the nature of the electron and the muon, it created additional puzzles with respect to the Tau itself.

154

7  Feyerabend’s Counterinduction and Science Textbooks

Perl (2004) considers this as an irony of his discovery and suggested the following as a possible explanation: “The discovery of the last decade that one type of neutrino can change into another type of neutrino may lead eventually to an understanding of what constitutes the intrinsic nature of elementary particles. But we are not there yet” (p. 424). This shows how the resolution of one puzzle leads to another and consequently the need for further research. Kragh (1999) has referred to the additional puzzles produced by Perl’s discovery of the Tau in the following terms: “The accompanying tau neutrino was assumed, rather than detected. The acceptance of the tau lepton (or tauon) implied that the symmetry between quarks and leptons—or between strong and weak interactions—was no longer satisfied. Two new quark flavors were needed and, as we shall see shortly, they were later found” (p. 343). Even in present day experimental particle physics, Perl speculated (even dreamed) that other leptons (heavy charge particles) existed. A constant interaction between Perl’s presuppositions and experimental verification of the Tau Lepton over a period of almost 16 years, clearly shows the importance of counterinduction. Interestingly, on receiving the Nobel Prize for Physics in 1995, Martin Perl started his acceptance speech with the following words: “My first thoughts in writing this lecture are about the young women and young men who are beginning their lives in science: students and those beginning scientific research” (Perl, 1997, p. 168). Indeed, Perl’s research experience extending over 50 years can have important implications for teaching science. Next, in his speech he referred to human vicissitudes involved in doing experimental work: (a) Pleasure, when an experiment is completed and the data safely recorded; (b) Anxiety, when an experiment does not work well or breaks; (c) Pain, when an experiment fails or when an experimenter does something stupid; (d) In the discovery of the Tau Lepton, the ups and downs of his emotions extended over many years; (e) There is nothing wrong with an experimental idea as long as you are the first to use it; (f) For a speculative experimenter it is good to have two experiments going, one in operation and the other being built; (g) Avoid obsession with a particular experiment; (h) In order to understand physics, look for simplicity; (i) Good speculative experimental work requires: a broad general plan, specific research methods, new technology, first-class research companions and some degree of luck; (j) Interpreting experimental data is difficult and leads to uncertainties; and (k) Role of creativity and the freedom to try out new ideas. Most undergraduate and graduate school programs are overloaded with course and lab work. Perl (2007) is critical of such curriculum practice as it does not motivate the students to do creative work: “There is little time for the student to play with ideas, to dream about discoveries and inventions” (p.  5). How many of our students would be willing to follow Perl’s advice on cutting-edge experimental work. Needless to say, Perl clearly brings to the forefront the idea that ‘science in the making’ is a human enterprise (for further details, see Niaz, 2012, Chapter 7). This chapter provides a wide range of episodes from the history of science that can be considered as examples of counterinduction as understood by Feyerabend. It provides a glimpse of how science works (science in the making) and enriches traditional philosophy of science. Conclusions based on these findings along with those of Chapters 3, 4, 5 and 6 will be presented in Chap. 8.

Chapter 8

Conclusion: Feyerabend and Challenges of the Twenty-First Century

In both philosophy of science and science education, Feyerabend is generally considered to be against rationalism, anti-science and for having espoused anything goes. Based on material presented in the previous chapters, here I will attempt to show that this image is erroneous and that on the contrary Feyerabend was presenting a picture of science that represented “how science really works.”

8.1  Feyerabend’s Hyperbolic Flourishes At the end of his fourth lecture at the Faculty of Sociology, University of Trent, Italy in May 1992 (about 2 years before his death), someone from the audience asked Feyerabend to clarify his position with respect to his controversial thesis of “Anything goes”. Feyerabend provided the following scenario: To be situated on a solid Earth is a first and basic experience. However, Anaximander said that the Earth floats in the midst of air. Just think, it is surprising as nothing floats in air. Still, Anaximander insists that the heavy Earth floats in air. According to modern criteria this was certainly anarchism, as it developed and led to something. Thus, “Anything goes only means do not put a limit on your imagination” (Feyerabend, 1999a, p. 157, Spanish edition of the Trent Lectures). In his Killing Time, Feyerabend (1995) clarified that he was not against rationalism per se, but only against those forms that were rigid and pompous. J.  Agassi (2014) in his Popper and his popular critics: Thomas Kuhn, Paul Feyerabend and Imre Lakatos, has highlighted the fact that some of Feyerabend’s anti-rationalistic and hyperbolic flourishes (e.g., science is not superior to magic) were primarily teasers or challenges to defend diversity, and Feyerabend acknowledged this aspect in his correspondence with Agassi. Ben-Ari (2005) has suggested that Feyerabend’s (1975a/1993, p. 14), anything goes, is helpful when it comes to choosing pedagogic methods that will improve © Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3_8

155

156

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

learning. Anything goes does not mean that all knowledge is equally good. On the contrary, it means that the scientist is free to employ all methodological approaches that seem fruitful, with the understanding that some of the attempts may not succeed (Quale, 2007). According to Hodson (1992, p.  131) anything goes implies the absence of an algorithm, rather than the absence of methods and should not be taken too literally. Anything goes also means that the methodological strategy is not decided a priori (Kalman, 2009b). Perhaps, for a naïve realist it is difficult to understand anything goes, especially if she/he is looking for “absolute truths.” Recent research in philosophy of science has endorsed Feyerabend’s vision of science for the twenty-first century: “Certainly his [Feyerabend’s] broad vision of the scientific enterprise as a pluralistic, value-laden, socially structured, politically invested project sits comfortably alongside many current trends in philosophy of science—a view which, though ‘anarchistic’ forty years ago, is, today, increasingly recognized as a vision of science that is fit to meet the practical, epistemic, and socially challenges of the twenty-first century” (Brown & Kidd, 2016, p. 7).

8.2  Feyerabend’s Epistemological Anarchism Table 8.1 provides an overview of the classification of all the articles evaluated in this book. Following are some of the salient features of the results obtained: (a) Science & Education was the only journal in which around 10% of the articles were classified in Level V, which approximates to an understanding of Feyerabend as a philosopher of science trying to investigate how science really works; (b) In all the chapters most of the articles were classified in Levels II and III; (c) In all the chapters some articles were classified in Level I (none for HPST), which approximates to the traditional view of Feyerabend as a relativist and postmodern philosopher of science; (d) Articles classified in Level III recognized the problematic nature of understanding Feyerabend and hence the need for alternative interpretations; and (e) Of the 120 articles evaluated in this book, only 9% approximated to an u­ nderstanding Table 8.1  Comparison of the levels of classification of articles in different chapters of this book Number of articles in level Chapter (Journal) n 3 (S&E) 78 4 (JRST) 21 5 (Interchange) 15 6 (HPST) 6

I 9 2 1 –

II 29 10 – –

III 22 5 11 6

IV 11 2 1 –

V 7 2 2 –

Notes: 1. For a description of Levels I – V, see Chap. 3 2. n: Total number of articles evaluated 3. S&E: Science & Education 4. JRST: Journal of Research in Science Teaching 5. HPST: International Handbook of Research in History, Philosophy & Science Teaching

8.2  Feyerabend’s Epistemological Anarchism

157

of Feyerabend as promoting a plurality of perspectives, which shows the need for a reappraisal of Feyerabend’s contribution to philosophy of science. Following are some aspects for facilitating an understanding of Feyerabend’s epistemological anarchism based on articles evaluated in S&E, JRST, Interchange and HPST. Furthermore, it includes Feyerabend’s thesis of counterinduction and its importance for science textbooks and teaching strategies (Chap. 7). It is plausible to suggest that these findings have implications for science education which are synthesized and discussed in the following sections (based on Chaps. 3, 4, 5, 6 and 7 and presented in alphabetical order): 1. Counterinduction (Accepting unsupported hypotheses) 2. Current view of a science may soon be voted out of office 3. Does science always provide the one “correct” model (theory) 4. Epistemological anarchism: Lifting the lid off the Pandora’s Box? 5. History of a science becomes an inseparable part of the science itself 6. Inferring objectivity from empirical approaches 7. Methodological pluralism: Diversity of methods 8. Nature of science 9. Role of genius in science 10. Scientific expertise needs a critical appraisal 11. Scientific method: Stockpiling and ordering of observations 12. Skeletons in the Newtonian cupboard 13. The new grew out of the old 14. Unnatural nature of science 15. Was Feyerabend a postmodern or a perspectival realist?

8.2.1  Counterinduction (Accepting Unsupported Hypotheses) In the late nineteenth and early twentieth century, atomic theory was still being questioned by some leading physicists (e.g., P. Duhem & W. Ostwald) and consequently Brownian motion could not be explained by the kinetic-molecular theory, or in other words support could be accepted provided we considered it as an example of counterinduction. Starting in 1857 based on the work of Clausius, Maxwell and other theorists, kinetic theory of gases developed through simplifying assumptions (idealization) and in most cases these were speculative with no experimental evidence, leading to counterinduction. The Michelson-Morley experiment was first conducted in 1887 and provided a “null” result with respect to the ether-drift hypothesis, namely, that there was no observable velocity of the earth with respect to the ether (Michelson & Morley, 1887). In this context, with this background it is interesting to reconsider the following two interpretations: (a) According to Lakatos, it took 25 years for the ether-drift hypothesis to be refuted; and (b) According to Feyerabend, Einstein’s special theory

158

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

of relativity (STR) was retained despite empirical evidence (Michelson, Miller) and including Lorentz contraction, to the contrary. Lakatos implies that empirical evidence was necessary for refuting the ether-drift hypothesis, whereas Feyerabend would imply that despite empirical evidence to the contrary, STR was not refuted. Does this have educational implications (especially for writing textbooks)? In other words, a review of the literature at present shows that researchers were interested in finding: Michelson-Morley (MM) experiment led Einstein to postulate his special theory of relativity (STR). With Feyerabend’s perspective of counterinduction it would be interesting to find if science textbooks explore: Despite empirical evidence to the contrary, STR was not refuted. To the best of my knowledge, research with this Feyerabendian perspective has not been conducted. The determination of the elementary electrical charge (e), based on the oil drop experiment aroused considerable interest in the scientific community and controversy between both Robert Millikan and Felix Ehrenhaft. Based on a historical reconstruction it is plausible to suggest that: (a) Ehrenhaft allowed his theory (subelectrons) to be dictated by experimental data and hence in a sense followed the scientific method (accepted all the data); (b) Millikan discarded data that did not support his guiding assumption—this coincides with Feyerabend’s (1975a) claim that scientific theories are not consistent with all the experimental data (p. 43); and (c) Millikan supported a theory that was not supported by at least 59% of his data, and hence accepted a theory that was at least partially unsupported, and Feyerabend would consider this as counterinduction. During the first decade of the twentieth century, both J.J.  Thomson and E.  Rutherford conducted alpha particle scattering experiments in their respective laboratories. Although, results from both laboratories were similar, interpretations of Thomson and Rutherford were entirely different. Thomson propounded the hypothesis of compound scattering, according to which a large angle deflection of an alpha particle resulted from successive collisions between the alpha particle and the positive charges distributed throughout the atom. Rutherford, in contrast, propounded the hypothesis of single scattering, according to which a large angle deflection resulted from a single collision between the alpha particle and the massive positive charge in the nucleus. The rivalry led to a bitter dispute between the proponents of the two hypotheses. Given, Thomson’s credentials (a world master in the design of atomic models) most scientists could argue that Rutherford’s hypothesis of single scattering, perhaps constituted Feyerabend’s counterinduction. Historical events, however, turned out to be otherwise and the scientific community eventually accepted Rutherford’s arguments and supported his hypothesis of single scattering. It is even plausible to suggest that Thomson’s hypothesis of compound scattering provided counterinduction and thus helped Rutherford to strengthen his arguments (a role foreseen by Feyerabend). N. Bohr did not have experimental evidence for all the postulates of his model of the atom in 1913. Bohr followed a different way of understanding Rutherford’s nuclear atom, namely introduction of Planck’s quantum of action that would constitute a bold new hypothesis accounting for more observations, such as stability of Rutherford nuclear atom and the hydrogen line spectra (Balmer, Paschen etc.). It is

8.2  Feyerabend’s Epistemological Anarchism

159

plausible to suggest that this strategy comes quite close to Feyerabend’s counterinduction. Most critics of Bohr’s model (such as Lorentz, Rayleigh, Rutherford, & Stern) would implicitly or explicitly reject counterinduction. The one who came closest to Feyerabend’s hypothesis was Lakatos, who suggested that Bohr’s model was based on an inconsistent foundation, thus endorsing counterinduction. Millikan (1916) described succinctly the photoelectric effect, its explanation by Einstein, experimental determination of Planck’s constant (h) based on Einstein’s photoelectric equation, the controversy surrounding the photoelectric effect and the classical wave theory of light. Interestingly, many scientists both in the U.S.A., and Europe shared Millikan’s thoughts on the subject, which is the fact that Millikan referred to—energy of the ejected electrons in photoelectric effect depends on the frequency of light (Einstein’s hypothesis) and not on the intensity of light (classical wave theory). Despite this acknowledgment, based on his presupposition, Millikan considered Einstein’s hypothesis as reckless—a clear example of counterinduction, viz., accepting an unsupported hypothesis (wave theory), as Einstein’s hypothesis did not explain interference, diffraction and other phenomena. Of course, it can be argued that the wave theory explained some phenomena (e.g., interference, diffraction) that Einstein’s lightquantum could not. It is plausible to suggest that the inclusion of the following aspects related to the photoelectric effect can facilitate a better understanding of the dynamics of scientific progress: (a) Millikan considered Einstein’s hypothesis as reckless—in other words accepting Einstein’s photoelectric equation constituted a clear example of counterinduction, viz., accepting unsupported hypothesis; (b) Einstein’s hypothesis was not accepted by the scientific community, including Planck, the ‘originator’ of the quantum hypothesis, for many years; (c) Millikan presented experimental evidence to support Einstein’s photoelectric equation and still rejected his quantum hypothesis; (d) scientific theories are underdetermined by experimental evidence, that is, no amount of experimental evidence can provide conclusive proof for a theory; (e) scientists customarily have prior theoretical beliefs or presuppositions before doing an experiment, and they resist any change in those epistemological beliefs; (f) Holton (1999) was right (based on general chemistry and physics textbooks) that although at present we may consider Millikan’s (1916) paper an experimental proof of the quantum theory of light, it was not considered as such by Millikan himself. The origin of wave-particle duality can be traced to Einstein’s (1905) hypothesis of the light quantum to explain the photoelectric effect. Later (starting in 1922), this interest in the properties of quanta motivated De Broglie’s search for a theory that would unify the wave and particle aspects. De Broglie’s contribution posed/pondered the question: if light can have both wave and particle properties then why particles of matter (for example, electrons) cannot also have both properties. Despite Einstein’s prestige, authority and support of De Broglie’s hypothesis, duality remained a controversial hypothesis, until conclusive experimental evidence was presented by Davisson and Germer (1927), based on their work with nickel crystals. How scientists went about resolving this dilemma, follows Feyerabend’s perspective, who would endorse that science involves not only “hypothesizing” but also “pondering” that leads to “counterinduction.”

160

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

A historical reconstruction shows that the discovery of argon in 1894, led to: a controversy with respect to its place in the periodic table ➔Mendeleev was reluctant to accept argon as an element ➔ Mendeleev even presented an alternative hypothesis, namely argon as triatomic nitrogen: counterinduction (based on unsupported hypothesis) ➔ Discussions and consultations among members of the scientific community finally led to the placing of argon in a separate group between the halogens and the alkali metals. Mendeleev’s periodic law (ascending order of atomic weights), in some instances, was not supported by the experimental data (e.g., atomic weights of Te and I), and consequently he switched their order and in so doing violated his own periodic law and thus accepted an unsupported hypothesis— counterinduction. In the context of the dilemma faced by one of the teacher-student in the study by Mugaloglu (2014, chap. 3), it is plausible to suggest that Feyerabend’s perspective (counter induction, see Chaps. 2 and 7) leading to diversity and epistemological anarchism approximates to a better understanding of knowledge and belief, while teaching both evolution and intelligent design. The latter, of course, would be considered as having no or little empirical evidence. In 1916 when G.N.  Lewis first presented his theory of covalent bond (shared pair) it was completely “out of tune” with established belief, that is the ionic bond based on transfer of electrons. Furthermore, the idea of sharing electrons was considered to be not only “absurd” but also “bizarre.” Indeed, some of Lewis’s ideas, such as the “cubic atom” and the “octet rule” were speculative. The hypothesis of cubic atom was at best partly supported by experimental evidence and consequently constituted counterinduction. Interestingly, Feyerabend emphasized the role of counterinduction as especially important when the previous theory was firmly established (in the present case Thomson’s theory of the ionic bond).

8.2.2  C  urrent View of a Science May Soon Be Voted Out of Office Feyerabend (2011) in his Tyranny of Science had complained that students are given a false picture of science, in so far as they are not told that the current view of science may soon be voted out of office (p. 125). Rowbottom (2013), while reviewing the Tyranny of Science, considers that, on the contrary science education needs to emphasize the Kuhnian conceptualization of “normal science” based on a dogmatic hegemony of only one paradigm. Implementation of normal science has been the subject of considerable controversy in science education (cf. Siegel, 1979; Niaz, 2010). Actually, instead of presenting a false picture of science Feyerabend’s advice leads to understanding how science really works, namely the tentative nature of science (for a recent appraisal, see Abd-El-Khalick, Belarmino, et al., 2017; Abd-El-­ Khalick, Myers, et al., 2017). More recently, a philosopher of science has presented a scenario in which in the late nineteenth century many Newtonian physicists must have thought that they had discovered the real structure of the physical world, and

8.2  Feyerabend’s Epistemological Anarchism

161

must have been taken by surprise to contemplate how just a few years later (1905 and onwards), the current view of science was voted out of office by the postulation of relativity theory and quantum mechanics (cf. Giere, 2006a, 2006b). This helps to understand Feyerabend’s advice to students. At this stage it is interesting to consider an in-service science teacher’s response who was asked to compare the contributions of Newton and Einstein, after taking a course in the history and philosophy of science (HPS): Newton’s laws are not false. I believe that at that time in history his laws revolutionized ways of thinking and helped to explain many phenomena that could not be understood earlier. Similarly, he was objective as his laws helped to solve many problems … In the quest to understand further, Einstein based on the theory of relativity questioned Newton’s laws. Now, in the 21st century someone with a new theory would do the same to what Einstein did to Newton’s theory. The ‘truth’ has no end and it is ‘true’ until someone discovers a new theory that is accepted by the scientific community (Participant #9, italics added, reproduced in Niaz, 2016, p. 66).

Let us compare what Feyerabend had suggested, namely, “current view of science may soon be voted out of office”, with what the participant wrote, “Now, in the 21st century someone with a new theory would do the same to what Einstein did to Newton’s theory.” Perhaps, there could be no better way of understanding Feyerabend’s advice, which was not a subject of discussion in class. How did this participant come to have this understanding? One plausible answer is that the overall HPS based classroom discussions provided the opportunity and stimulus to think beyond the traditional science curriculum. At this stage, it is important to consider how a philosopher of science has endorsed a similar thesis much more forcefully (cf. Worrall, 2010, p. 288). It seems that Participant #9 (cited above) had perhaps previously read Worrall—that possibility, however, is not even remotely probable. Winchester (1993) has expressed some concern with respect to introducing, history and philosophy of science (HPS) to science teachers and science educators in general. It is plausible to suggest that courses on HPS can provide the impetus and background for developing an understanding of “how science works.” Of course, for such courses to be successful it is essential that a critical appraisal of alternative frameworks of HPS be included (cf. Niaz, 2004, see Chapter 5). I am sure Winchester would agree that cultivating such ideas and thinking is what science education is all about. Consequently, contrary to what Winchester suggested, questioning the presuppositions (predominately empiricist) of our students would not “kill” science but rather facilitate an understanding of how science works. Interestingly, even in his early work Feyerabend (1968) questioned methodological rules and instead emphasized the practice of science based on thoroughgoing pluralism (e.g., Newton, Einstein, what next) in order to check dogmatism. Despite the reform efforts, science education in most parts of the world continues to be dogmatic, especially with respect to its empiricist foundations. Similarly, according to Brush (1989) traditional science teaching looks for objective facts which can be helpful in learning science. However, it has also been recognized that scientific research does not provide immutable truths but rather working hypotheses useful for future research. Brush, then goes on to argue that science educators need

162

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

to strike a balance between these two extremes—a thesis that comes quite close to Feyerabend’s advice that current view of science may soon be voted out of office.

8.2.3  D  oes Science Always Provide the One “Correct” Model (Theory) Most chemists and general chemistry textbooks use one of the three acid-base models (Arrhenius, Brønsted-Lowry and Lewis) to understand different aspects of acid-­ base equilibria (Kousathana, Demerouti & Tsaparlis, 2005). The legacy of logical positivism, would of course, require that the topic of acid-base equilibria should preferably facilitate students’ understanding that of the three models only “one” is “correct” or “true”. In contrast, a general chemistry textbook has suggested that there is no single “correct” definition of acids and bases (Tro, 2008). Similarly, chemists use both the Valence bond (VB) and Molecular orbital (MO) theories in understanding covalent bonding. Although both theories were first proposed in the 1930s, the rivalry between the two has continued up to recent days, which has been recognized by Hoffmann, Shaik and Hiberty (2003) in cogent terms: Taken together, MO and VB theories constitute not an arsenal, but a tool kit, simple gifts from the mind to the hands of chemists. Insistence on a journey through the perfervid bounty of modern chemistry equipped with one set of tools and not the other puts one at a disadvantage. Discarding any one of the two theories undermines the intellectual heritage of chemistry (p. 755).

This clearly shows that the postulation and use of more than one “correct” theory to understand a phenomena can even be an asset and part of the intellectual heritage of chemistry. One general chemistry textbook (see Appendix 9) reiterated the same point with a slightly different emphasis: “Whenever two different theories are used to explain the same concept, the question comes up: Which theory is better? The question isn’t easy to answer, because it depends on what is meant by “better.” Valence bond theory is better because of its simplicity, but the MO theory is better because of its accuracy. Best of all, though, is a blend of the two theories that combines the strengths of both” (McMurry & Fay, 2001, p. 285). Copenhagen and Bohm’s theories of quantum mechanics, provide yet another example of two theories that can be considered as “correct” at the same time (Cushing, 1998). In a critical review of the literature, a physicist has conceded that, “At the turn of the century, it is probably fair to say that we are no longer sure that the Copenhagen interpretation is the only possible consistent attitude for physicists … Alternative points of view are considered as perfectly consistent: theories including additional variables (or ‘hidden variables’)” (Laloë, 2001, p. 656; hidden variables refers to Bohm’s theory). At this stage it is plausible to suggest that science can provide more than one “correct” theory of the same phenomena and that would represent what Feyerabend referred to as epistemological pluralism. Interestingly, according to van Strien

8.2  Feyerabend’s Epistemological Anarchism

163

(2019) Feyerabend developed his arguments for pluralism in science (also anarchism) in the context of debates on David Bohm’s alternative approach (hidden variables) to quantum physics. Bohm and Feyerabend were colleagues at the University of Bristol during 1957–58. Feyerabend was particularly concerned to see how the community of quantum physicists (Bohr and colleagues) dogmatically excluded alternative/rival approaches.

8.2.4  E  pistemological Anarchism: Lifting the Lid Off the Pandora’s Box? Epistemological anarchism basically refers to: “How science really works”—anything goes means that a scientist employs all sorts of methodological strategies to see what works. It is important to note that the methodological strategy is not decided a priori (cf. Kalman, 2009b, Chapter 3). There is some concern among science educators that Feyerabend was anti-science (Matthews, 2009). Recent research in philosophy of science, however, has shown that this was not the case and that Feyerabend’s major objective was to understand, “how science really works” (for details see Chap. 2). Winchester (1989) has referred to a deep rooted problem in both science and science teaching, namely accumulating “findings” based on two centuries of empiricism. Despite efforts by philosophers of science (Popper, Whitehead, Kuhn and Feyerabend), this is deeply misleading for science students as it ignores what actually goes on, and hence the need to understand how science really works. The idea of two centuries of empiricism is of course, laden with Baconian methods and Hodson (2014) has presented a dilemma for science educators by juxtaposing the traditional science curriculum and Feyerabend’s epistemological anarchism and thus opening the possibility of lifting the lid off the Pandora’s Box. Furthermore, Hodson (2014) wonders if we embrace anarchism would it be possible to retain what is still good and useful in the traditional view of science, namely conceptual clarity and stringent testing. On the contrary a deeper understanding of Feyerabend’s philosophy of science shows that it does not rule out stringent testing. For example, the strategy of counterinduction can lead to the consideration of hypotheses with even no empirical support and thus precisely facilitate conceptual clarity (for details see Chap. 7).

8.2.5  H  istory of a Science Becomes an Inseparable Part of the Science Itself In the context of the Lavoisier-Priestley controversy, de Berg (2014) has endorsed Feyerabend’s (1993, p. 21) thesis that history of a science becomes an inseparable part of the science itself. The controversy between Priestley’s phlogiston theory and

164

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

Lavoisier’s oxygen theory has been the subject of considerable discussion in the history and philosophy of science literature (cf. De Berg, 2014). A historical reconstruction shows the role played by experiments, contradictions and their respective worldviews (theories). Priestley’s adherence to the phlogiston theory is well known. However, it is not so well known that in reporting his results Lavoisier was guided by the law of conservation of mass. Next, the question arises, whether Lavoisier formulated or assumed the law. It seems, that after almost 100 years, it is now possible to understand that Lavoisier could not formulate the law inductively and consequently followed in the footsteps of Newton and assumed the law. De Berg (2014) has formulated a series of questions that can help students to understand the historical controversy and following is one example: Some phlogistonists suggested that phlogiston might actually carry negative weight and increase the buoyancy of a metal in air. Suggest why they may have come to this conclusion. Do you think their suggestion was feasible? (p. 2063). De Berg (2014, p. 2066) considers such questions as an example of embedding history in a controversy, and thus facilitate students’ interest and understanding. In order to understand the Priestley-Lavoisier controversy, Koertge (1996) has postulated the need for understanding the “clash” between two comprehensive theoretical systems, namely Lavoisier’s oxygen theory and Priestley’s phlogiston theory. The latter being a comprehensive theory treated various other subjects successfully and thus could not be refuted easily. Koertge suggests that in order to understand the controversy students must first know the “empirical comprehensiveness and the metaphysical attractiveness of phlogiston chemistry” (p. 399). The “clash” between Lavoisier’s oxygen theory and a scientifically unacceptable theory (Priestley’s phlogiston theory) today, approximates to Feyerabend’s counterinduction. In order to understand the controversy successful students must become “knowledgable phlogistonists.” The two examples (presented above), related to the Priestley-Lavoisier controversy, clearly show the importance of Feyerabend’s (1993) advice that history of a science becomes an inseparable part of the science itself (p. 21), which has educational implications. Other examples from the history of science can also play the same role, such as: (a) Millikan-Ehrenhaft controversy with respect to the determination of the elementary electrical charge (cf. Holton, 1978a, 1978b; Niaz, 2000a, 2000b, 2005, 2015); (b) Thomson-Rutherford controversy with respect to the alpha particle scattering experiments (Heilbron, 1981a, 1981b; Niaz, 1998, 2009; Wilson, 1983). At this stage it is interesting to consider Taber’s (2014) suggestion that some controversies do not form part of the core commitments of a research tradition and thus their inclusion in Kuhn’s normal science would not undermine Kuhn’s views on science education. It is not clear how that would redress the criticism that, following Kuhn science textbooks “distort the history of science” (Siegel, 1979) or for that matter present a “falsified history of science” (Collins, 2000). This is all the more important if we consider Feyerabend’s (1993) suggestion that the history of science becomes an inseparable part of a science itself.

8.2  Feyerabend’s Epistemological Anarchism

165

8.2.6  Inferring Objectivity from Empirical Approaches It is generally assumed that empirical testing and scientific method provide the best means for achieving (for example in educational evaluation and scientific progress) rationality, truth and hence objective knowledge. A basic idea of the modern natural sciences is bound with an appreciation that they are objective rather than subjective accounts (Shapin, 1996). Bailin (1990) has suggested that both objective and subjective aspects play an important role in scientific progress. On the other hand, Robottom (1989) has questioned the assumption of empirical approaches, on the grounds that an empirical data may presuppose a theoretical background. Furthermore, the relationship between observations and internal cognitive structures has been recognized (Piaget, 1971), leading to theory laden observations. This provides an opportunity to understand the constant confrontation between objectivity and subjectivity in the history of science and thus the need to go beyond an exclusively empirical approach. Based on a historical reconstruction, according to Daston and Galison (2007), what is knowledge and how scientific progress is attained can be understood by the following sequence of practices, that helped to understand objectivity in the making: truth-to-nature (eighteenth century), mechanical objectivity (nineteenth century), structural objectivity (late nineteenth century) and trained judgment (twentieth century). Truth-to-nature refers to science as practiced by Enlightenment naturalists, based on selecting, comparing, judging, and generalizing. It was important for the scientists to be steeped in but not enslaved to nature as it appeared. Those who followed mechanical objectivity were particularly critical of truth-to-nature and considered it to be a subjective distortion. By the late nineteenth century, although mechanical objectivity did not drive out truth-to-nature, it became firmly established as a guide for scientific representation across a wide range of disciplines (Daston & Galison, 2007, p. 111). The photograph became the emblem for all aspects of noninterventionist mechanical objectivity, and this was primarily due to the fact that the camera apparently eliminated human and thus subjective agency. Photography had its own problems with respect to reflecting the object objectively, and this was recognized as early as 1898 by Richard Neuhauss, an expert on photomicrography, as too much light or too little light changed the details in a photograph. The light sensitive photographic plate copies everything even if something does not belong to the object, such as impurities, diffraction edges, dust particles, plate defects and many other artifacts. After working for 40 years in the service of scientific photography Neuhauss became convinced that mechanical objectivity, based on automaticity and noninterference by the scientist was difficult to achieve. (cf. Daston & Galison, 2007, pp. 187–189). Early in the twentieth century scientists came to see the limitations of mechanical objectivity and the need for going beyond by employing trained judgment based on an interpretative vision of the scientific enterprise. Just like structural objectivity, trained judgment was another response to the limitations of the empirical data and photographs used by mechanical objectivity. Within a historical perspective scientists following truth-to-nature (idealized objects) were led to

166

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

mechanical objectivity (actual images and photographs), which in turn led to structural objectivity (relational invariants) and finally came trained judgment, through interpreted images (based on ‘trained’ or ‘seeing’ eye). Of course, this historical transition does not mean that each replaced the other that is instead of supplanting, these different forms of understanding science supplemented each other. Within the historical perspective, the different forms of objectivity represent alternatives that can be supported by groups with different philosophical orientations. Daston and Galison (2007) have provided a detailed overview of how trained judgment came to be an important part of scientific understanding and included various examples from the history of science. One of the most interesting examples is the determination of the elementary electrical charge, the electron (e), which was the subject of considerable controversy between Robert Millikan and Felix Ehrenhaft that lasted for many years (around 1910 to 1923, when Millikan was awarded the Physics Nobel Prize). Both physicists had very similar experimental data and still Millikan postulated the existence of a universal charged particle (the electron) and Ehrenhaft postulated the existence of subelectrons based on fractional charges. Almost 55 years later, Holton (1978a, 1978b) added a new dimension to the controversy with his discovery of Millikan’s two laboratory notebooks at the California Institute of Technology, Pasadena. In these notebooks, Holton found data from 140 drops (from the oil drop experiments), but the published article (Millikan, 1913) reported results from only 58 drops. What happened to the other 82 drops? It seems that Millikan made a rough calculation for the value of e as soon as the data for the times of descent/ascent of the oil drops started coming in and ignored any drop that did not give the value of e that he expected according to his presuppositions (for details see Niaz, 2005). Interestingly, according to Daston and Galison (2007, p. 478) the exercise of scientific judgment in the Millikan-Ehrenhaft controversy can be considered as an example of trained judgment, and not mechanical objectivity. Galison (2015) explicitly endorsed that trained judgment was fundamental for Millikan’s work. More recently, Holton (2014b) has clarified that, “So even if Millikan had included all drops and yet had come out with the same result, the error bar of Millikan’s final result would not have been remarkably small, but large—the very thing Millikan did not like (p. 1, italics in the original). This discussion clearly shows the controversial nature of inferring objectivity based on entirely empirical approaches (for further details see Niaz, 2018).

8.2.7  Methodological Pluralism: Diversity of Methods No single sequence of prescribed activities will guarantee for philosophers or scientists the discovery of new theories and solutions, hence Feyerabend’s emphasis on a diversity of methods (Park, Nielsen & Woodruff, 2014). A “single” best approach to solving a problem is self-defeating as it blinds us to valuable alternative possibilities (Geelan, 1997). In Feyerabend’s view rival theories do not supplant one another. On the contrary they are necessary to one another, and their dialectical interaction reveals powerful epistemological perspectives that are helpful for the development

8.2  Feyerabend’s Epistemological Anarchism

167

of educational theory and practice. Although, the idea of proliferation of theories is also implicit in the writings of Popper and Lakatos, in the case of Feyerabend this is attributed to violation of some rule or method that is generally considered to be part of the scientific method (Loving, 1991). Similar to Feyerabend, Popper also endorsed the idea that there is no scientific method. Some scholars have suggested that “anything goes” means that all knowledge is equally good. On the contrary, Quale (2007) has suggested that it means that the scientist is free to employ all methodological approaches that seem fruitful, with the understanding that some of the attempts may not succeed, and the scientist has to be prepared to accept the results. This precisely leads to what Feyerabend referred to as methodological pluralism. For example, while teaching physics one should not monopolize theoretical physics under only one theory, namely classical mechanics or thermodynamics, but rather use a plural (instead of a Unitarian) approach, in which two strategies are possible (Drago, 1994). Interestingly, this does not necessarily agree with Feyerabend’s ideas on incommensurability, but does coincide with his advice with respect to following diversity, rather than a Unitarian approach. Furthermore, in order to look for breakthroughs in science one does not have to be complacent about the truth of the theories but rather look for opportunities to “break rules” or “violate categories” (cf. Hoffmann, 2012).

8.2.8  Nature of Science Tentative nature of scientific knowledge is an important part of the nature of science. To take an example from physics, Galileo only arrived at the modern theory of inertia by a critical examination of the tower experiment in the light of two alternative frameworks; that of Ptolemy and that of Copernicus (Kalman, 2002). In the case of biology, current views of evolutionary processes are based more on punctuated equilibrium (Gould, 1980) despite their origin in Darwin’s theory of natural selection (Lederman, 1992). In order to pursue the suggestion of Efflin, Glennan and Reisch (1999), it is interesting to consider the study conducted by Tolvanen, Jansson, Vesterinen and Aksela (2014). Based on the participation of Finnish secondary school chemistry teachers, these authors explored the introduction of materials from the chemistry curriculum in order to facilitate an understanding of the following aspects of nature of science (NOS): tentative, difference between theories and laws, empirical, model based, inferential, creativity, social and societal dimensions and instrumentation. The most important innovation of Tolvanen, et al. (2014) is the inclusion of domain-specific aspects in each of the nine domain-general elements. In the case of tentative nature of chemistry the work of A. Lavoisier was discussed to show that some of his ideas were correct and others were later proven wrong. Actually, all accepted theories may eventually change, showing the tentative nature of science. This comes quite close to Feyerabend’s (2011) advice that current view of science may soon be voted out of office.

168

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

8.2.9  Role of Genius in Science The role of genius in science continues to be controversial. Agassi (1975) considered views of the Enlightenment as democratic, whereas those of Polanyi (1972) as elitist and romantic. In contrast, Hattiangadi (1977), described Feyerabend (1975a) as an “egalitarian romantic.” Within a historical perspective a more nuanced position would suggest that as compared to their contemporaries, Columbus, Newton and Einstein, broke the rules (in Feyerabend’s terminology) and thus could be considered as geniuses.

8.2.10  Scientific Expertise Needs a Critical Appraisal Feyerabend was generally both critical and skeptical of scientific expertise (especially under the aegis of the state), and hence his recommendation that science and state be separated (Finocchiaro, 2010). Historically, Feyerabend supported Cardinal Bellarmine’s arguments against Galileo who represented the scientific expertise (cf. Finocchiaro, 2011, 2019 for a recent appraisal; also see Chap. 2). However, Feyerabend did support some of Galileo’s methodological innovations, that he later called “counterinduction” (cf. Kalman, 2009a). More recently, based on the conflict between Cumbrian sheep farmers and British Agriculture authorities after the Chernobyl fallout, Sorgner (2016) has not only found support for Feyerabend’s thesis but also some coincidence with Collins and Evans (2007).

8.2.11  S  cientific Method: Stockpiling and Ordering of Observations Some philosophers of science continue to reiterate that without the scientific method and methodological rules, the self-corrective nature of science becomes a mystery (cf. Irzik & Nola, 2011). On the contrary, Feyerabend would attribute the self-­ corrective nature of science to a diversity and plurality of methods. The traditional scientific method based on the lecture, lab and demo approach is generally followed by most science educators. In contrast, Feyerabend’s approach would emphasize how science really works, namely uncertainty, consideration of wasted efforts and critical encounters with peers (cf. O’Neill & Polman, 2004). Despite the contributions of various philosophers of science (Kuhn, Lakatos, Popper and Feyerabend), the positivist myth with respect to scientific method as the source of “true” scientific knowledge, holds it sway (Rampal, 1992). As part of the reform effort in science education it is important to consider the following question: Should inductive procedures be taught in school science? This is a controversial question as Dewey responded affirmatively, whereas Popper

8.2  Feyerabend’s Epistemological Anarchism

169

responded in the negative. With this background, Swartz (1985) has suggested that this question be rejected as it is based on the erroneous assumption that inductive procedures form a significant part of the method scientists use to test ideas. However, if Feyerabend’s (1981) methodological pluralism is accepted as an intricate part of school science then it will be possible to study the ideas of Dewey, Popper or for that matter anyone else that may facilitate learning, and would approximate to an epistemological anarchism. According to Heering and Höttecke (2014) some of the difficulties associated with the traditional scientific method can be attributed to the fact that some of the methodological procedures are tacit in nature (Polanyi, 1966) and cannot be communicated explicitly. Instead of stockpiling of experimental data, it is the tacit knowledge (Polanyi), among other sources, that helps a scientist to understand its real significance. In the oil drop experiment, Millikan discarded experimental data for which he provided no explicit warrant in his published article (Millikan, 1913). In a study designed to understand the scientific method in the context of the oil drop experiment, an in-­ service science teacher argued that the method was not followed in a strict and rigorous manner and still the scientific community accepted Millikan’s findings (for details see Niaz, 2016). It is plausible to suggest that this teacher’s response approximates to what Hoffmann (2014) has referred to as “humanizing the scientific method.”

8.2.12  Skeletons in the Newtonian Cupboard It is generally believed that based on the inductive method Newton formulated his laws, including the law of gravitation, and this is even considered as a model for other sciences to follow. One of the postulates of the Newtonian method (Rule IV) explicitly stipulated that metaphysical criticism must not be allowed to make us reject inductive proofs. Some philosophers of science (e.g., Whewell, Feyerabend, Norton) have even argued that Newton formulated his method a posteriori to protect and strengthen the law of gravitation (Karam, 2014). Interestingly, several other philosophers of science have endorsed a similar thesis (e.g., Lakatos, Kuhn, Popper, Cartwright, Giere). Lakatos (1978) has pointed out that as early as 1905 (of course leaving aside the criticisms of Newton’s own contemporaries), Duhem had presented a “crushing” criticism of the Newtonian method that revealed some of the skeletons in the Newtonian cupboard.

8.2.13  The New Grew Out of the Old There has been a significant amount of controversy with respect to incommensurability among Feyerabend, Kuhn, Lakatos, Popper and others (cf. Hoyningen-Huene, 1993). However, in the context of classical and quantum mechanics, according to

170

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

Bunge (2003) the new grew out of the old, and that there have been continuities along with discontinuities in the development of scientific theories. For example, Lakatos (1970) emphasized continuity by pointing out that scientists in the nineteenth century recognized the anomalous motion of Mercury (which counted against Newtonian mechanics), but still continued to use it. Kuhn (1970), despite in general agreement with Feyerabend on incommensurability, emphasized that paradigm rejection is a three-step process: an established paradigm, a rival paradigm and the observational evidence. Actually, there are some common elements in the Kuhnian and Lakatosian frameworks. Lakatos had pointed out that refutation of a theory is not necessarily followed by rejection, and that a theory may even flourish within an “ocean of anomalies.” For science education it is important to consider Cartwright’s (1999) advice on this issue: “The easiest way to ensure that no contradictions arise is to become a quantum imperialist and assume that there are no properties of interest besides those studied by quantum mechanics” (p. 233). In other words there are systems whose behavior is adequately described by classical physics. For a science teacher it is important to note that if the new grew out of the old (Bunge, 2003), very soon the new itself may be voted out of office (Feyerabend, 1993), and consequently it sounds good policy to follow some of the guidelines of the old (Cartwright, 1999). This sounds a plausible strategy despite the different philosophical perspectives of Bunge, Cartwright and Feyerabend. Interestingly, in his Conquest of abundance, Feyerabend (1999b) expressed reservations with respect to Kuhn’s views on incommensurability and came quite close to those of Bunge and Cartwright: Following Kuhn … some have asserted that the transition from a comprehensive physical theory to its historical successor involves an act of conversion and that the converted no longer understand the older faith. That is not what we find when we look at history. The transition from classical physics—to the quantum theory—certainly was one of the most radical transformations in the history of science. Yet every stage of the transformation was discussed—Did Bohr and Einstein talk past each other? No. Einstein raised an objection; Bohr was mortified, thought intensely, found an answer, told Einstein, and Einstein accepted the answer. Einstein raised another objection—and so on. Looking at such details, we realize that the conversion philosophy simply does not make sense (Feyerabend, 1999b, p. 267).

This clearly shows that Feyerabend endorsed the intricate relationship between the old and the new, or for that matter between two competing perspectives, perhaps even world views.

8.2.14  Unnatural Nature of Science Cordero (2001) and Nanda (2003) have referred to “voodoo” and “vedic” science and raised the issue of “Isn’t science better than such primitive practices?” At first sight this is a very simple question. However, teaching and doing science is a very complex issue, especially when these are enmeshed with various socio-political and cultural controversies, and this is precisely the context in which Feyerabend is situating the problems. To take one example, the Evolution/Intelligent Design contro-

8.2  Feyerabend’s Epistemological Anarchism

171

versy illustrates the issues involved (for details see the section “Evolution, knowledge and belief” in Chap. 3). According to a survey conducted by the Pew Research Center, in March–April, 2013, 33% of U.S. adults reject the idea of evolution saying that, “humans and other living things have existed in their present form since the beginning of time”. On the other hand, 60% of U.S. adults say that, “humans and other living things have evolved over time.” Interestingly, however, of those who believe in evolution 24% say that “a supreme being guided the evolution of living things for the purpose of creating humans and other life in the form it exists today” (downloaded on 24 December, 2018 from www.pewresearch.org, a nonpartisan fact tank that informs the public about different issues, and this is the latest survey on this subject). With this background (skepticism about evolution) it is easier to understand the unnatural nature of science. Wolpert (1993) a developmental biologist has referred to the difficulties involved in understanding scientific practice, precisely due to the unnatural nature of science: […] both the ideas that science generates and the way in which science is carried out are entirely counter intuitive and against common sense—by which I mean that scientific ideas cannot be acquired by simple inspection of phenomena and that they are often outside everyday experience. Science does not fit with our natural expectations (p. 1, italics added).

No wonder, Feyerabend considered that astrology and voodoo are attractive alternatives to evolution, especially if these are presented as based on “simple inspection of phenomena” or “dogmatic scientism.” Precisely, this was Mugaloglu’s (2014) dilemma while teaching evolution in Turkey (see Chap. 3). Indeed, most science curricula and textbooks reduce “scientific practice” to a “simple inspection of phenomena”. History of science provides many similar examples, such as: Alpha particle scattering experiments and the ensuing controversy between J.J. Thomson and E. Rutherford (cf. Niaz, 2009; Wilson, 1983).

8.2.15  Was Feyerabend a Postmodern or Perspectival Realist? Most science educators and researchers in science education consider Feyerabend to be a postmodern philosopher of science. In a recent appraisal, Mackenzie, Good and Brown, (2014), consider Feyerabend to be a postmodern as he questioned some of the central features of the standard picture of science, namely scientific reason and objectivity, among others. Actually, Feyerabend’s philosophy of science was based on how science really works and recent literature in philosophy of science has recognized that his oeuvre was more modern (in the Enlightenment tradition) than postmodern (e.g., Brown & Kidd, 2016). Philosophy of science itself has explored new territory in this context and Giere (2006a, p. 95) considers that it is presentist hubris to think that we can have an objectively correct or true theories, and that progress in science is perspectival. Furthermore, in a historicized vision of scientific judgment, Daston and Galison (2007) have concluded that there is no objectivity without subjectivity to suppress and vice versa (p. 33). It is precisely in this context

172

8  Conclusion: Feyerabend and Challenges of the Twenty-First Century

that Feyerabend (1974/1988) claimed that “true belief limits freedom” and led McCarthy (2014) to suggest that this would lead to “wholesale rejection of modern science.” On the contrary, Feyerabend often expressed his love and admiration for science. Finally, according to Giere (2016) Feyerabend is not only a perspectivist but also a perspectival realist.

8.3  Educational Implications Based on different chapters of this book, following educational implications can provide students, teachers and researchers an opportunity to reflect and design new teaching strategies: • As progress in science is based on controversy, established theories and ways of thinking are difficult to change, thus making counterinduction not only essential but also challenging. • Accepting unsupported hypotheses (counterinduction) can help the rival hypothesis to strengthen its arguments. • An unsupported hypothesis can even be based on an inconsistent foundation. • Feyerabend’s advice to students: “Current view of science may soon be voted out of office” approximates to what science educators refer to as the “tentative nature of science.” • Traditional science teaching emphasizes objective facts. These are, however, not immutable truths but rather working hypotheses, useful for future research. • Science can provide more than one “correct” theory of the same phenomena and that would represent what Feyerabend referred to as “epistemological anarchism.” • Feyerabend’s “epistemological anarchism” basically refers to “how science really works” and the need to go beyond two centuries of empiricism. • “Anything goes” means that a scientist employs all sorts of methodological strategies to see what works and that a methodological strategy is not decided a priori. • Based on a historical reconstruction, history of science becomes an inseparable part of the science itself. • History of science shows a constant confrontation between objectivity and subjectivity and thus the need to go beyond an exclusively empirical approach. • For breakthroughs in science one does not have to be complacent about the “truth” of the theories but rather look for opportunities to “break rules” or “violate categories.” • If science looks for “truth”, Feyerabend claimed that would limit/curtail freedom of thought. Indeed, if scientists have found the “truth” there is nothing left for students to do in the future.

8.3  Educational Implications

173

• A dialectical interaction among rival theories leads to methodological pluralism. • Scientific expertise needs to be subject to a critical appraisal by the society. • Tacit knowledge is more helpful to a scientist rather than stockpiling of experimental data. • Methodological pluralism facilitates the inclusion of opposing ideas in school science leading to epistemological anarchism. • Diversity and plurality of methods and not the scientific method facilitates the self-corrective nature of science. • Uncertainty with respect to scientific knowledge need not be a constraint in learning science. • Newtonian method based on inductive generalization and espoused in school science is difficult to follow in real science. • If the new grew out of the old (quantum mechanics out of Newtonian mechanics) and soon the new itself, may be voted out of office, it is plausible to follow some of the guidelines of the old.

Appendices

 ppendix 1: Articles from the Journal Science & Education A (Springer) Evaluated in This Study (n = 78) Arriassecq, I., & Greca, I.M. (2012). A teaching-learning sequence for the special relativity theory at high school level historically and epistemologically contextualized. Science & Education, 21(6), 827–851. Ben-Ari, M. (2005). Situated learning in “this high-technology world”. Science & Education, 14(3–5), 367–376. Blake, D.D. (1994). Revolution, revision or reversal: Genetics — ethics curriculum. Science & Education, 3(4), 373–391. Bunge, M. (2003). Twenty-five centuries of quantum physics: From Pythagoras to us, and from subjectivism to realism. Science & Education, 12(5–6), 445–466. Bunge, M. (2012). Does quantum physics refute realism, materialism and determinism? Science & Education, 21(10), 1601–1610. Chang, H. (2011). How historical experiments can improve scientific knowledge and science education: The cases of boiling water and electrochemistry. Science & Education, 20(3–4), 317–341. Clarke, S.W. (2016). Review of Agassi’s Popper and his popular critics: Thomas Kuhn, Paul Feyerabend and Imre Lakatos. Science & Education, 25(1–2), 221–227. Cobern, W.W. (1995). Science education as an exercise in foreign affairs. Science & Education, 4(3), 287–302. Cordero, A. (1992). Science, objectivity and moral values. Science & Education, 1(1), 49–70. Cordero, A. (2001). Scientific culture and public education. Science & Education, 10(1–2), 71–83. Davson-Galle, P. (2004). Philosophy of science, critical thinking and science education. Science & Education, 13(6), 503–517. © Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3

175

176

Appendices

De Berg, K.C. (2011). Joseph Priestley across theology, education, and chemistry: An interdisciplinary case study in epistemology with a focus on the science education context. Science & Education, 20(7–8), 805–830. De Berg, K. (2014). Teaching chemistry for all its worth: The interaction between facts, ideas, and language in Lavoisier’s and Priestley’s chemistry practice: The case of the study of the composition of air. Science & Education, 23(10), 2045–2068. Develaki, M. (2007). The model-based view of scientific theories and the structuring of school science programmes. Science &Education, 16(7–8), 725–749. Develaki, M. (2008). Social and ethical dimension of the natural sciences, complex problems of the age, interdisciplinarity, and the contribution of education. Science & Education, 17(8–9), 873–888. Develaki, M. (2012). Integrating scientific methods and knowledge into the teaching of Newton’s theory of gravitation: An instructional sequence for teachers’ and students’ nature of science education. Science & Education, 21(6), 853–879. Drago, A. (1994). Mach’s thesis: Thermodynamics as the basic theory for physics teaching. Science & Education, 3(2), 189–198. Duschl, R.A., & Grandy, R. (2013). Two views about explicitly teaching nature of science. Science & Education, 22(9), 2109–2139. Eger, M. (1993). Hermeneutics as an approach to science: Part I. Science & Education, 2(1), 1–29. Ernest, P. (1993). Constructivism, the psychology of learning, and the nature of mathematics: Some critical issues. Science & Education, 2(1), 87–93. Finocchiario, M.A. (2011). A Galilean approach to the Galileo affair. Science & Education, 20(1), 51–66. Gauch, H.G. (2009). Science, worldviews and education. Science &Education, 18(6–7), 667–695. Geelan, D.R. (1997). Epistemological anarchy and the many forms of constructivism. Science & Education, 6(1–2), 15–28. Gil-Pérez, D., Vilches, A., Fernández, I., Cachapuz, A., Praia, J., Valdés, P., & Salinas, J. (2005). Technology as ‘applied science.’ Science & Education, 14(3– 5), 309–320. Ginev, D.J. (2008). Hermeneutics of science and multi-gendered science education. Science & Education, 17(10), 1139–1156. Glass, R.J. (2013). Tacit beginnings towards a mode of scientific thinking. Science & Education, 22(10), 2709–2725. Good, R., & Shymansky, J. (2001). Nature-of-science literacy in benchmarks and standards: Post-modern/relativist or modern/realist? Science & Education, 10(1–2), 173–185. Goodney, D.E., & Long, C.S. (2003). The collective classic: A case for the reading of science. Science & Education, 12(2), 167–184. Guerra-Ramos, M.T. (2012). Teachers’ ideas about the nature of science: A critical analysis of research approaches and their contribution to pedagogical practice. Science & Education, 21(5), 631–655.

Appendices

177

Hafner, R., & Culp, S. (1996). Elaborating the structures of a science discipline to improve problem-solving instruction: An account of classical genetics’ theory structure, function, and development. Science & Education, 5(4), 331–355. Heffron, J.M. (1995). The knowledge most worth having: Otis W. Caldwell (1869– 1947) and the rise of the general science course. Science & Education, 4(3), 227–252 Hodson, D. (1992). Assessment of practical work. Science & Education, 1(2), 115–144. Howard, D. (2009). Better red than dead — Putting an end to the social irrelevance of postwar philosophy of science. Science & Education, 18(2), 199–220. Irzik, G. (2000). Back to basics: A philosophical critique of constructivism. Science & Education, 9(6), 621–639. Irzik, G., & Nola, R. (2011). A family resemblance approach to the nature of science for science education. Science & Education, 20(7–8), 591–607. Izquierdo, M., & Adúriz-Bravo, A. (2003). Epistemological foundations of school science. Science & Education, 12(1), 27–43. Jacobs, S. (2000). Michael Polanyi on the education and knowledge of scientists. Science & Education, 9(3), 309–320. Jung, W. (1994). Toward preparing students for change: A critical discussion of the contribution of the history of physics in physics teaching. Science & Education, 3(2), 99–130. Kalman, C.S. (2002). Developing critical thinking in undergraduate courses: A philosophical approach. Science & Education, 11(1), 83–94. Kalman, C.S., & Aulls, M.W. (2003). Can an analysis of the contrast between pre-­ Galilean and Newtonian theoretical frameworks help students develop a scientific mindset. Science & Education, 12(8), 761–772. Kalman, C.S. (2009a). A role for experiment in using the law of inertia to explain the nature of science: A comment on Lopes Coelho. Science & Education, 18(1), 25–31. Kalman, C.  S. (2009b). The need to emphasize epistemology in teaching and research. Science & Education, 18(3–4), 325–347. Kalman, C.S. (2011). Enhancing students’ conceptual understanding by engaging science text with reflective writing as a hermeneutical circle. Science & Education, 20(2), 159–172. Kanderakis, N.E. (2010). When is a physical concept born? The emergence of ‘work’ as a magnitude of mechanics. Science & Education, 19(10), 995–1012. Karakostas, V., & Hadzidaki, P. (2005). Realism vs. constructivism in contemporary physics: The impact of the debate on the understanding of quantum theory and its instructional process. Science & Education, 14(7–8), 607–629. Karam, R. (2014). Review of Achinstein’s Evidence and Method: Scientific strategies of Isaac Newton and James Clerk Maxwell. Science &Education, 23(10), 2137–2148. Kendig, C. (2013). Integrating history and philosophy of the life sciences in practice to enhance science education: Swammerdam’s historia insectorum generalis and the case of the water flea. Science & Education, 22(8), 1939–1961.

178

Appendices

Koertge, N. (1996). Toward an integration of content and method in the science curriculum. Science & Education, 5(4), 391–406 (first published in 1969). Kousathana, M., Demerouti, M., & Tsaparlis, G. (2005). Instructional misconceptions in acid-base equilibria: An analysis from a history and philosophy of science perspective. Science & Education, 14(2), 173–194. Kötter, M., & Hammann, M. (2017). Controversy as a blind spot in teaching nature of science. Science & Education, 26(5), 451–482. Lamont, J. (2009). Fall and rise of Aristotelian metaphysics in the philosophy of science. Science & Education, 18(6–7), 861–884. Lederman, N.G. (1995). Suchting on the nature of scientific thought: Are we anchoring curricula in quicksand. Science & Education, 4(4), 371–377. Matthews, M.R. (1992). History, philosophy, and science teaching: The present rapprochement. Science & Education, 1(1), 11–47. Matthews, M.R. (2004). Reappraising positivism and education: The arguments of Philipp Frank and Herbert Feigl. Science & Education, 13(1–2), 7–39. Matthews, M.R. (2009). Science, worldviews and education: An introduction. Science & Education, 18(6–7), 641–666. Mugaloglu, E.Z. (2014). The problem of pseudoscience in science education and implications of constructivist pedagogy. Science & Education, 23(4), 829–842. Nola, R. (2004a). Review of Meera Nanda’s Prophets facing backwards: Postmodern critiques of science and Hindu nationalism in India. Science & Education, 13(3), 243–249. Nola, R. (2004b). Pendula, models, constructivism and reality. Science & Education, 13(4–5), 349–377. Park, H., Nielsen, W., & Woodruff, E. (2014). Students’ conceptions of the nature of science: Perspectives from Canadian and Korean middle school students. Science & Education, 23(5), 1169–1196. Pennock, R.T. (2010). The postmodern sin of intelligent design creationism. Science & Education, 19(6–8), 757–778. Pinnick, C.L. (2005). The failed feminist challenge to “fundamental epistemology”. Science & Education, 14(2), 103–116. Quale, A. (2007). Radical constructivism, and the sin of relativism. Science & Education, 16(3–5), 231–266. Reydon, T.A.C. (2013). Classifying life, reconstructing history and teaching diversity: Philosophical issues in the teaching of biological systematics and biodiversity. Science & Education, 22(2), 189–220. Rowbottom, D.P. (2013). Review of Feyerabend’s The tyranny of science. Science & Education, 22(5), 1229–1231. Schulz, R.M. (2009). Reforming science education: Part II. Utilizing Kieran Egan’s educational metatheory. Science & Education, 18(3–4), 251–273. Siemsen, H. (2011). Ernst Mach and the epistemological ideas specific for Finnish science education. Science & Education, 20(3–4), 245–291. Slezak, P. (1994). Sociology of scientific knowledge and science education, Part 2. Science & Education, 3(4), 329–355.

179

Appendices

Slezak, P. (2011). Review of Maurice A. Finocchiaro: Defending Copernicus and Galileo: Critical reasoning in the two affairs. Science & Education, 20(1), 71–81. Strömdahl, H.R. (2012). On discerning critical elements, relationships and shifts in attaining scientific terms: The challenge of polysemy/homonymy and reference. Science & Education, 21(1), 55–85. Suchting, W.A. (1992). Constructivism deconstructed. Science & Education, 1(3), 223–254. Suchting, W.A. (1994). Notes on the cultural significance of the sciences. Science & Education, 3(1), 1–56. Suchting, W.A. (1995). The nature of scientific thought. Science & Education, 4(1), 1–22. Taber, K.S. (2008). Towards a curricular model of the nature of science. Science & Education, 17(2–3), 179–218. Tala, S., & Vesterinen, V.-M. (2015). Nature of science contextualized: Studying nature of science with scientists. Science & Education, 24(4), 435–457. Tweney, R.D. (2012). Review of Nersessian’s Creating scientific concepts. Science & Education, 21(4), 591–596. Viana, H.E.B., & Porto, P.A. (2010). The development of Dalton’s atomic theory as a case study in the history of science: Reflections for educators in chemistry. Science & Education, 19(1), 75–90. Villani, A., Almeida Pacca, J.L., & Freitas, D. (2009). Science teacher education in Brazil: 1950–2000. Science & Education, 18(1), 125–148. Wendel, P.J. (2011). Object-based epistemology at a creationist museum. Science & Education, 20(1), 37–50. Woodcock, B.A. (2014). “The scientific method” as myth and ideal. Science & Education, 23(10), 2069–2093.

 ppendix 2: Distribution of Articles (Science & Education) A According to Author’s Area of Research, Context of the Study and Level (Classification), n = 78

No. Authors in the references 1 Arriassecq, I., & Greca, I.M. (2012). 2 Ben-Ari, M. (2005) 3 Blake, D.D. (1994) 4 Bunge, M. (2003)

Author’s area of research Science education Science education Science education Philosophy of science

Context of the study Special relativity theory and experimental verification Situated learning Genetics-ethics curriculum Quantum physics

Level III III II III

(continued)

Appendices

180

6

Chang, H. (2011)

7

Clarke, S.W. (2016)

8 9

Cobern, W.W. (1995) Cordero, A. (1992)

10

Cordero, A. (2001)

11

Davson-Galle, P. (2004)

12

De Berg, K.C. (2011)

Author’s area of research Philosophy of science History and philosophy of science Philosophy of science Science education Philosophy of science Philosophy of science Philosophy of education Science education

13 14 15

De Berg, K.C. (2014) Develaki, M. (2007) Develaki, M. (2008)

Science education Science education Science education

16 17 18

Science education History of physics Science education

19 20 21

Develaki, M. (2012) Drago, A. (1994) Duschl, R.A., & Grandy, R. (2013) Eger, M. (1993) Ernest, P. (1993) Finocchiaro, M.A. (2011)

22

Gauch, H.G. (2009)

23

Geelan, D.R. (1997)

Science education Education Philosophy of science Philosophy of science Science education

24 25

Gil-Pérez et al. (2005) Ginev, D.J. (2008)

Science education Philosophy

26 27

29

Glass, R.J. (2013) Good, R., & Shymansky, J. (2001) Goodney, D.E., & Long, C.S. (2003) Guerra-Ramos, M.T. (2012)

30 31 32

Heffron, J.M. (1995) Hodson, D. (1992) Howard, D. (2009)

No. Authors in the references 5 Bunge, M. (2012)

28

Context of the study Quantum physics and determinism Historical experiments and science education

Level I III

Diversity and pluralism in science Radical social constructivism Naïve falsificationism

IV IV II

Scientific culture and education IV Critical thinking and science education Theory-ladenness of observation Priestley-Lavoisier controversy Model-based scientific theories Social and ethical dimension of science Integrating scientific methods Mach and thermodynamics Naturalized philosophy of science Hermeneutics and science Radical constructivism Defending Copernicus and Galileo Science and worldviews

I III III II III II IV II III III V I

Education Science education

Epistemological anarchy and Constructivism Scientific method Multi-gendered science education Tacit knowledge Nature of science literacy

IV I

Science education

Reading classics of science

II

Science education

Epistemological anarchism and scientific method Inductive method Learning to do science Social epistemology of science

IV

Education Science education Philosophy of science

V I III

III IV II

(continued)

181

Appendices

No. Authors in the references 33 Irzik, G. (2000)

Author’s area of research Philosophy of science Philosophy of science Science education

34

Irzik, G., & Nola, R. (2011)

35 36

Izquierdo, M., & Adúriz-­ Bravo, A. (2003) Jacob, S. (2000)

37 38

Jung, W. (1994) Kalman, C.S. (2002)

39 40

Kalman, C.S., & Aulls, M.W. (2003) Kalman, C.S. (2009a)

Science education

41 42

Kalman, C. S. (2009b) Kalman, C.S. (2011)

Science education Science education

43

Kanderakis, N.E. (2010)

Education

44 45 46

Karakostas, V., & Hadzidaki, P. (2005) Karam, R. (2014) Kendig, C. (2013)

Philosophy of science Science education Philosophy

47

Koertge, N. (1996)

Philosophy of science Science education Science education Science education

51 52

Philosophy of science Kousathana, M., Demerouti, Science education M., & Tsaparlis, G. (2005) Kötter, M., & Hammann, Science education M. (2017) Lamont, J. (2009) Philosophy of science Lederman, N.G. (1995) Science education Matthews, M.R. (1992) Science education

53 54

Matthews, M.R. (2004) Matthews, M.R. (2009)

Science education Science education

55

Mugaloglu, E.Z. (2014)

Science education

56

Nola, R. (2004a)

57

Nola, R. (2004b)

Philosophy of science Philosophy of science

48 49 50

Context of the study Radical constructivism

Level II

Family resemblance and nature of science Theory-laden observations

II II

Polanyi and science education

III

Incommensurability Critical thinking and conceptual change Theoretical frameworks and scientific mindset Role of experiment and nature of science Epistemology of science Conceptual understanding and hermeneutical circle Emergence of a physical concept Realism and constructivism

II V V V V II II I

Newtonian method History and philosophy of life sciences Competing theoretical systems

III II IV

Acid-base equilibria

III

Teaching nature of science

II

Theory and experience

III

Nature of science History, philosophy, and science teaching Reappraising positivism Enlightenment and science education Pseudoscience and constructivist pedagogy Hindu nationalism and Feyerabend Subversion of experience by reason

IV II II III II I III

(continued)

Appendices

182

No. Authors in the references 58 Park, H., Nielsen, W., & Woodruff, E. (2014) 59 Pennock, R.T. (2010) 60 61 62

Pinnick, C.L. (2005) Quale, A. (2007) Reydon, T.A.C. (2013)

63

Rowbottom, D.P. (2013)

64 65

Schulz, R.M. (2009) Siemsen, H. (2011)

66

Slezak, P. (1994)

67

Slezak, P. (2011)

68 69

Strömdahl, H.R. (2012) Suchting, W.A. (1992)

70

Suchting, W.A. (1994)

71

Suchting, W.A. (1995)

72 73

Taber, K.S. (2008) Tala, S., & Vesterinen, V.-M. (2015) Tweney, R.D. (2012)

74 75 76 77 78

Viana, H.E.B., & Porto, P.V. (2010) Villani, A., Almeida Pacca, J.L., & Freitas, D. (2009) Wendel, P.J. (2011) Woodcock, B.A. (2014)

Author’s area of research Science education Philosophy of biology Philosophy Teacher education Philosophy of biology Philosophy of science Science education Science education History & philosophy History & philosophy Science education Philosophy of science Philosophy of science Philosophy of science Science education Science education Cognitive psychology Science education Science education Science education Philosophy of science

Context of the study Diversity of methods

Level IV

Intelligent design creationism

II

Feminism Radical constructivism Meaning change

II V II

Normal science, dogmatism & science education Science education reform Mach and Finnish science education Sociological constructivism

III

Defending Copernicus and Galileo Referents in conceptual change Natural interpretations

III II I III III II

Cultural significance of science IV Nature of scientific thought

II

Nature of science Contextualized nature of science Practice of science

II II II

Atomic theory and history of science Teacher education in Brazil

I

Object-based epistemology Scientific method

III II

Notes: 1. In the case of more than one author, area of research refers to that of the first author 2. For a description of Levels (I – V) see Chap. 3

II

Appendices

183

 ppendix 3: Articles from the Journal of Research in Science A Teaching (Wiley Blackwell) Evaluated in This Study, n = 21 Abd-El-Khalick, F., & Lederman, N.G. (2000). The influence of history of science courses on students’ views of nature of science. Journal of Research in Science Teaching, 37(10), 1057–1095. Abd-El-Khalick, F., Myers, J.Y., Summers, R., Brunner, J., Waight, N., Wahbeh, N., Zeineddin, A.A., & Belarmino, J. (2017). A longitudinal analysis of the extent and manner of representations of nature of science in U.S. high school biology and physics textbooks. Journal of Research in Science Teaching, 54(1), 82–120. Abd-El-Khalick, F., Waters, M., & Le, A.-P. (2008). Representations of nature of science in high school chemistry textbooks over the past four decades. Journal of Research in Science Teaching, 45(7), 835–855. Beghetto, R.A. (2007). Factors associated with middle and secondary students’ perceived science competence. Journal of Research in Science Teaching, 44(6), 800–814. Duschl, R.A., & Wright, E. (1989). A case study of high school teachers’ decision making models for planning and teaching science. Journal of Research in Science Teaching, 26(6), 467–501. Eflin, J.T., Glennan, S., & Reisch, G. (1999). The nature of science: A perspective from the philosophy of science. Journal of Research in Science Teaching, 36(1), 107–116. Good, R.G. (1993). Editorial: The slippery slopes of postmodernism. Journal of Research in Science Teaching, 30(5), 427. Hawkins, J., & Pea, R.D. (1987). Tools for bridging the cultures of everyday and scientific thinking. Journal of Research in Science Teaching, 24(4), 291–307. Lederman, N.G. (1992). Students’ and teachers’ conceptions of the nature of science: A review of the research. Journal of Research in Science Teaching, 29(4), 331–359. Lederman, N.G., Abd-El-Khalick, F., Bell, R.L., & Schwartz, R.S. (2002). Views of nature of science questionnaire: Toward valid and meaningful assessment of learners’ conceptions of nature of science. Journal of Research in Science Teaching, 39(6), 497–521. Loving, C.C. (1991). The Scientific Theory Profile of science model for science teachers. Journal of Research in Science Teaching, 28(9), 823–838. Martin, M. (1970). Anomaly-recognition and research in science education. Journal of Research in Science Teaching, 7(3), 187–190. Moshman, D., & Thompson, P.A. (1981). Hypothesis testing in students: Sequences, stages, and instructional strategies. Journal of Research in Science Teaching, 18(4), 341–352. O’Neill, D.K., & Polman, J.L. (2004). Why educate “little scientists?” Examining the potential of practice-based scientific literacy. Journal of Research in Science Teaching, 41(3), 234–266.

184

Appendices

Palmquist, B.C., & Finley, F.N. (1997). Preservive teachers’ views of the nature of science during a postbaccalaureate science teaching program. Journal of Research in Science Teaching, 34(6), 595–615. Robottom, I. (1989). Social critique or social control: Some problems for evaluation in environmental education. Journal of Research in Science Teaching, 26(5), 435–443. Roth, W.-M. (1993). Heisenberg’s uncertainty principle and interpretive research in science education. Journal of Research in Science Teaching, 30(7), 669–680. Roth, W.-M., & Roychoudhury, A. (1994). Physics students’ epistemologies and views about knowing and learning. Journal of Research in Science Teaching, 31(1), 5–30. Roth, W.-M., McRobbie, C.J., Lucas, K.B., & Boutonné, S. (1997). Why may students fail to learn from demonstrations? A social practice perspective on learning physics. Journal of Research in Science Teaching, 34(5), 509–533. Roth, W.-M., & McGinn, M.K. (1998). Knowing, researching, and reporting science education: Lessons from science and technology studies. Journal of Research in Science Teaching, 35(2), 213–235. Roychoudhury, A., Tippins, D.J., & Nichols, S.E. (1995). Gender-inclusive science teaching: A feminist-constructivist approach. Journal of Research in Science Teaching, 32(9), 897–924.

 ppendix 4: Distribution of Articles (Journal of Research A in Science Teaching) According to Author’s Area of Research, Context of the Study and Level (Classification), n = 21

Author’s area of No. Authors in the reference research 1 Abd-El-Khalick & Lederman (2000) Science education 2 Abd-El-Khalick et al (2017) Science education 3 Abd-El-Khalick, Waters & Le Science (2008) education 4 Beghetto, R.A. (2007) Science education 5 Duschl & Wright (1989) Science education 6 Eflin, Glennan & Reisch (1999) Philosophy of science 7 Good, R.G. (1993) Science education

Context of the study Students’ views of nature of science Nature of science in science textbooks Nature of science in chemistry textbooks Science competence

Level II

III

Worldviews

I

Nature of science

II

Postmodernism

I

II II

(continued)

Appendices

No. Authors in the reference 8 Hawkins, J., & Pea, R.D. (1987) 9

Lederman (1992)

10 11

Lederman, Abd-El-Khalick, Bell & Schwartz (2002) Loving (1991)

12

Martin, M. (1970)

13

Moshman, D., & Thompson, P.A. (1981) O’Neill & Polman (2004)

14 15 16

Palmquist, B.C., & Finley, F.N. (1997) Robottom, I. (1989)

17

Roth, W.-M. (1993)

18

Roth, W.-M. & McGinn, M.K. (1998) Roth, W.-M., McRobbie, C.J., Lucas, K.B., & Boutonné, S. (1997) Roth, W.-M., & Roychoudhury, A. (1994) Roychoudhury, Tippins, & Nichols (1995)

19 20 21

185 Author’s area of research Science education Science education Science education Science education Philosophy of science Psychology Science education Science education Science education Science education Science education Science education Science education Science education

Context of the study Everyday and Scientific thinking Nature of science

Level III III

Nature of science

II

Scientific theory profile

V

Proliferation of theories

V

Hypothesis testing

III

Scientific method

IV

Nature of science

II

Evaluation

III

Heisenberg’s uncertainty principle Science and technology studies Teacher demonstrations

II

II

Students’ epistemologies

II

Feminism

II

IV

Notes: 1. In the case of more than one author, area of research refers to that of the first author 2. For a description of Levels (I –V), see Chap. 3

 ppendix 5: Articles from the Journal Interchange Evaluated A in This Study (n = 15) Agassi, J. (1996). Book review of Feyerabend’s Killing Time. Interchange, 27(1), 85–88. Bailin, S. (1990). Creativity, discovery, and science education: Kuhn and Feyerabend revisited. Interchange, 21(3), 34–44. Brown, J.R. (1997). Academic freedom, affirmative action, and the advance of knowledge. Interchange, 28(4), 381–388.

Appendices

186

Brush, S.G. (1989). History of science and science education. Interchange, 20(2), 60–70. Fried, M.N. (2011). Theories for, in, and of mathematics education. Interchange, 42(1), 81–95. Hattiangadi, J.N. (1985). Novelty, creation and society. Interchange, 16(1), 40–50. Niaz, M. (2004). Exploring alternative approaches to methodology in educational research. Interchange, 35(2), 155–184. Niaz, M. (2011). How to facilitate teachers’ understanding of hypotheses and predictions? Interchange, 42(1), 51–58. Pope, M. L. (1982). Personal construction of formal knowledge. Interchange, 13(4), 3–14. Rampal, A. (1992). Maintaining the status quo — a response to Fred Wilson and John Wilson. Interchange, 23(3), 309–314. Sriraman, B. (2008). Let Lakatos be! Commentary on “Would the real Lakatos please stand up”. Interchange, 39(4), 483–492. Swartz, R. (1985). Dewey and Popper on learning from induction. Interchange, 16(4), 29–51. Winchester, I. (1989). Editorial: History, science and science teaching. Interchange, 20(2), i–vi. Winchester, I. (1993). “Science is dead. We have killed it, you and I” — How attacking the presuppositional structures of our scientific age can doom the interrogation of nature. Interchange, 24(1–2), 191–198. Woodhouse, H., & Ndongko, T.M. (1993). Women and science education in Cameroon: Some critical reflections. Interchange, 24(1–2), 131–158.

 ppendix 6: Distribution of Articles (Interchange) According A to Author’s Area of Research, Context of the Study and level (Classification), n = 15

No Authors in the reference 1 Agassi, J. (1996) 2

Bailin, S. (1990)

3 4 5

Brown, J.R. (1997) Brush, S.G. (1989) Fried, M.N. (2011)

6

Hattiangadi, J.N. (1985)

7

Niaz, M. (2004)

Author’s area of research Philosophy of science Education Philosophy History of science Mathematics education Philosophy of science Science education

Context of the study Rationalism

Level III

Feyerabend, Kuhn and science education Rival theories Working hypotheses Theories for mathematics education Genius in science

III

III

Growth of knowledge

III

IV III I

(continued)

Appendices

No Authors in the reference 8 Niaz, M. (2011) 9 Pope, M.L. (1982) 10 Rampal, A. (1992) 11 Sriraman, B. (2008) 12 13 14 15

187 Author’s area of research Science education

Psychology Science studies Mathematics education Swartz, R. (1985) Philosophy of education Winchester, I. (1989) Philosophy of science Winchester, I. (1993) Philosophy of science Woodhouse, H., & Ndongko, Education T.M. (1993)

Context of the study Understanding hypotheses and predictions Constructive alternativism Scientific method Competing theoretical perspectives Induction and learning

Level III

V

History of science

III

Presuppositional structures

III

African and modern medicine

III

V III III

Notes: 1. In the case of more than one author, area of research refers to that of the first author 2. For a description of Levels (I – V), see Chap. 3

 ppendix 7: Articles from the International Handbook A of Research in History, Philosophy and Science Teaching (Springer), n = 6 Heering, P., & Höttecke, D. (2014). Historical-investigative approaches in science teaching. In M.R.  Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. II, pp. 1473–1502). Dordrecht: Springer. Hodson, D. (2014). Nature of science in the science curriculum: Origin, development, implications and shifting emphases. In M.R. Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. II, pp. 911–970). Dordrecht: Springer. Mackenzie, J., Good, R.G., & Brown, J.R. (2014). Postmodernism and science education: An appraisal. In M.R.  Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. II, pp. 1057–1086). Dordrecht: Springer. McCarthy, C.L. (2014). Cultural studies in science education: Philosophical considerations. In M.R.  Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. III, pp. 1927–1964). Dordrecht: Springer. Schulz, R.M. (2014). Philosophy of education and science education: A vital but underdeveloped relationship. In M.R. Matthews (Ed.), International Handbook

188

Appendices

of Research in History, Philosophy and Science Teaching (Vol. II, pp.  1259– 1316). Dordrecht: Springer. Taber, K.S. (2014). Methodological issues in science education research: A perspective from the philosophy of science. In M.R.  Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. III, pp. 1839–1893). Dordrecht: Springer.

 ppendix 8: Distribution of Articles (International Handbook A of Research in History, Philosophy and Science Teaching) According to Author’s Area of Research, Context of the Study and Level (Classification), n = 6.

No. Authors in the reference 1 Heering, P., & Höttecke, D. (2014) 2 Hodson, D. (2014) 3 Mackenzie, J., Good, R.G., & Brown, J.R. (2014) 4 McCarthy, C.L. (2014) 5 6

Schulz, R.M. (2014) Taber, K.S. (2014)

Author’s area of research Science education

Context of the study Historical-investigative approach Science education Nature of science Educational theory Postmodernism Philosophy of education Science education Science education

Cultural studies in science education Philosophy of education Methodological issues

Level III III III III III III

Notes: 1. For a description of Levels (I – V) see Chap. 3 2. In the case of more than one author, area of research refers to that of the first author

 ppendix 9: List of General Chemistry Textbooks Evaluated A in Different Studies of This Book (n = 128) Ander, P. & Sonnessa A. (1968). Principles of chemistry (Spanish ed.). New York: Macmillan. Andrews, D.H. & Kokes, R.  J. (1962). Fundamental chemistry (Spanish ed.). New York: Wiley. Atkins, P.W. (1989). General chemistry. New York: Scientific American Books. Atkins, P. & Jones, L. (2002). Chemical principles: The quest for insight (2nd ed.). New York: Freeman.

Appendices

189

Atkins, P. & Jones, L. (2008). Chemical principles: The quest for insight (4th ed.). New York: Freeman. Barrow, G. M. (1972). General chemistry (Spanish ed.). Belmont, CA: Wadsworth. Basset, L., Bounce, S., Carter, A., Clark, H. & Holinger, H. (1966). Principles of chemistry. Englewood Cliffs, NJ: Prentice Hall. Becker, R.S. & Wentworth, W.E. (1977). General chemistry (Spanish ed.). Boston: Houghton Mifflin. Bishop, M. (2002). An introduction to chemistry. San Francisco: Benjamin Cummings. Bodner, G. & Pardue, H. (1989). Chemistry: An experimental science. New York: Wiley. Brady, J. (2000). General chemistry: Principles and structures (2nd ed., Spanish). New York: Wiley. Brady, J. & Humiston, G. (1996). General Chemistry: Principles and structure (Spanish ed.). New York: Wiley. Brady, J., Russell, J. & Holum, J. (2000). Chemistry: The study and its changes (3rd ed.). New York: Wiley. Breck, W.G., Brown, R.J.C. & McCowan, J.D. (1981). Chemistry for science and engineering. New York: McGraw-Hill. Brescia, F., Arents, J., Meislich, H. & Turk, A. (1970). Fundamentals of chemistry. A modern introduction (Spanish ed.). New York: Academic Press. Brown, L.S. & Holme, T.A. (2011). Chemistry for engineering students (2nd ed.). Belmont, CA: Brooks/Cole (Cengage Learning). Brown, T.L., Le May, H.E. & Bursten, B. (1998). Chemistry: The central science (7th ed., Spanish). Englewood Cliffs, NJ: Prentice Hall. Brown, T.L., Le May, H.E, Bursten, B. & Burdge, J.R. (2003). Chemistry: The central science (9th ed., Spanish). Englewood Cliffs, NJ: Prentice Hall. Brown, T.L., LeMay, H.E., Bursten, B.E., & Murphy, C.J. (2009). Chemistry: The central science (11th ed., Spanish). Englewood Cliffs, NJ: Prentice Hall (Pearson Education). Burman, H. (1968). Principles of general chemistry. Boston: Allyn and Bacon. Burns, R. (1995). Fundamentals of chemistry (Spanish ed.). Englewood Cliffs, NJ: Prentice Hall. Chang, R. (1998). Chemistry (6th ed., Spanish). New York: McGraw-Hill. Chang, R. (2007). Chemistry (9th ed., Spanish). New York: McGraw-Hill. Chang, R. (2010). Chemistry (10th ed., Spanish). New York: McGraw-Hill. Choppin, G.R., Jaffe, B., Summerlin, L. & Jackson, L. (1976). Chemistry. Morristown, NJ: Silver Burdett . Compton, C. (1964). An introduction to chemistry. New  York: Van Nostrand Company. Cracolice, M.S. & Peters, E.I. (2011). Introductory chemistry (4th ed.). Belmont, CA: Brooks/Cole (Cengage Learning). Daub, G.W. & Seese, W. (1996). Basic chemistry. (8th ed., Spanish). Englewood Cliffs, NJ: Prentice Hall. Deming, H. (1957). General chemistry (6th ed., Spanish). New York: Wiley.

190

Appendices

Dickerson, R. E. & Geis, I. (1976). Chemistry, matter and the universe. Menlo Park, CA: Benjamin Cummings. Dickerson, R., Gray, H., Darensbourg, M. & Darensbourg, D. (1984). Chemical principles (4th ed.). Menlo Park, CA: Benjamin Cummings. Dickson, T. (2000). Introduction to chemistry (8th ed.). New York: Wiley. Dillard, C. & Goldberg, D. (1971). Chemistry: Reactions, structure and properties (Spanish ed.). New York: MacMillian. Dull, C., Brooks, W. Y. & Metcalfe, H.C. (1954). Modern chemistry. New York: Henry Colt. Ebbing, D.  D. (1996). General chemistry (5th ed., Spanish). New  York: McGraw-Hill. Fine, L. & Beall, H. (1990). Chemistry for engineers and scientists. Philadelphia: Saunders. Frey, P. (1965). College Chemistry (Spanish ed.). Englewood Cliffs, NJ: Prentice Hall. Gillespie, R.J., Baird, N.C., Humphreys, D. A. & Robinson, E. A (1990). Chemistry (Spanish ed.). Newton: Allyn and Bacon. Goldberg, D. (2001). Fundamentals of chemistry (3rd ed.). New  York: McGraw-Hill. Gray, H. & Haight, G.(1974). Basic principles of chemistry (Spanish ed.). New York: Benjamin. Hein, M. (1990). Foundations of college chemistry (Spanish ed.). Pacific Grove, CA: Brooks/Cole. Hein, M. & Arena, S. (2011). Foundations of college chemistry (13th ed.). Hoboken, NJ: Wiley. Hepler, L. (1968). Chemical principles (Spanish ed.). New  York: Blaisdell Publishing. Hildebrand, J. & Powel, R. (1964). Principles of chemistry (Spanish ed.). New York: MacMillian. Hill, J. (1975). Chemistry for changing times (2nd ed.). Minneapolis: Burgess Publishing Company. Hill, J. & Kolb, D. (1998). Chemistry for changing times (8th ed., Spanish). Upper Saddle River, NJ: Prentice Hall. Hill, J. & Petrucci, R. (1999). General chemistry: An integrated approach (2nd ed.). Upper Saddle River, NJ: Prentice Hall. Hiller, L. & Heber, R. (1960). Principles of chemistry (4th ed., Spanish). New York: McGraw- Hill. Hogg, J., Bickel, C., Nicholson, M. & Wick, H. (1963). Chemistry : A modern approach (Spanish ed.). New York: Van Nostrand. Holum, J.R. (1969). Introduction to principles of chemistry. New York: Wiley. Hutchinson, E. (1968). Chemistry: The elements and their reactions (2nd ed., Spanish). Philadelphia: Saunders. Joesten, M., Johnston, D., Netterville, J. & Wood, J. (1991). World of chemistry. Philadelphia: Saunders.

Appendices

191

Jones, L. & Atkins, P. (2000). Chemistry: Molecules, matter and change (4th ed.). New York: Freeman. Kask, U. (1969). Chemistry: Structure and change of matter (Spanish ed.). New York: McGraw-Hill. Keenan, C., Kleinfelter, D. & Wood, J. (1980). General college chemistry (6th ed., Spanish). New York: Harper & Row. Kneen, W.R., Rogers, M.J.W. & Simpson, P. (1972). Chemistry: Facts, patterns and principles. Reading, MA: Addison-Wesley. Kotz, J.C. & Purcell, K. (1991). Chemistry and chemical reactivity (2nd ed.). Philadelphia: Saunders. Kotz, J.C., & Treichel, P.M. (1999). Chemistry and chemical reactivity (4th ed.). Philadelphia: Saunders. Kotz, J.  C. & Treichel, P.  M. (2003). Chemistry and chemical reactivity (5th ed., Spanish). Pacific Grove, CA: Brooks/Cole. Kotz, J.C., Treichel, P.M. & Townsend, J. (2011). Chemistry and chemical reactivity (7th ed.). Belmont, CA: Brooks/Cole (Cengage Learning). Laidler, K. (1971.) Principles of chemistry (Spanish). New York: Hartcourt, Brace and World. Leddy, E. & Roach, D. (1972). Introductory chemistry. San Francisco: Rinehart Press. Lee, G. & Van Order H. (1965). General Chemistry (Spanish ed.). Philadelphia: Saunders. Lewis, J.R. (1979). College chemsitry. (9th ed., Spanish). New  York: Academic Press. Lippincott, W.T., Garrett, A. & Verhoek, F. (1968). Chemistry: A study of matter (3rd ed.). New York: Wiley. Longo, F. (1974). General chemistry (Spanish ed.). New York: McGraw-Hill. Mahan, B. (1975). University chemistry (3rd ed.). Reading, MA: Addison-Wesley. Mahan, B. & Myers, R. (1987). University chemistry (4th ed., Spanish). Menlo Park, CA: Benjamin Cummings. Malone, L. (2001). Basic concepts of chemistry (6th ed.). New York: Wiley. Malone, L.J., & Dolter, T. (2010). Basic concepts of chemistry (8th ed.). New York: Wiley. Masterton, W.L. & Hurley, C.N. (1997). Chemistry: Principles and reactions (3rd ed.). Philadelphia: Saunders. Masterton, W.L., Slowinski, E.J. & Stanitski, C. L. (1986). Chemical principles (5th ed., Spanish) Philadelphia: Saunders. McMurry, J. & Fay, R. (2001). Chemistry (3rd ed.). Upper Saddle River, NJ: Prentice Hall. McQuarrie, D.A., Rock, P.A., & Gallogly, E.B. (2011). General chemistry (4th ed.). Mill Valley, CA: University Science Books. Metcalfe, H., Williams, J. & Castka, J. (1981). Modern chemistry (Spanish ed.). New York: Holt, Rinehart & Winston. Miller, F.M. (1984). Chemistry: Structure and dynamics. New York: McGraw-Hill.

192

Appendices

Miller, A. & Augustine, L. (1977). Basic chemistry (Spanish ed.). New York: Harper & Row. Moore, J.W., Davies, W.  G. & Collins, R.W. (1978). Chemistry (Spanish ed.). New York: McGraw-Hill. Moore, J. W., Kotz, J.C., Stanitski, C.L., Joesten, M.D. & Wood, J.L. (1998). The chemical world: Concepts and applications (2nd ed., Spanish). Orlando, FL: Harcourt Brace. Moore, J.W., Stanitski, C.L. & Jurs, P. C. (2002). Chemistry: The molecular science. Orlando, FL: Harcourt College Publishers. Moore, J. W., Stanitski, C. L. & Jurs, P. C. (2010). Chemistry: The molecular science (4th ed.). Belmont, CA: Brooks/Cole (Cengage Learning). Mortimer, C. (1979). Chemistry a conceptual Approach (4th ed.). New York: Van Nostrand. Murphy, D. & Rousseau, V. (1980). Foundations of college chemistry (3rd ed.). New York: Wiley. Nebergall, W., Holtzclaw, H. & Robinson, W. (1968). College chemistry (6th ed.). Lexington, MA: Heath. Nebergall, W., Schmidt, F. & Holtzclaw, H. (1968). College chemistry with qualitative analysis. (3rd ed.). Lexington, MA: Heath. O’Connor, R. (1974). Fundamentals of chemistry: A learning systems approach. New York: Harper & Row. Oxtoby, D., Gillis, H.P. & Nachtrieb, N. (1999). Principles of modern chemistry (4th ed.). Philadelphia: Saunders. Oxtoby, D., Nachtrieb, N. & Freeman, W. (1990). Chemistry: Science of change. (2nd ed.). Philadelphia: Saunders. Pauling, L. (1965). General chemistry. (2nd ed., Spanish). San Francisco: Freeman. Petrucci, R.H. & Harwood W.S. (1997). General chemistry: Principles and modern applications (7th ed.). Upper Saddle River, NJ: Prentice Hall. Pierce, J.B. (1970). The chemistry of matter (Spanish ed.).Boston: Houghton Mifflin. Pilar, F. (1979). Chemistry: The universal science. Reading, MA: Addisson-Wesley. Quagliano J.V. & Vallarino, L.M. (1969). Chemistry (3rd ed.). Englewood Cliffs, NJ: Prentice Hall. Redmore, F. (1970). The chemistry of matter. Englewood Cliffs, NJ: Prentice Hall. Reger, D.L., Goode, S.R. & Mercer, E. (1997). Chemistry: Principles and practice. Philadelphia: Saunders. Rosenberg, J.L. & Epstein, L.M. (1991). College chemistry (7th ed., Spanish). New York: McGraw-Hill. Russell, J.  B. & Larena, A.(1988). General chemistry (Spanish ed.). New  York: McGraw-Hill. Russo, S. & Silver, M. (2002). Introductory chemistry (2nd ed.). San Francisco: Benjamin Cummings. Seager, S.L., & Slabaugh (2011). Introductory chemistry for today (7th ed.). Belmont, CA: Brooks/Cole (Cengage Learning).

Appendices

193

Segal, B. (1989). Chemistry: Experiment and theory ( 2nd ed.). New York: Wiley. Sherman, A., Sherman, S.J. & Russikoff, L. (1995). Basic concepts of chemistry (6th ed., Spanish). Boston: Houghton Mifflin. Sienko, M. & Plane, R. (1966). Chemistry: Principles and properties (Spanish ed.). New York: McGraw-Hill. Sienko, M. & Plane, R. (1979). Chemistry: Principles and applications (3rd ed., Spanish). New York: McGraw-Hill. Silberberg, M. (2000). Chemistry: The molecular nature of matter and change (2nd ed.). New York: McGraw-Hill. Sisler, H., Dresdner, R. & Mooney, W. (1980). Chemistry: A systematic approach. New York: Oxford University Press. Sisler, H., Vander Werf, C. & Davidson, A. (1959). General chemistry: A systematic approach. New York: McMillan. Slabaugh, W.H. & Parsons, T.D. (1971). General chemistry (2nd ed.). New  York: Wiley. Smoot, R.C. & Price, J. (1975). Chemistry: A modern course (Spanish ed.). New York: Bell & Howell. Sneed, M., Maynard, L. & Brasted, R. (1960). General college chemistry (Spanish ed.). New York: Van Nostrand. Sorum, C.H (1963). General chemistry (Spanish ed.). Englewood Cliffs, NJ: Prentice Hall. Spencer, J.N., Bodner, G.M., & Rickard L.H. (1999). Chemistry: Structure and dynamics. New York: Wiley. Spencer, J.N., Bodner, G.M., & Rickard, L.H. (2008). Chemistry: Structure and dynamics (4th ed.). New York: Wiley. Stoker, H.S. (1990). Introduction to chemical principles (3rd ed.). New  York: McMillan. Timm, J.A. (1966). General chemistry (Spanish ed.). New York: McGraw-Hill. Tro, N. (2008). Chemistry: A molecular approach. Upper Saddle River, NJ: Prentice Hall (Pearson Education). Turk, A., Meislich, H. Brescia, F., & Arents, F. (1968). Introduction to chemistry. New York: Academic Press. Umland, J.B. (1993). General chemistry. St. Paul, MN: West Publishing Co. Umland, J. & Bellama, J. (1999). General chemistry (3rd ed.). Pacific Grove, CA: Brooks/Cole. Waser, J., Trueblood, K.N. & Knobler, C.  M. (1976). Chem one. New  York: McGraw-Hill. Weller, P.F. & Supple, J.H. (1971). Chemistry: Elementary principles. Reading, MA: Addison Wesley. Whittaker, R. (1964). General Chemistry (Spanish ed.). New  York: Chemical Publishing. Whitten, K.W., Davis, R. & Peck, M.L. (1996). General chemistry (5th ed., Spanish). Philadelphia: Saunders. Whitten, K.W., Davis, R.E., Peck, M.L. & Stanley, G.G. (2009). Chemistry (9th ed.). Belmont, CA: Brooks/Cole (Cengage Learning).

194

Appendices

Whitten, K.W., Gailey, K.D. & Davis, R.E. (1980). General chemistry (3rd ed., Spanish). Philadelphia: Saunders. Wolfe, D. (1988). Introduction to college chemistry (2nd ed.). New  York: McGraw-Hill. Young, L.E. & Porter, C.W. (1958). General chemistry: A first course. Englewood Cliffs, NJ: Prentice Hall. Zumdahl, S. S. (1990). Introductory chemistry: A foundation (Spanish ed.). New York: McGraw-Hill. Zumdahl, S. S. & DeCoste, D.J. (2010). Introductory chemistry (7th ed.). Belmont, CA: Brooks/Cole (Cengage Learning). Zumdahl, S.S., & Zumdahl, S.A. (2010). Chemistry (8th ed.). Belmont, CA: Brooks/ Cole (Cengage Learning).

 ppendix 10: List of General Physics Textbooks Evaluated A in Different Studies of This Book (n = 103) Acosta, V., Cowan, C.L., & Graham, B.J. (1973). Essentials of Modern Physics (1st ed.). New York, NY: Harper & Row. Arfken, G.B., Griffing, D.F., Kelly, D.C. & Priest, J. (1989). University Physics Volume II (2nd ed.). San Diego, CA: Harcourt Brace Jovanovich. Atkins, K.R. (1976). Physics (3rd ed.). New York, NY: Wiley. Atkins, K.R. (1965). Physics (1st ed.). New York, NY: Wiley. Beiser, A. (1991). Physics (5th ed.). Reading, MA: Addison-Wesley. Beiser, A. (1979). Modern Technical Physics (3rd ed.). Menlo Park, CA: Benjamin/ Cummings Pub. Co.. Blackwood, O. & Kelly, W. (1955). General Physics (2nd ed.). New  York, NY: Wiley. Blanchard, C.H., et. al. (1958). Introduction to Modern Physics (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Borowitz, S. & Bornstein, L.A. (1968). A Contemporary View of Elementary Physics (1st ed.). New York, NY: McGraw-Hill. Bueche, F. (1977). Principles of Physics (3rd ed.). New York, NY: McGraw-Hill. Bueche, F. (1969). Introduction to Physics for Scientists and Engineers (1st ed.). New York, NY: McGraw-Hill. Bueche, F. (1965). Principles of Physics (1st ed.). New York, NY: McGraw-Hill. Bueche, F.J. (1986). Introduction to Physics for Scientists and Engineers (4th ed.). New York, NY: McGraw-Hill. Bueche, F.J. (1980). Introduction to Physics for Scientists and Engineers (3rd ed.). New York, NY: McGraw-Hill. Bueche, F.J. (1975). Introduction to Physics for Scientists and Engineers (2nd ed.). New York, NY: McGraw-Hill.

Appendices

195

Bueche, F. & Wallach, D.L. (1994). Technical Physics (4th ed.). New  York, NY: John Wiley. Coletta, V.P. (1995). College Physics (1st ed.). St. Louis, MO: Mosby. Cooper, L.N. (1992). Physics: Structure and Meaning (1st ed.). Hanover, NH: University Press of New England. Cooper, L.N. (1968). An Introduction to the Meaning and Structure of Physics (1st ed.). New York, NY: HarperCollins. Copeland, P.L. & Bennett, W.E. (1961). Elements of Modern Physics (1st ed.). New York, NY: Oxford University Press. Cutnell, J.D. & Johnson, K.W. (2007). Physics (7th ed.). Hoboken, NJ: Wiley. Dull, C.E., Metcalfe, H.H. & Williams, J.E. (1964). Modern Physics (1st ed.). New York, NY: Holt, Rinehart and Winston. Durbin, F.M. (1955). Introduction to Physics (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Eisberg, R.M. & Lerner, L.S. (1981). Physics Foundations and Applications Combined Volume (1st ed.). New York, NY: McGraw-Hill. Fishbane, P.M., Gasiorowicz, S.G. & Thornton, S.T. (2005). Physics for Scientists and Engineers with Modern Physics (3rd ed.). Upper Saddle River, NJ: Pearson Prentice-Hall. Fishbane, P.M., Gasiorowicz, S.G. & Thornton, S.T. (1993). Physics for Scientists and Engineers (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Gamow, G. & Cleveland, J.M. (1960). Physics: Foundations and Frontiers (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Gettys, W.E., Keller, F.J. & Skove, M.J. (1989). Physics Classical and Modern (1st ed.). New York, NY: McGraw-Hill. Giancoli, D.C. (1998). Physics Principles with Applications (5th ed.). Upper Saddle River, NJ: Prentice-Hall. Giancoli, D.C. (1984). General Physics (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Greenberg, L.H. (1978). Physics with Modern Applications (1st ed.). Philadelphia, PA: Saunders. Greene, E.S. (1962). Principles of Physics (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Hagelberg, M.P. (1973). Physics An Introduction for Students of Science and Engineering (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Halliday, D. & Resnick, R. (1981). Fundamentals of Physics (2nd ed.). New York, NY: Wiley. Halliday, D. & Resnick, R. (1974). Fundamentals of Physics (1st ed.). New York, NY: Wiley. Halliday, D. & Resnick, R. (1962). Physics for Students of Science and Engineering Combined Edition (1st ed.). New York, NY: Wiley. Halliday, D., Resnick, R. & Walker, J. (2008). Fundamentals of Physics (8th ed.). New York, NY: Wiley. Halliday, D., Resnick, R. & Walker, J. (1997). Fundamentals of Physics Part 5 (5th ed.). New York, NY: Wiley.

196

Appendices

Halliday, D., Resnick, R. & Walker, J. (1993). Fundamentals of Physics (4th ed.). New York, NY: Wiley. Hazen, W.E., & Pidd, R.W. (1965). Physics (1st ed.). Reading, MA: Addison-Wesley. Hecht, E. (1998). Physics: Algebra/Trig (2nd ed.). Pacific Grove, CA: Brooks/Cole. Hecht, E. (2003). Physics: Algebra/Trig (3rd ed.). Pacific Grove, CA: Thomson Brooks/Cole. Hooper, H.O. & Gwynne, P. (1980). Physics and the Physical Perspective (2nd ed.). San Fransisco, CA: Harper & Row. Jones, E.R. & Childers, R.L. (1990). Contemporary College Physics (1st ed.). Reading, MA: Addison-Wesley. Knight, R.D. (2004). Physics for Scientists and Engineers: a strategic approach (1st ed.). San Fransisco, CA: Pearson Addison-Wesley. Lea, S.M. & Burke, J.R. (1997). Physics: The Nature of Things (1st ed.). Pacific Grove, CA: Brooks/Cole Pub. Co.. Lindsay, R.B. & Margenau, H. (1957). Foundations of Physics (1st ed.). New York, NY: Dover Publications Inc.. Marion, J.B. & Hornyak, W.F. (1982). Physics for Science and Engineering Part 2 (1st ed.). Philidelphia, PA: Saunders College Pub.. Marshall, J.S. & Pounder, E.R. (1957). Physics (1st ed.). Toronto, ON: The Macmillan Company. McCormick, W.W. (1965). Fundamentals of College Physics (1st ed.). New York, NY: Macmillan. McGervey, J.D. (1983). Introduction to Modern Physics (2nd ed.). New York, NY: Academic Press. Melissinos, A.C. & Lobkowicz, F. (1975). Physics for Scientists and Engineers Volume II (1st ed.). Philidelphia, PA: W.B. Saunders Company. Miller, F. (1972). College Physics (3rd ed.). New  York, NY: Harcourt Brace Jovanovich. Morgan, J. (1964). Introduction to University Physics Volume II (1st ed.). Boston, MA: Allyn and Bacon. Nolan, P.J. (1995). Fundamentals of College Physics Volume Two (2nd ed.). Dubuque, IA: Wm. C. Brown Publishers. Ohanian, H.C. (1989). Physics (2nd ed.). New York, NY: Norton. Ohanian, H.C. & Markert, J.T. (2007). Physics for Scientists and Engineers Volume III (3rd ed.). New York, NY: W.W. Norton & Company. Orear, J. (1979). Physics (1st ed.). New York, NY: Macmillan. Orear, J. (1967). Fundamental Physics (2nd ed.). New York, NY: Wiley. Orear, J. (1961). Fundamental Physics (1st ed.). New York, NY: Wiley. Ostdiek, V.J. & Bord, D.J. (1991). Inquiry in to Physics (2nd ed.). St Paul, MN: West Publishing Company. Physical Science Study Committee (1968). College Physics (1st ed.). Boston, MA: Raytheon Education Company. Priestley, H. (1958). Introductory Physics a historical approach (1st ed.). Boston, MA: Allyn and Bacon.

Appendices

197

Radin, S.H. & Folk, R.T. (1982). Physics for Scientists and Engineers (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Reese, R.L. (2000). University Physics (1st ed.). Pacific Grove, CA: Brooks/Cole Pub. Co.. Richards, J.A., Sears, F.W., Wehr, M.R. & Zemansky, M.W. (1960). Modern University Physics (1st ed.). Reading, MA: Addison-Wesley. Richtmyer, F.K., Kennard, E.H., Lauritsen, T. (1955). Introduction to Modern Physics (5th ed.). New York, NY: McGraw-Hill. Sears, F.W. & Zemansky, M.W. (1970). University Physics (4th ed.). Reading, MA: Addison-Wesley. Sears, F.W. & Zemansky, M.W. (1964). University Physics (3rd ed.). Reading, MA: Addison-Wesley. Sears, F.W. & Zemansky, M.W. (1953). College Physics (2nd ed.). Cambridge, MA: Addison-Wesley. Sears, F.W., Zemansky, M.W. & Young, H.D. (1991). College Physics (7th ed.). Reading, MA: Addison-Wesley. Sears, F.W., Zemansky, M.W. & Young, H.D. (1982). University Physics (6th ed.). Reading, MA: Addison-Wesley. Sears, F.W., Zemansky, M.W. & Young, H.D. (1974). College Physics (4th ed.). Reading, MA: Addison-Wesley. Semat, H. (1957). Fundamentals of Physics (3rd ed.). New York, NY: Holt, Rinehart and Winston. Serway, R.A. (1990). Physics for Scientists and Engineers Volume II (3rd ed.). Fort Worth: Saunders College Pub. Serway, R.A. & Beichner, R.J. (2000). Physics for Scientists and Engineers Volume 2 (5th ed.). Pacific Grove, CA: Brooks/Cole-Thomson Learning. Serway, R.A. & Jewett, J.W. (2008). Physics for Scientists and Engineers (7th ed.). Belmont, CA: Thomson. Serway, R.A., Faughn, J.S., Vuille, C., & Bennett, C.A. (2006). College Physics (7th ed.). Pacific Grove, CA: Thomson Brooks/Cole. Serway, R.A. & Faughn, J.S. (1995). College Physics (4th ed.). Fort Worth: Saunders College Publishing. Shortley, G. & Williams, D. (1971). Elements of Physics for Students of Science and Engineering (5th ed.). Englewood Cliffs, NJ: Prentice-Hall. Shortley, G. & Williams, D. (1959). Principles of College Physics (1st ed.). Englewood Cliffs, NJ: Prentice-Hall. Shortley, G. & Williams, D. (1950). Physics Fundamental Principles for Students of Science and Engineering: Volume II (1st ed.). New York, NY: Prentice-Hall. Tillery, B.W. (1996). Physical Science (3rd ed.). Dubuque, IA: Wm. C. Brown. Tilley, D.E. (1976). University Physics for Science and Engineering (1st ed.). Menlo Park, CA: Cummings Publishing. Tipler, P.A. (1991). Physics for Scientists and Engineers (3rd ed.). New York, NY: Worth. Tipler, P.A. (1987). College Physics (1st ed.). New York, NY: Worth. Tipler, P.A. (1978). Modern Physics (1st ed.). New York, NY: Worth.

198

Appendices

Tippens, P.E. (2007). Physics (7th ed.). Boston, MA: McGraw-Hill. Touger, J. (2006). Introductory Physics Building Understanding (1st ed.). Hoboken, NJ: Wiley. Urone, P.P. (2001). College Physics (2nd ed.). Pacific Grove, CA: Brooks/Cole. Van Name, F.  W. (1962). Modern Physics (2nd ed.). Englewood Cliffs, NJ: Prentice-Hall. Walker, J.S. (2002). Physics (1st ed.). Upper Saddle River, NJ: Prentice-Hall. Walker, J.S. (2007). Physics Volume II (3rd ed.). Upper Saddle River, NJ: Pearson Prentice-Hall. Weidner, R.T. & Browne, M.E. (1985). Physics (1st ed.). Boston, MA: Allyn and Bacon. Weidner, R.T. & Sells, R.L. (1968). Elementary Modern Physics (1st ed.). Boston, MA: Allyn and Bacon. White, H.E. (1969). Introduction to College Physics (1st ed.). New York, NY: Van Nostrand Reinhold. Williams, D. & Spangler, J. (1981). Physics for Science and Engineering (1st ed.). New York, NY: Van Nostrand. Wilson, J.D., Buffa, A.J. & Lou, B. (2007). College Physics (6th ed.). Upper Saddle River, NJ: Pearson Prentice-Hall. Wolfson, R. & Pasachoff, J.M. (1995). Physics for Scientists and Engineers (2nd ed.). New York, NY: HarperCollins. Young, H.D. (1992). University Physics Extended Version with Modern Physics (8th ed.). Reading, MA: Addison-Wesley. Young, H.D. & Freedman, R.A. (1996). University Physics (9th ed.). Reading, MA: Addison-Wesley. Young, H.D. & Gellar, R.M. (2007). Sears & Zemansky’s College Physics (8th ed.). San Fransisco, CA: Pearson Addison-Wesley. Zafiratos, C. (1976). Physics (1st ed.). New York, NY: Wiley.

References

Aarriassecq, I., & Greca, I. M. (2007). Approaches to the teaching of special relativity theory in high school and university textbooks. Science & Education, 16(1), 65–86. Abd-El-Khalick, F., Belarmino, J. J., Brunner, J. L., Le, A.-P., Myers, J. Y., Summers, R. G., et al. (2017). A longitudinal analysis of the extent and manner of representations of nature of science in U.S. high school chemistry, biology, and physics textbooks. In C. V. McDonald & F. Abd-­ El-­Khalick (Eds.), Representations of nature of science in school science textbooks: A global perspective (pp. 20–60). New York: Routledge. Abd-El-Khalick, F., Myers, J. Y., Summers, R., Brunner, J., Waight, N., Wahbeh, N., et al. (2017). A longitudinal analysis of the extent and manner of representations of nature of science in U.S. high school biology and physics textbooks. Journal of Research in Science Teaching, 54(1), 82–120. Abrams, G. S., & Perl, M. L. (1979). Measurement of the branching fraction for tau ➔ rho nu. Physics Review Letters, 43, 1555–1558. Achinstein, P. (1987). Scientific discovery and Maxwell’s kinetic theory. Philosophy of Science, 54, 409–434. Achinstein, P. (2004). Science rules: A historical introduction to scientific methods. Baltimore, MD: Johns Hopkins University Press. Achinstein, P. (2013). Evidence and method: Scientific strategies of Isaac Newton and James Clerk Maxwell. New York: Oxford University Press. Agassi, J. (1975). Genius in science. Philosophy of the Social Sciences, 5(2), 145–161. Agassi, J.  (2014). Popper and his popular critics: Thomas Kuhn, Paul Feyerabend and Imre Lakatos. Dordrecht, the Netherlands: Springer. Alters, B. J. (1997). Whose nature of science? Journal of Research in Science Teaching, 34(1), 39–55. American Association for the Advancement of Science (AAAS). (1993a). Science for all Americans. New York: Oxford University Press. American Association for the Advancement of Science (AAAS). (1993b). Benchmarks for science literacy: Project 2061. Washington, DC: Oxford University Press. Arabatzis, T., & Gavroglu, K. (2015). That the Michelson-Morley experiment paved the way for the special theory of relativity. In R. L. Numbers & K. Kampourakis (Eds.), Newton’s apple and other myths about science (pp. 149–156). Cambridge, MA: Harvard University Press. Arrhenius, S. (1912). Theories of solutions. New Haven, CT: Yale University Press. Atkinson, P., & Delamont, S. (2005). Analytic perspectives. In N.  K. Denzin & Y.  S. Lincoln (Eds.), The Sage handbook of qualitative research (3rd ed., pp.  821–840). Thousand Oaks, CA: Sage. © Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3

199

200

References

Bailin, S. (1990). Creativity, discovery, and science education: Kuhn and Feyerabend revisited. Interchange, 21(3), 34–44. Barnes, B., Bloor, D., & Henry, J. (1996). Scientific knowledge: A sociological analysis. Chicago: University of Chicago Press. Bauer, H.  H. (1994). Scientific literacy and the myth of the scientific method. Champaign, IL: University of Illinois Press. Bazghandi, P., & Hamrah, S. Z. (2011). The principles of teaching science based on the ideas of Feyerabend regarding the nature of science and the manner of its expansion. Procedia Social and Behavioral Sciences Elsevier, 29, 969–975. Ben-Ari, M. (2005). Situated learning in “this high-technology world”. Science & Education, 14(3–5), 367–376. Bensaude-Vincent, B. (1986). Mendeleev’s periodic system of chemical elements. British Journal for the History of Science, 19, 3–17. Bird, A. (2010). The historical turn in the philosophy of science. In S. Psillos & M. Curd (Eds.), The Routledge companion to philosophy of science (pp. 67–77). London: Routledge. Bjelic, D., & Lynch, M. (1992). The work of a (scientific) demonstration: Respecifying Newton’s and Goethe’s theories of prismatic color. In G. Watson & R. M. Seiler (Eds.), Text in context: Contributions to ethnomethodology (pp. 52–78). Newbury Park, CA: Sage. Bloor, D. (1976). Knowledge and social imagery. London: Routledge & Kegan Paul. Bloor, D. (1978). Polyhedra and the abominations of Leviticus. British Journal for the History of Science, 11, 245–272. Bohr, N. (1913). On the constitution of atoms and molecules. Part I. Philosophical Magazine, 26(series 6), 1–25. Bransford, J.  D., & Donovan, S.  M. (2005). Scientific inquiry and how people learn. In S.  M. Donovan & J. D. Bransford (Eds.), How students learn: History, mathematics, and science in the classroom (pp. 397–420). Washington, DC: The National Academies Press. Bray, W. C., & Branch, G. E. K. (1913). Valence and tautomerism. Journal of American Chemical Society, 35, 1440–1447. Brito, A., Rodríguez, M. A., & Niaz, M. (2005). A reconstruction of development of the periodic table based on history and philosophy of science and its implications for general chemistry textbooks. Journal of Research in Science Teaching, 42, 84–111. Brock, W. H. (2008). Joseph Priestley, enlightened experimentalist. In D. L. Wykes & I. Rivers (Eds.), Joseph Priestley, scientist, philosopher, and theologian (pp.  49–79). Oxford, UK: Oxford University Press. Brown, M. J. (2016). The abundant world: Paul Feyerabend’s metaphysics of science. Studies in History and Philosophy of Science, 57, 142–154. Brown, M. J., & Kidd, I. J. (2016). Introduction: Reappraising Paul Feyerabend. Studies in History and Philosophy of Science, 57, 1–8. Brush, S. G. (1974). Should the history of science be rated X? Science, 183(4130), 1164–1172. Brush, S. G. (1976). The kind of motion we call heat: A history of the kinetic theory of gases in the 19th century. New York: North-Holland. Brush, S. G. (1978). Why chemistry needs history — And how it can get some. Journal of College Science Teaching, 7, 288–291. Brush, S. G. (1979). Comments on ‘On the distortion of the history of science in science education’. Science Education, 63(2), 277–278. Brush, S. G. (1996). The reception of Mendeleev’s periodic law in America and Britain. Isis, 87, 595–628. Brush, S. G. (2000). Thomas Kuhn as a historian of science. Science & Education, 9, 39–58. Bunge, M. (2003). Twenty-five centuries of quantum physics: From Pythagoras to us, and from subjectivism to realism. Science & Education, 12(5–6), 445–466. Campbell, D. T. (1988a). Can we be scientific in applied social science? In E. S. Overman (Ed.), Methodology and epistemology for social science (pp.  315–333). Chicago: University of Chicago.

References

201

Campbell, D.  T. (1988b). Qualitative knowing in action research. In E.  S. Overman (Ed.), Methodology and epistemology for social science (pp.  360–376). Chicago: University of Chicago Press. Cartwright, N. (1983). How the laws of physics lie? Oxford: Clarendon Press. Cartwright, N. (1999). The dappled world: A study of the boundaries of science. Cambridge, UK: Cambridge University Press. Cavalli-Sforza, M., Goggi, G., Mantovani, G. C., Piazzoli, A., Rossini, B., Scannicchio, D., et al. (1976). Anomalous production of high-energy muons in e+e- collisions at 4.8 GeV. Physics Review Letters, 36, 558–561. Cavazos, L., Hazelwood, C.  C., Howes, E.  V., Kurth, L., Lane, P., Markham, L., et  al. (1998). Response to guest editorial: The WISE group: Connecting activism, teaching, and research. Journal of Research in Science Teaching, 35(4), 341–344. Chalmers, A. F. (1982). What is this thing called science? (2nd ed.). St. Lucia, QLD: University of Queensland Press. Chalmers, A. F. (1985). Galileo’s telescopic observations of Venus and Mars. British Journal for the Philosophy of Science, 36, 175–183. Charmaz, K. (2005). Grounded theory in the 21st century: Applications for advancing social justice studies. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (3rd ed., pp. 507–535). Thousand Oaks, CA: Sage Publications. Clark, P. (1976). Atomism and thermodynamics. In C. Howson (Ed.), Method and appraisal in the physical sciences: The critical background to modern science, 1800–1905 (pp. 41–105). Cambridge, MA: Cambridge University Presss. Clausius, R. (1857). On the nature of the motion we call heat. Philosophical Magazine, 14, 108–127. Cobern, W.  W. (1994). Point: Belief, understanding, and the teaching of evolution. Journal of Research in Science Teaching, 31(5), 583–590. Cobern, W. W. (2004). Apples and oranges: A rejoinder to Smith and Siegel. Science & Education, 13(6), 583–589. Collins, H. M. (1981). Stages in the empirical programme of relativism. Social Studies of Science, 11, 3–10. Collins, H. M. (1985). Changing order: Replication and induction in scientific practice. London: Sage. Collins, H. M. (2000). On beyond 2000. Studies in Science Education, 35, 169–173. Collins, H. M., & Evans, R. (2002). The third wave of science studies: Studies of expertise and experience. Social Studies of Science, 32(2), 235–296. Collins, H. M., & Evans, R. (2007). Rethinking expertise. Chicago: University of Chicago Press. Collins, H. M., & Pinch, T. (1998). The Golem: What you should know about science (2nd ed.). Cambridge, MA: Cambridge University Press. Conant, J. B. (1951). Science and common sense. New Haven, CT: Yale University Press. Cooper, L.  N. (1970). An introduction to the meaning and structure of physics (short edn). New York: Harper & Row. Cooper, L. N. (1992). Physics: Structure and meaning. Hanover, NH: University Press of New England. Cordero, A. (2001). Scientific culture and public education. Science & Education, 10(1-2), 71–83. Cowan, C. L., et al. (1956). Detection of the free neutrino: A confirmation. Science, 124, 103–104. Crease, R. M. (2002). Critical point: The most beautiful experiment. Physics World, 15(9), 19–20. Crowther, J. G. (1910). Proceedings of the Royal Society (Vol. lxxxiv). London: Royal Society. Cushing, J. T. (1989). The justification and selection of scientific theories. Synthese, 78, 1–24. Cushing, J. T. (1998). Philosophical concepts in physics: The historical relation between philosophy and scientific theories. Cambridge: Cambridge University Press. Darrigol, O. (2009). A simplified genesis of quantum mechanics. Studies in History and Philosophy of Modern Physics, 40, 151–166. Daston, L., & Galison, P. L. (2007). Objectivity. New York: Zone Books.

202

References

Davisson, C., & Germer, L. H. (1927). Diffraction of electrons by a crystal of Nickel. Physical Review, 30(6), 705–740. De Berg, K. C. (2003). The development of the theory of electrolytic dissociation: A case study of a scientific controversy and the changing nature of chemistry. Science & Education, 12, 397–419. De Berg, K. C. (2014). Teaching chemistry for all its worth: The interaction between facts, ideas, and language in Lavoisier’s and Priestley’s chemistry practice: The case of the study of the composition of air. Science & Education, 23(10), 2045–2068. De Broglie, L. (1922). Journal de Physique, 3 (Series VI), 422. De Broglie, L. (1923a). Ondes et quanta. Comptes Rendus, 177, 507–510, 548–550, 630–632. De Broglie, L. (1923b). Waves and quanta. Nature, 112, 540. De Broglie, L. (1924). A tentative theory of light quanta. Philosophical Magazine, 47(Series 6), 446–458. De Milt, C. (1951). The congress at Karlsruhe. Journal of Chemical Education, 28, 421–425. Denzin, N. K., & Lincoln, Y. S. (2005). Introduction: The discipline and practice of qualitative research. In N. K. Denzin & Y. S. Lincoln (Eds.), The Sage handbook of qualitative research (3rd ed., pp. 1–32). Thousand Oaks, CA: Sage. Dobzhansky, T. (1973). Nothing in biology makes sense except in the light of evolution. American Biology Teacher, 35, 125–129. Drago, A. (1994). Mach’s thesis: Thermodynamics as the basic theory for physics teaching. Science & Education, 3(2), 189–198. Duhem, P. (1914). The aim and structure of physical theory (2nd ed., & P.  P. Wiener, Trans.). New York: Atheneum (First published in 1905). Dupré, J.  (1993). The disorder of things: Metaphysical foundations for the disunity of science. Harvard, MA: Harvard University Press. Earman, J., & Glymour, C. (1980). Relativity and eclipses: The British eclipse expeditions of 1919 and their predecessors. Historical Studies in the Physical Sciences, 11(1), 49–85. Eflin, J. T., Glennan, S., & Reisch, G. (1999). The nature of science: A perspective from the philosophy of science. Journal of Research in Science Teaching, 36(1), 107–116. Einstein, A. (1905). Über einen erzeugung und verwandlung des lichtes betreffenden heuristischen gesichtspunkt. Annalen de Physik, 17, 132–148. Einstein, A. (1916). Zur quantentheorie der strahlung. Zürich Mitteilungen, 18, 47–62. Einstein, A. (1926). Investigations on the theory of the Brownian movement (R. Furth, (Ed.), A. D. Cowper, Trans.). London: Methuen. Reprinted: New York, Dover, 1956. Einstein, A. (1949). Remarks concerning the essays brought together in this cooperative volume (P. Schilpp, Trans.). In P. Schilpp (Ed.), Albert Einstein: Philosopher-scientist (pp. 65–688). La Salle, PA: Open Court. Erickson, F.  E. (1986). Qualitative methods in research on teaching. In M.  C. Wittrock (Ed.), Handbook of research on teaching (3rd ed., pp. 119–161). New York: Macmillan. Farrell, R. P. (2003). Feyerabend and scientific values. Tightrope-Walking rationality. Dordrecht, the Netherlands: Kluwer. Feldman, G. J., & Perl, M. L. (1977). Inclusive anomalous muon production in e+e- annihilation. Physics Review Letters, 38, 117–120. Feyerabend, P. K. (1962/1981). Explanation, reduction and empiricism. Minnesota Studies in the Philosophy of Science, 3, 28–97. Feyerabend, P. K. (1963). How to be a good empiricist. In B. Baumrin (Ed.), Delaware seminar in philosophy of science (Vol. 2, pp. 3–40). New York: Interscience. Feyerabend, P. K. (1968). Science, freedom, and the good life. Philosophical Forum, 1(2), 127–135. Feyerabend, P.  K. (1970a). Against method: Outline of an anarchistic theory of knowledge. In M.  Radner & S.  Winokur (Eds.), Minnesota studies in the philosophy of science (Vol. IV, pp. 17–130). Mineapolis, MN: University of Minnesota Press. Feyerabend, P. K. (1970b). Problems of empiricism, Part II. In R. G. Colodny (Ed.), The nature and function of scientific theories (pp. 275–353). Pittsburgh, PA: University of Pittsburgh Press.

References

203

Feyerabend, P. K. (1970c). Consolations for the specialist. In I. Lakatos & A. Musgrave (Eds.), Criticism and the growth of knowledge (pp. 197–230). Cambridge, UK: Cambridge University Press. Feyerabend, P. K. (1970d). Classical empiricism. In R. E. Butts & J. W. Davis (Eds.), The methodological heritage of Newton (pp. 150–170). Oxford, UK: Basil Blackwell. Feyerabend, P. K. (1974/1975b/1988). How to defend society against science. In E. D. Klemke, R. Hollinger, & A. D. Kline (Eds.), Introductory readings in the philosophy of science. Buffalo, NY: Prometheus. Feyerabend, P. K. (1975a). Against method. Outline of an anarchist theory of knowledge. Londond: New Left Books. Feyerabend, P.  K. (1975b/1988). How to defend society against science. Radical Philosophy, 11(1), 3-9. Feyerabend, P.  K. (1975c). “Science”: The myth and its role in society. Afterword: Theses on anarchism. Inquiry, 18, 167–181. Feyerabend, P. K. (1978/1982). Science in a free society. London: Verso. Feyerabend, P. K. (1980). How to defend society against science. In E. D. Klemke, R. Hollinger, & A.  D. Kline (Eds.), Introductory readings in the philosophy of science. Buffalo, NY: Prometheus. Feyerabend, P. K. (1987). Farewell to reason. London: Verso. Feyerabend, P.  K. (1991a). Concluding unphilosophical conversation. In G.  Munévar (Ed.), Beyond reason: Essays on the philosophy of Paul Feyerabend (pp. 487–527). Dordrecht, the Netherlands: Kluwer. Feyerabend, P. K. (1991b). Three dialogues on knowledge. Oxford, UK: Blackwell. Feyerabend, P. K. (1993). Against method. Outline of an anarchistic theory of knowledge (3rd Rev and enlarged edn). New York: Verso. Feyerabend, P.  K. (1994). Potentially every culture is all cultures. Common Knowledge, 2(3), 16–22. Feyerabend, P. K. (1995). Killing time (autobiography). Chicago: University of Chicago Press. Feyerabend, P. K. (1999a). Ambigüedad y armonía. Barcelona, Spain: Ediciones Paidós (Based on Lectures delivered at the University of Trent in 1992, published in Italian in 1996 and English in 2011). Feyerabend, P. K. (1999b). Conquest of abundance: A tale of abstraction versus the richness of being. Chicago: University Of Chicago Press. Feyerabend, P. K. (2011). The tyranny of science. Cambridge, UK: Polity Press (Based on Trent lectures delivered in 1992). Finocchiario, M. A. (2011). A Galilean approach to the Galileo affair. Science & Education, 20(1), 51–66. Finocchiario, M. A. (2019). On trial for reason: Science, religion, and culture in the Galileo affair. New York: Oxford University Press. Finocchiaro, M.  A. (Trans.: Ed.). (1989). The Galileo affair: A documentary history. Berkeley, CA: University of California Press. Finocchiaro, M. A. (Trans.: Ed.). (1997). Galileo on the world systems: A new abridged translation and guide. Berkeley, CA: University of California Press. Finocchiaro, M.  A. (2010). Defending Copernicus and Galileo: Critical reasoning in the two affairs. Dordrecht, the Netherlands: Springer. Fosnot, C. T. (1993). Rethinking science education: A defense of Piagetian constructivism. Journal of Research in Science Teaching, 30(9), 1189–1201. Fuller, S. (1992). Social epistemology and the research agenda of science studies. In A. Pickering (Ed.), Science as practice and culture (pp. 390–428). Chicago: University of Chicago Press. Galileo, G. (1960). On motion and on mechanics. Madison, WI: University of Wisconsin Press. Galileo, G. (1967). Dialogue concerning the two chief world systems (S. Drake, Trans.). Berkeley, CA: University of California Press. Galison, P. (2015). Email to author, Nov. 17.

204

References

Gattei, S. (2016). Feyerabend, truth and relativisms: Footnotes to the Italian debate. Studies in History and Philosophy of Science, 57, 87–95. Garber, E., Brush, S.  G., & Everitt, C.  W. F. (Eds.). (1986). Maxwell on molecules and gases. Cambridge, MA: MIT Press. Gavroglu, K. (2000). Controversies and the becoming of physical chemistry. In P.  Machamer, M. Pera, & A. Baltas (Eds.), Scientific controversies: Philosophical and historical perspectives (pp. 177–198). New York: Oxford University Press. Geelan, D. (2006). Undead theories: Constructivism, eclecticism and research in education. Rotterdam, the Netherlands: Sense Publishers. Geelan, D. (2019). Email to author, March 2, reproduced with permission. Geelan, D. R. (1997). Epistemological anarchy and the many forms of constructivism. Science & Education, 6(1–2), 15–28. Geiger, H., & Marsden, E. (1909). On a diffuse reflection of the alpha particles. In Proceedings of the Royal Society (Vol. lxxxii). London: Royal Society. Giere, R. N. (1988). Explaining science. Chicago: University of Chicago Press. Giere, R. N. (1999). Science without laws. Chicago: University of Chicago Press. Giere, R. N. (2006a). Scientific perspectivism. Chicago: University of Chicago Press. Giere, R.  N. (2006b). Perspectival pluralism. In S.  H. Kellert, H.  E. Longino, & C.  K. Waters (Eds.), Scientific pluralism (pp. 26–41). Minneapolis, MN: University of Minnesota Press. Giere, R.  N. (2010). Naturalism. In S.  Psillos & M.  Curd (Eds.), The Routledge companion to philosophy of science (pp. 213–223). London: Routledge. Giere, R. N. (2016). Feyerabend’s perspectivism. Studies in History and Philosophy of Science, 57, 137–141. Gillispie, C. C. (1960). The edge of objectivity. Princeton, NJ: Princeton University Press. Glaser, B. G., & Strauss, A. L. (1967). The discovery of grounded theory. Chicago: Aldine. Godfrey-Smith, P. (2003). Theory and reality: An introduction to the philosophy of science. Chicago: University of Chicago Press. Good, R.  G. (1993). Editorial: The slippery slopes of postmodernism. Journal of Research in Science Teaching, 30(5), 427. Good, R. G. (2001). Habits of mind associated with science and religion: Implications for science education. Paper presented at the tri-annual meeting of the International History, Philosophy and Science Teaching Group, Denver, CO. Gordin, M. D. (2004). A well-ordered thing: Dmitrii Mendeleev and the shadow of the periodic table. New York: Basic Books. Gould, S. J. (1977). Ever since Darwin. New York: Norton. Gould, S.  J. (1980). The promise of paleobiology as a nomothetic, evolutionary discipline. Paleobiology, 6(1), 96–118. Gower, B. (1997). Scientific method: An historical and philosophical introduction. London: Routledge. Grosser, M. (1962). The discovery of Neptune. Cambridge, MA: Harvard University Press (Reprint, New York: Dover, 1979). Hacking, I. (1983). Representing and intervening. Cambridge, UK: Cambridge University Press. Hanson, N. R. (1958). Patterns of discovery: An inquiry into the conceptual foundations of science. Cambridge, UK: Cambridge University Press. Hanson, N. R. (1964). On the structure of physical knowledge. In S. Elam (Ed.), Education and the structure of knowledge (pp. 148–187). Chicago: Rand McNally & Company. Hanson, N.  R. (1969). Perception and discovery: An introduction to scientific inquiry. San Francisco: Freeman, Cooper and Co.. Harvard Project Physics Course. (1975). Text (2nd ed.). New York: Holt, Rinehart & Winston. Hattiangadi, J.  N. (1977). The crises in methodology: Feyerabend. Philosophy of the Social Sciences, 7, 289–302.

References

205

Heering, P., & Höttecke, D. (2014). Historical-investigative approaches in science teaching. In M. R. Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. II, pp. 1473–1502). Dordrecht, the Netherlands: Springer. Heilbron, J. L. (1981a). Rutherford-Bohr atom. American Journal of Physics, 49, 223–231. Heilbron, J. L. (1981b). Historical studies in the theory of atomic structure. New York: Arno Press. Heilbron, J. L., & Kuhn, T. (1969). The genesis of the Bohr atom. Historical Studies in the Physical Sciences, 1, 211–290. Hertz, H. (1883). Versuche über die Glimmentladung. Annalen der Physik, 19, 782–816. Hertz, H. (1887). Ueber einen Einfluss des Ultravioletten Lichtes auf die Elektrische Entladung. Annalen der Physik, 31, 983–1000. Hildebrand, G. M. (1998). Disrupting hegemonic writing practices in school science: Contesting the right way to write. Journal of Research in Science Teaching, 35(4), 345–362. Hodson, D. (1992). Assessment of practical work. Science & Education, 1(2), 115–144. Hodson, D. (1998). Science fiction: The continuing misrepresentation of science in the school curriculum. Curriculum Studies, 6(2), 191–216. Hodson, D. (2009). Teaching and learning about science: Language, theories, methods, history, traditions and values. Rotterdam, the Netherlands: Sense Publishers. Hodson, D. (2014). Nature of science in the science curriculum: Origin, development, implications and shifting emphases. In M. R. Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. II, pp. 911–970). Dordrecht, the Netherlands: Springer. Hoffmann, R. (2012). In J. Kovac & M. Weisberg (Eds.), Roald Hoffmann on the philosophy, art, and science of chemistry. Oxford, UK: Oxford University Press. Hoffmann, R. (2014). The tensions of scientific storytelling: Science depends on compelling narratives. American Scientist, 102, 250–253. Hoffmann, R. (2016). Email sent to author, 23 February, 2016. Hoffmann, R., Shaik, S., & Hiberty, P. C. (2003). A conversation on VB vs MO theory: A never-­ ending rivalry? Accounts of Chemical Research, 36(10), 750–756. Holton, G. (1969a). Einstein and “crucial” experiments. American Journal of Physics, 37(10), 968–982. Holton, G. (1969b). Einstein, Michelson and the “crucial” experiment. Isis, 60, 133–197. Holton, G. (1974). On being caught between Dionysians and Appollonians. Daedalus, 103, 65–81. Holton, G. (1978a). Subelectrons, presuppositions, and the Millikan-Ehrenhaft dispute. Historical Studies in the Physical Sciences, 9, 161–224. Holton, G. (1978b). On the educational philosophy of the Project Physics Course. In G. Holton (Ed.), The scientific imagination (pp. 294–298). New York: Cambridge University Press. Holton, G. (1986). The advancement of science and its burdens. Cambridge, UK: Cambridge University Press. Holton, G. (1992). Ernst Mach and the fortunes of positivism in America. Isis, 83, 27–60. Holton, G. (1993). Science and anti-science. Cambridge, MA: Harvard University Press. Holton, G. (1999). R.A.  Millikan’s struggle with the meaning of Planck’s constant. Physics in Perspective, 1, 231–237. Holton, G. (2000). Personal communication, September 25. Holton, G. (2014a). The neglected mandate: Teaching science as part of our culture. Science & Education, 23, 1875–1877. Holton, G. (2014b). Personal communication, August 3, italics in the original. Holtzclaw, H.F., & Robinson, W.R. (1988). General chemistry (8th ed.). Lexington, MA: Heath House, E. R. (1991). Realism in research. Educational Researcher, 20(6), 2–9. 25. Hoyningen-Huene, P. (1993). Reconstructing scientific revolutions: Thomas S. Kuhn’s philosophy of science. Chicago: University of Chicago Press. Jackson, D.  F., Doster, E.  C., Meadows, L., & Wood, T. (1995). Hearts and minds in the science classroom: The education of a confirmed evolutionist. Journal of Research in Science Teaching, 32(6), 585–611.

206

References

Jammer, M. (1966). The conceptual development of quantum mechanics. New York: McGraw-Hill. Jenkins, E. (2007). School science: A questionable construct? Journal of Curriculum Studies, 39(3), 265–282. Kadvany, J. (2001). Imre Lakatos and the guises of reason. Durham, NC: Duke University Press. Kalman, C.  S. (2002). Developing critical thinking in undergraduate courses: A philosophical approach. Science & Education, 11(1), 83–94. Kalman, C. S. (2009a). A role for experiment in using the law of inertia to explain the nature of science: A comment on Lopes Coelho. Science & Education, 18(1), 25–31. Kalman, C. S. (2009b). The need to emphasize epistemology in teaching and research. Science & Education, 18(3–4), 325–347. Kalman, C. S. (2013). Science and religion, separate pursuits. Physics Today, 66(8), 10–12. Kalman, C.  S. (2017). Successful science and engineering teaching: Theoretical and learning perspectives (2nd ed.). Dordrecht, the Netherlands: Springer. Kalman, C.  S. (2019a). Personal communication to author, February 18, reproduced with permission. Kalman, C.  S. (2019b). Personal communication to author, February 26, reproduced with permission. Karam, R. (2014). Review of Achinstein’s Evidence and Method: Scientific strategies of Isaac Newton and James Clerk Maxwell. Science &Education, 23(10), 2137–2148. Kellert, S. H., Longino, H. E., & Waters, C. K. (Eds.). (2006). Scientific Pluralism. Minneapolis, MN: University of Minnesota Press. Kelly, G. A. (1955). The psychology of personal constructs. New York: W.W. Norton & Co.. Kidd, I. J. (2016). Was Feyerabend a postmodernist? International Studies in the Philosophy of Science, 30(1), 1–14. Kitchener, R. F. (1993). Piaget’s epistemic subject and science education: Epistemological versus psychological issues. Science & Education, 2, 137–148. Kitcher, P. (1993). The advancement of science: Science without legend, objectivity without illusions. Oxford, UK: Oxford University Press. Kitcher, P. (2011). Science in a democratic society. New York: Prometheus. Klassen, S. (2006). A theoretical framework for contextual science teaching. Interchange, 37, 31–62. Koertge, N. (1996). Toward an integration of content and method in the science curriculum. Science & Education, 5(4), 391–406 (First published in 1969). Koestler, A. (1964). The act of creation. London: Hutchinson. Kohler, R. E. (1971). The origin of Lewis’s theory of the shared pair bond. Historical Studies in the Physical Sciences, 3, 343–376. Kousathana, M., Demerouti, M., & Tsaparlis, G. (2005). Instructional misconceptions in acid-­ base equilibria: An analysis from a history and philosophy of science perspective. Science & Education, 14(2), 173–194. Kragh, H. (1999). Quantum generations: A history of physics in the twentieth century. Princeton, NJ: Princeton University Press. Krane, K. S. (1996). Modern physics (2nd ed.). New York: Wiley. Kuhn, T. S. (1962). The structure of scientific revolutions. Chicago: University of Chicago Press. Kuhn, T. S. (1963). The function of dogma in scientific research. In A. C. Crombie (Ed.), Scientific change (pp. 347–369). New York: Basic Books. Kuhn, T.  S. (1970). The structure of scientific revolutions (2nd ed.). Chicago: University of Chicago Press. Kuhn, T. S. (1977). The function of measurement in modern physical science. In T.S. Kuhn (Ed.), Essential tension (pp. 178–224). Chicago: University of Chicago Press. (Originally published in Isis, 52, 161–190, 1961). Laats, A., & Siegel, H. (2016). Teaching evolution in a creation nation. Chicago: University of Chicago Press.

References

207

Lakatos, I. (1970). Falsification and the methodology of scientific research programmes. In I.  Lakatos & A.  Musgrave (Eds.), Criticism and the growth of knowledge (pp.  91–195). Cambridge, UK: Cambridge University Press. Lakatos, I. (1971a). History of science and its rational reconstructions. In R. Buck & R. Cohen (Eds.), P.S.A. 1970, Boston Studies in the Philosophy of Science (Vol. 8, pp.  102–138). Dordrecht, the Netherlands: Reidel. Lakatos, I. (1971b). Replies to critics. In R. Buck & R. Cohen (Eds.), P.S.A. 1970, Boston Studies in the Philosophy of Science (Vol. 8, pp. 174–182). Dordrecht, the Netherlands: Reidel. Lakatos, I. (1978). Newton’s effect on scientific standards. In J.  Worrall & G.  Currie (Eds.). The methodology of scientific research programmes. Vol I (pp.  193–236). Cambridge, UK: Cambridge University Press. (Early drafts of this paper were written in 1963–64, and published posthumously). Lakatos, I., & Musgrave, A. (Eds.). (1970). Criticism and the growth of knowledge. Cambridge, UK: Cambridge University Press. Laloë, F. (2001). Do we really understand quantum mechanics? Strange correlations, paradoxes, and theorems. American Journal of Physics, 69, 655–701. Latour, B. (1987). Science in action. Boston, MA: Harvard University Press. Latour, B. (1992). Aramis ou l’amour des techniques [Aramic or the love of technology]. Paris: Éditions la Découverte. Laudan, L. (1977). Progress and its problems. Berkeley, CA: University of California Press. Laudan, R., Laudan, L., & Donovan, A. (1988). Testing theories of scientific change. In A. Donovan, L. Laudan, & R. Laudan (Eds.), Scrutinizing science: Empirical studies of scientific change (pp. 3–44). Dordrecht, the Netherlands: Kluwer. Lave, J., & Wenger, E. (1991). Situated learning: Legitimate peripheral participation. Cambridge, UK: Cambridge University Press. Lederman, N. G. (1992). Students’ and teachers’ conceptions of the nature of science: A review of the research. Journal of Research in Science Teaching, 29(4), 331–359. Lederman, N. G. (2004). Syntax of nature of science within inquiry and science instruction. In L. B. Flick & N. G. Lederman (Eds.), Scientific inquiry and nature of science (pp. 301–317). Dordrecht, the Netherlands: Kluwer. Lewis, G.  N. (1916). The atom and the molecule. Journal of American Chemical Society, 38, 762–785. Lewis, G. N. (1919). Letter to Robert Millikan (Reproduced in Kohler, 1971). Lewis, G.  N. (1923). Valence and the structure of atoms and molecules. New  York: Chemical Catalog. Longino, H. E. (1990). Science as social knowledge: Values and objectivity in scientific inquiry. Princeton, NJ: Princeton University Press. Lorentz, H. A. (1895). Versuch einer theorie de electrischen und optischen erscheinunungen in bewegten Körpern (collected papers, vol. 5). Leiden, the Netherlands: Brill. Losee, J.  (2001). A historical introduction to the philosophy of science. Oxford, UK: Oxford University Press. Loving, C. C. (1991). The scientific theory profile: A philosophy of science models for science teachers. Journal of Research in Science Teaching, 28(9), 823–838. Loving, C. C. (1997). From the summit of truth to its slippery slopes: Science education’s journey through positivist-postmodern territory. American Educational Research Journal, 34(3), 421–452. Machamer, P. K. (1973). Feyerabend and Galileo: The interaction of theories, and the reinterpretation of experience. Studies in History and Philosophy of Science, 4(1), 1–46. Machamer, P. K., Pera, M., & Baltas, A. (Eds.). (2000). Scientific controversies: Philosophical and historical perspectives. New York: Oxford University Press. Mackenzie, J., Good, R. G., & Brown, J. R. (2014). Postmodernism and science education: An appraisal. In M. R. Matthews (Ed.), International Handbook of Research in History, Philosophy and Science Teaching (Vol. II, pp. 1057–1086). Dordrecht, the Netherlands: Springer.

208

References

Margenau, H. (1950). The nature of physical reality. New York: McGraw-Hill. Matthews, M.  R. (1987). Experiment as the objectification of theory: Galileo’s revolution. In Proceedings of the second international seminar on misconceptions and educational strategies in science and mathematics (Vol. 1, pp. 289–298). Ithaca, NY: Cornell University. Matthews, M.  R. (2009). Science, worldviews and education: An introduction. Science & Education, 18(6–7), 641–666. Matthews, M. R. (Ed.). (2014a). International handbook of research in history, philosophy and science teaching (3 volumes). Dordrecht, the Netherlands: Springer. Matthews, M.  R. (2014b). Introduction: The history, purpose and content of the Springer. In International handbook of research in history, philosophy and science teaching (Vol. I, pp. 1–15). Dordrecht, the Netherlands: Springer. Matthews, M. R. (2015). Science teaching: The contribution of history and philosophy of science (20th Anniversary Rev. and Exp. ed.). New York: Routledge. Maxwell, J. C. (1860). Illustrations of the dynamical theory of gases. Philosophical Magazine, 19, 19–32 (Reproduced in Scientific Papers, 1965, pp. 377–409, New York: Dover). Maxwell, G. (1962). The ontological status of theoretical entities. In H. Feigel & G. Maxwell (Eds.), Minnesota Studies in the Philosophy of Science, Vol III (pp. 3–27). Minneapolis: University of Minnesota Press. Mayo, D. G. (1988). Brownian motion and the appraisal of theories. In A. Donovan, L. Laudan, & R. Laudan (Eds.), Scrutinizing science: Empirical studies of scientific change (pp. 219–243). Dordrecht, the Netherlands: Springer. Mayr, E. (1982). The growth of biological thought: Diversity, evolution and inheritance. Cambridge, MA: Harvard University Press. McCarthy, C. L. (2014). Cultural studies in science education: Philosophical considerations. In M. R. Matthews (Ed.), International handbook of research in history, philosophy and science teaching (Vol. III, pp. 1927–1964). Dordrecht, the Netherlands: Springer. McDonald, C.  V., & Abd-El-Khalick, F. (2017). Where to from here? Implications and future directions for research on representations of nature of science in school science textbooks. In C. V. McDonald & F. Abd-El-Khalick (Eds.), Representations of nature of science in school science textbooks: A global perspective (pp. 215–231). New York: Routledge. McMullin, E. (1985). Galilean idealization. Studies in History and Philosophy of Science, 16, 247–273. Medawar, P. B. (1967). The art of the soluble. London: Methuen. Medicus, H. A. (1974). Fifty years of matter waves. Physics Today, 27, 38–45. Mendeleev, D. (1869). Ueber die beziehungen der eigenschaften zu den atom gewichten der elemente. Zeitschrift für Chemie, 12, 405–406. Mendeleev, D. (1879). The periodic law of the chemical elements. The Chemical News, 40, No. 1042. Mendeleev, D. (1889). The periodic law of the chemical elements. Journal of the Chemical Society, 55, 634–656 (Faraday lecture, delivered on 4 June 1889). Mendeleev, D. (1897). The principles of chemistry (2nd English ed., trans. of 6th Russian ed.). New York: American Home Library Company. Michelson, A. A. (1927). Studies in optics. Chicago: University of Chicago Press. Michelson, A. A., & Morley, E. W. (1887). On the relative motion of the earth and the luminiferous ether. American Journal of Science, 34(3rd series), 333–345. Miller, D. C. (1925). Science, 61, 617. Miller, D. C. (1933). The ether-drift experiment and the determination of the absolute motion of the earth. Review of Modern Physics, 5, 204–242. Millikan, R. A. (1913). On the elementary electrical charge and the Avogadro constant. Physical Review, 2, 109–143. Millikan, R. A. (1916). A direct photoelectric determination of Planck’s “h”. Physical Review, 7, 355–388.

References

209

Moseley, H. G. J. (1913). High frequency spectra of the elements. Philosophical Magazine, 26, 1025–1034. Moseley, H. G. J. (1914). High frequency spectra of the elements. Part II. Philosophical Magazine, 27, 703–713. Motterlini, M. (1999). Ed. For and against method: Including Lakatos’s lectures on scientific method and the Lakatos-Feyerabend correspondence. Chicago: University of Chicago Press. Motterlini, M. (2002a). Professor Lakatos between the Hegelian devil and the Popperian deep blue sea. In G. Kampis, L. Kvasz, & M. Stöltzner (Eds.), Appraisng Lakatos: Mathematics, methodology and the man (pp. 23–52). Dordrecht, the Netherlands: Kluwer Academic Publishers. Motterlini, M. (2002b). Reconstructing Lakatos: A reassessment of Lakatos’s epistemological project in the light of the Lakatos archive. Studies in History and Philosophy of Science, 33A(3), 487–509. Motterlini, M. (2016). Paul Karl Feyerabend. In J. Pfeifer & S. Sarkar (Eds.), The philosophy of science: An encyclopedia (pp. 1–22). London: Routledge. Mugaloglu, E. Z. (2014). The problem of pseudoscience in science education and implications of constructivist pedagogy. Science & Education, 23(4), 829–842. Munévar, G. (2016). Historical antecedents to the philosophy of Paul Feyerabend. Studies in History and Philosophy of Science, 57, 9–16. Musgrave, A. (1976a). Method or madness? Can the methodology of research programmes be rescued from epistemological anarchism? In R. S. Cohen, P. K. Feyerabend & M. W. Wartofsky (Eds.), Essays in memory of Imre Lakatos (Boston Studies in the Philosophy of Science, Vol 39, pp. 457–491). Dordrecht, the Netherlands: Reidel. Musgrave, A. (1976b). Why did oxygen supplant phlogiston? Research programmes in the chemical revolution. In C. Howson (Ed.), Method and appraisal in the physical sciences: The critical background to modern science, 1800–1905 (pp.  181–209). Cambridge, UK: Cambridge University Press. Nanda, M. (2003). Prophets facing backward: Postmodern critiques of science and Hindu nationalism in India. New Brunswick, NJ: Rutgers University Press. National Research Council (NRC). (1996). National science education standards. Washington, DC: National Academy Press. NGSS Lead States. (2013). Next generation science standards: For states, by states. Washington, DC: The National Academies Press. Niaz, M. (1993). If Piaget’s epistemic subject is dead, shall we bury the scientific methodology of idealization. Journal of Research in Science Teaching, 30, 809–812. Niaz, M. (1998). From cathode rays to alpha particles to quantum of action: A rational reconstruction of structure of the atom and its implications for chemistry textbooks. Science Education, 82, 527–552. Niaz, M. (2000a). A rational reconstruction of the kinetic molecular theory of gases based on history and philosophy of science and its implications for chemistry textbooks. Instructional Science, 28, 23–50. Niaz, M. (2000b). The oil drop experiment: A rational reconstruction of the Millikan-Ehrenhaft controversy and its implications for chemistry textbooks. Journal of Research in Science Teaching, 37(5), 480–508. Niaz, M. (2001). A rational reconstruction of the origin of the covalent bond and its implications for general chemistry textbooks. International Journal of Science Education, 23(6), 623–641. Niaz, M. (2004). Exploring alternative approaches to methodology in educational research. Interchange, 35(2), 155–184. Niaz, M. (2005). An appraisal of the controversial nature of the oil-drop experiment: Is closure possible? British Journal of the Philosophy of Science, 56, 681–702. Niaz, M. (2009). Critical appraisal of physical science as a human enterprise: Dynamics of scientific progress. Dordrecht, the Netherlands: Springer.

210

References

Niaz, M. (2010). Science curriculum and teacher education: The role of presuppositions, contradictions, controversies and speculations vs Kuhn’s ‘normal science’. Teaching and Teacher Education, 26, 891–899. Niaz, M. (2011a). How to facilitate teachers’ understanding of hypotheses and predictions? Interchange, 42(1), 51–58. Niaz, M. (2011b). Innovating science teacher education: A history and philosophy of science perspective. New York: Routledge. Niaz, M. (2012). From ‘science in the making’ to understanding the nature of science: An overview for science educators. New York: Routledge. Niaz, M. (2015). That the Millikan oil-drop experiment was simple and straightforward. In R. L. Numbers & K. Kampourakis (Eds.), Newton’s apple and other myths about science (pp. 157– 163). Cambridge, MA: Harvard University Press. Niaz, M. (2016). Chemistry education and contributions from history and philosophy of science. Dordrecht, the Netherlands: Springer. Niaz, M. (2017). Relationship between domain-specific and domain-general aspects of nature of science in science textbooks. In C. V. McDonald & F. Abd-El-Khalick (Eds.), Representations of nature of science in school science textbooks (pp. 61–78). New York: Routledge. Niaz, M. (2018). Evolving nature of objectivity in the history of science and its implications for science education. Dordrecht, the Netherlands: Springer. Niaz, M., Abd-El-Khalick, F., Benarroch, A., Cardellini, L., Laburu, C. E., Marín, N., et al. (2003). Constructivism: Defense or a continual critical appraisal  — A response to Gil-Pérez, et  al. Science & Education, 12(8), 787–797. Niaz, M., Klassen, S., McMillan, B., & Metz, D. (2010). Reconstruction of the history of the photoelectric effect and its implications for general physics textbooks. Science Education, 94, 903–931. Niaz, M., & Luiggi, M. (2014). Facilitating conceptual change in students’ understanding of the periodic table. Dordrecht, the Netherlands: Springer. Niaz, M., & Marcano, C. (2012). Reconstruction of wave-particle duality and its implications for general chemistry textbooks. Dordrecht, the Netherlands: Springer. Niaz, M., & Maza, A. (2011). Nature of science in general chemistry textbooks. Dordrecht, the Netherlands: Springer. Niaz, M., & Rivas, M. (2016). Students’ understanding of research methodology in the context of dynamics of scientific progress. Dordrecht, the Netherlands: Springer. Niaz, M., & Rodríguez, M. A. (2001). Do we have to introduce history and philosophy of science or is it already “inside” chemistry? Chemistry Education: Research and Practice in Europe, 2, 159–164. Niaz, M., Rodríguez, M. A., & Brito, A. (2004). An appraisal of Mendeleev’s contribution to the development of the periodic table. Studies in History and Philosophy of Science, 35, 271–282. Niño, L. M. (2009). Reconstrucción racional del origen de las teorías acido-base y sus implicaciones en los textos de química universitarios: Un enfoque Lakatosiano. Master of Science Thesis (chemistry education). Cumaná, Venezuela: Universidad de Oriente, Venezuela. Nola, R. (1996). Review of Feyerabend’s Killing Time. British Journal for the Philosophy of Science, 47(3), 467–473. Nola, R. (1997). Constructivism in science and science education: A philosophical critique. Science & Education, 6, 55–83. Novak, J. D. (1990). Concept mapping: A useful tool for science education. Journal of Research in Science Teaching, 27(10), 937–949. Oberheim, E. (2016). Rediscovering Einstein’s legacy: How Einstein anticipates Kuhn and Feyerabend on the nature of science. Studies in History and Philosophy of Science, 57, 17–26. O’Loughlin, M. (1992). Rethinking science education: Beyond Piagetian constructivism toward a sociocultural model of teaching and learning. Journal of Research in Science Teaching, 29(8), 791–820.

References

211

O’Neill, D. K., & Polman, J. L. (2004). Why educate “little scientists?” Examining the potential of practice-based scientific literacy. Journal of Research in Science Teaching, 41(3), 234–266. Ospina, J. (2010). Efecto fotoeléctrico: Una reconstrucción racional basada en la historia y filosofía de la ciencia y sus implicaciones para los textos de química general. Master of Science Thesis (chemistry education), Cumaná, Venezuela: Universidad de Oriente. Pais, A. (1982). “Subtle is the Lord …” The science and the life of Albert Einstein. Oxford, UK: Oxford University Press. Park, H., Nielsen, W., & Woodruff, E. (2014). Students’ conceptions of the nature of science: Perspectives from Canadian and Korean middle school students. Science & Education, 23(5), 1169–1196. Patton, M. (1990). Qualitative evaluation and research methods (2nd ed.). Newbury Park, CA: Sage. Perl, M. L. (1997). Reflections on the discovery of the Tau Lepton. In G. Ekspong (Ed.), Nobel lectures, physics 1991–1995 (pp. 168–195). Singapore: World Scientific Publishing Co. (Nobel prize acceptance speech delivered on December 8, 1995). Perl, M. L. (2004). The discovery of the Tau Lepton and the changes in elementary-particle physics in forty years. Physics in Perspective, 6, 401–427. Perl, M. L. (2007). A contrarian view of how to develop creativity in science and engineering. Paper presented at The Eighth Olympiad of the Mind, The National Academies, Washington, DC, November (SLAC-PUB-12850). Perl, M.  L., Abrams, G.  S., Boyarski, A.  M., Breidenbach, M., Briggs, D.  D., Bulos, F., et  al. (1975). Evidence for anomalous lepton production in e+-e- annihilation. Physics Review Letters, 35, 1489–1492. Perrin, J. (1923). Atoms (D. L. Hammick, Trans.). London: Constable. Piaget, J. (1971). Biology and knowledge. Edinburgh, Scotland: Edinburgh University Press. Piaget, J., & Garcia, R. (1989). Psychogenesis and the history of science. New York: Columbia University Press. PLUTO-Collaboration, et al. (1977). Anomalous muon production e+e- annihilations as evidence for heavy leptons. Physics Letters B, 68, 297–300. Polanyi, M. (1964). Personal knowledge: Towards a post-critical philosophy. New York: Harper & Row (First published 1958 by the University of Chicago Press). Polanyi, M. (1966). The tacit dimension. London: Routledge & Kegan Paul. Polanyi, M. (1972). Genius in science. Encounter, 38(1), 43–50. Popper, K. R. (1957). The aim of science. Ratio, 1, 24–35. Popper, K. R. (1959). The logic of scientific discovery. New York: Harper & Row. Popper, K. R. (1963a). Conjectures and refutations. London: Routledge and Kegan Paul. Popper K. R. (1963b). The open society and its enemies (4th ed., first published 1945). New York: Harper Torchbooks. Popper, K. (1970). Normal science and its dangers. In I. Lakatos & A. Musgrave (Eds.), Criticism and the growth of knowledge (pp. 51–58). Cambridge: Cambridge University Press Preston, J. (1997). Feyerabend: Philososphy, science and society. Cambridge, UK: Polity Press. Preston, J., Munévar, G., & Lamb, D. (Eds.). (2000). The worst enemy of science? Essays in memory of Paul Feyerabend. New York: Oxford University Press. Priestley, J. (1790). Experiments and observations on different kinds of air, and other branches of natural philosophy (Vol. I). Birmingham, UK: Thomas Pearson. Olenick, R. P., Apostol, T. M., & Goodstein, D. L. (1985). Beyond the mechanical universe. From electricity to modern physics. New York: Cambridge University Press. Quale, A. (2007). Radical constructivism, and the sin of relativism. Science & Education, 16(3–5), 231–266. Rampal, A. (1992). Maintaining the status quo — A response to Fred Wilson and John Wilson. Interchange, 23(3), 309–314. Ramsay, W. (1897). An undiscovered gas (address to the Section of Chemical Sciences of the British Association). Nature, 56, 378–382.

212

References

Ramsay, W., & Travers, M. W. (1901). Philosophical Transactions of the Royal Society A, 197, 47. Reines, F. (1997). The neutrino: From poltergeist to particle. In G. Ekspong (Ed.), Nobel lectures, physics, 1991–1995 (pp. 202–221). Singapore, Singapore: World Scientific Publishing Co.. Rey, J. (1630). Essays on an enquiry into the cause wherefore tin and lead increase in weight on calcination. Alembic Club Reprints, No. 11, 1953. Robottom, I. (1989). Social critique or social control: Some problems for evaluation in environmental education. Journal of Research in Science Teaching, 26(5), 435–443. Rodebush, W. H. (1928). The electron theory of valence. Chemical Review, 5, 509–531. Rodríguez, M. A., & Niaz, M. (2004). A reconstruction of structure of the atom and its implications for general physics textbooks. Journal of Science Education and Technology, 13, 409–424. Rorty, R. (1979). Philosophy and the mirror of nature. Princeton, NJ: Princeton University Press. Rorty, R. (1989). Contingency, irony, and solidarity. Cambridge: Cambridge University Press. Roth, W.-M., & Roychoudhury, A. (1994). Physics students’ epistemologies and views about knowing and learning. Journal of Research in Science Teaching, 31(1), 5–30. Rowbottom, D. P. (2013). Review of Feyerabend’s The tyranny of science. Science &Education, 22(5), 1229–1231. Rucker, W. A., & Kelvin, L. (1895). Contribution to untitled comments. Chemical News, 71(1836), 62. Rudolph, J. L. (2000). Reconsidering the ‘nature of science’ as a curriculum component. Journal of Curriculum Studies, 32(3), 403–419. Ruse, M. (Ed.). (2013). The Cambridge encyclopedia of Darwin and evolution. Cambridge, UK: Cambridge University Press. Rutherford, E. (1911). The scattering of alpha and beta particles by matter and the structure of the atom. Philosophical Magazine, 21, 669–688. Rutherford, E. (1915). The constitution of matter and the evolution of the elements. In Address to the annual meeting of the National Academy of Sciences (pp. 167–202). Washington, DC: Smithsonian Institution. Sandin, T. R. (1989). Modern physics. Reading, MA: Addison-Wesley. Sankey, H. (2012). Philosophical fairytales from Feyerabend. Metascience, 21(2), 471–476. Scheffler, I. (1967). Science and subjectivity. Indianapolis, IN: Bobbs-Merrill. Schwab, J. J. (1974). The concept of the structure of a discipline. In E. W. Eisner & E. Valance (Eds.), Conflicting conceptions of curriculum (pp.  162–175). Berkeley, CA: McCutchan Publishing Corp. Segre, M. (1997). Light on the Galileo case? Isis, 88, 484–504. Serway, R. A. (1996). Physics for scientists and engineers with modern physics (4th ed., Spanish). New York: McGraw-Hill. Shankland, R. S. (1963). Conversations with Albert Einstein. American Journal of Physics, 31, 47–57. Shankland, R. S. (1964). The Michelson-Morley experiment. American Journal of Physics, 32, 23. Shapere, D. (1977). Scientific theories and their domains. In F. Suppe (Ed.), The structure of scientific theories (2nd ed., pp. 518–565). Chicago: University of Illinois Press. Shapin, S. (1996). The scientific revolution. Chicago: University of Chicago Press. Shaw, J.  (2019). The revolt against rationalism: Feyerabend’s critical philosophy. Studies in History and Philosophy of Science (in press). Siegel, H. (1979). On the distortion of the history of science in science education. Science Education, 63, 111–118. Siegel, H. (1991). The rationality of science, critical thinking, and science education. In M.  R. Matthews (Ed.), History, philosophy and science teaching: Selected readings (pp.  45–62). Toronto, Canada: OISE Press. Smart, W. M. (1946). John Couch Adams and the discovery of Neptune. Nature, 158, 648–652. Smith, E. F. (1925). Observations on teaching the history of chemistry. Journal of Chemical Education, 2, 553–555.

References

213

Smith, M. U. (1994). Belief, understanding, and the teaching of evolution. Journal of Research in Science Teaching, 31, 591–597. Smith, M. U., Lederman, N. G., Bell, R. L., McComas, W. F., & Clough, M. P. (1997). How great is the disagreement about the nature of science? A response to Alters. Journal of Research in Science Teaching, 34(10), 1101–1103. Smith, M. U., & Siegel, H. (2004). Knowing, believing, and understanding: What goals for science education? Science & Education, 13(6), 553–582. Sorgner, H. (2016). Challenging expertise: Paul Feyerabend vs. Harry Collins and Robert Evans on democracy, public participation and scientific authority. Studies in History and Philosophy of Science, 57, 114–120. Swartz, R. (1985). Dewey and Popper on learning from induction. Interchange, 16(4), 29–51. Taber, K. S. (2014). Methodological issues in science education research: A perspective from the philosophy of science. In M. R. Matthews (Ed.), International handbook of research in history, philosophy and science teaching (Vol. III, pp. 1839–1893). Dordrecht, the Netherlands: Springer. Theocharis, T., & Psimopoulos, M. (1987). Where science has gone wrong. Nature, 329, 595–598. Thomason, N. (1994). The power of ARCHED hypotheses: Feyerabend’s Galileo as a closet rationalist. British Journal for the Philosophy of Science, 45(1), 255–264. Thomson, J. J. (1897). Cathode rays. Philosophical Magazine, 44, 293–316. Thomson, J. J. (1907). The corpuscular theory of matter. London: Constable. Thomson, J. J. (1914). The forces between atoms and chemical affinity. Philosophical Magazine, 27, 757–789. Tiberghein, A., Cross, D., & Sensevy, G. (2014). The evolution of classroom physics knowledge in relation to certainty and uncertainty. Journal of Research in Science Teaching, 51(7), 930–961. Tolvanen, S., Jansson, J., Vesterinen, V.-M., & Aksela, M. (2014). How to use historical approach to teach nature of science in chemistry education? Science & Education, 23(8), 1605–1636. Van Spronsen, J. (1969). The periodic system of chemical elements. A history of the first hundred years. Amsterdam, the Netherlands: Elsevier. Van Strien, M. (2019). Pluralism and anarchism in quantum physics: Paul Feyerabend’s writings on quantum physics in relation to his general philosophy of science. Studies in History and Philosophy of Science (in press). Wartofsky, M.  W. (1968). Conceptual foundations of scientific thought: An introduction to the philosophy of science. New York: Macmillan. Weinberg, S. (2001). Physics and history. In J. A. Labinger & H. M. Collins (Eds.), The one culture: A conversation about science. Chicago: University of Chicago Press. Weisberg, M. (2007). Who is a modeler? British Joural for the Philosophy of Science, 58, 207–233. Wheaton, B.  R. (1983). The tiger and the shark: Empirical roots of wave-particle dualism. Cambridge, UK: Cambridge University Press. Wilson, D. (1983). Rutherford: Simple genius. Cambridge, MA: MIT Press. Winchester, I. (1989). Editorial: History, science and science teaching. Interchange, 20(2), i–vi. Winchester, I. (1993). “Science is dead. We have killed it, you and I” — How attacking the presuppositional structures of our scientific age can doom the interrogation of nature. Interchange, 24(1–2), 191–198. Windschitl, M. (2004). Folk theories of “inquiry:” How preservice teachers reproduce the discourse and practices of an atheoretical scientific method. Journal of Research in Science Teaching, 41, 481–512. Wolpert, L. (1993). The unnatural nature of science. Cambridge, MA: Cambridge University Press. Worrall, J. (2010). Theory-change in science. In S. Psillos & M. Curd (Eds.), The Routledge companion to philosophy of science (pp. 281–291). New York: Routledge. Wynne, B. (1989). Sheep farming after chernobyl: A case study in communicating scientific information. Environment: Science and Policy for Sustainable Development, 31(2), 10–39. Ziman, J.  (1978). Reliable knowledge: An exploration of the grounds for belief in science. Cambridge, UK: Cambridge University Press.

Name Index

A Abd-El-Khalick, F., 60, 76, 77, 88, 91, 160 Achinstein, P., 1, 57–59, 112 Agassi, J., 64, 87, 94, 95, 155, 168 Aksela, M., 167 Alters, B.J., 74 Ampère, A.-M., 59 Anaximander, 155 Arabatzis, T., 115 Arrhenius, S., 42, 44, 162 Atkinson, P., 40 Aulls, M.W., 60, 61 B Bachelard, G., 56 Bailin, S., 165 Baltas, A., 99 Bauer, H.H., 77, 100 Bazghandi, P., 6 Beghetto, R.A., 73 Belarmino, J., 160 Bell, R.L., 74, 76 Bellama, J., 111, 118 Bellarmine, R., 30, 31 Ben-Ari, M., 69, 155 Bensaude-Vincent, B., 137 De Berg, K.C., 42, 66–68, 163, 164 Berthelot, M., 140 Bessel, F., 31 Bird, A., 4 Bjelic, D., 72 Bloor, D., 35, 88 Bodner, G.M., 146, 147

Bohr, N., 17, 18, 20, 24–26, 44, 46, 56, 84, 112, 119, 120, 125–127, 132, 133, 135, 146, 158, 163, 170 Boutonné, S., 77 Bransford,J.D., 73 Brito, A., 136, 139, 141, 144 De Broglie, L., 18, 131–133, 159 Brønsted, J.N., 43 Brown, J.R., 86, 102, 171 Brown, M.J., 32, 34 Brown, R., 110, 111 Bruno, G., 31 Brush, S.G., 13, 88, 104–106, 112, 115, 137, 140, 161 Bunge, M., 21, 56, 170 C Campbell, D.T., 12, 37, 83, 84 Cartwright, N., 5, 9, 21, 32, 57, 58, 92, 169, 170 Cavazos, L., 36 Chalmers, A.F., 32, 66, 67 Chang, H., 48 Charmaz, K., 40, 41, 71, 81, 97, 98 Clark, P., 113 Clarke, S.W., 64 Clausius, R., 111, 157 Clough, M.P., 74 Cobern, W.W., 46, 47, 51, 52, 54 Collins, H.M., 28–30, 35, 60, 88, 99, 164, 168 Collins, R.W., 8, 14 Conant, J.B., 25 Cooper, L.N., 52, 103, 124, 127 Cordero, A., 62, 170

© Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3

215

Name Index

216 Cowan, C.L., 150 Crease, R.M., 116 Cross, D., 55 Crowther, J.G., 123, 124 Currie, G., 58 Cushing, J.T., 35, 162 D Darensbourg, M., 127 Darrigol, O., 126 Darwin, C., 13 Daston, L., 13, 21, 33, 36, 53, 62, 74, 90, 103, 104, 165, 166, 171 Davisson, C., 18, 132, 133, 159 Delamont, S., 40 Demerouti, M., 43, 162 Denzin, N.K., 41, 71, 81, 97 Dickerson, R., 127 Dobzhansky, T., 51 Donovan, A., 110 Donovan, S.M., 73 Doster, E.C., 53 Drago, A., 55, 56, 167 Duhem, P., 4, 5, 15, 21, 56, 58, 59, 67, 89, 93, 94, 110, 136, 157, 169 Dupre, J., 9 Duschl, R.A., 78 E Earman, J., 33 Eflin, J.T., 75, 167 Ehrenhaft, F., 16, 33, 116, 158, 166 Einstein, A., 1, 4, 13, 16–18, 20, 53, 56, 85–87, 89, 91–95, 110, 111, 114–116, 128–133, 148, 149 Erickson, F.E., 12, 13, 83, 84 Ernest, P., 47 Evans, R., 8, 28–30, 168 F Farrell, R.P., 25, 28, 34 Fay, R., 140, 162 Feyerabend, P.K., 1, 22, 23, 37, 39, 69, 71–79, 81–107, 109–173 Finocchiaro, M.A., 30, 31, 63 Foscarini, F., 30 Fosnot, C.T., 46 G Galileo, G., 8, 10, 11, 13, 15, 24, 30, 32, 34, 42, 49, 60, 63, 64, 73

Galison, P., 165, 166, 171 Galison, P.L., 13, 21, 33, 36, 53, 62, 74, 90, 103, 104 Garcia, R., 73 Garrett, A., 139 Gattei, S., 33, 34 Gavroglu, K., 115 Geelan, D., 44, 46, 47 Geelan, D.R., 5, 6, 23, 30, 31, 166 Geiger, H., 123 Gergen, K., 46 Germer, L.H., 18, 132, 133, 159 Giere, R.N., 1, 5, 9, 14–16, 32, 35, 36, 46, 47, 57, 58, 65, 74, 91–93, 106, 107, 113, 161, 169, 171, 172 Gillispie, C.C., 67 Gil-Pérez, D., 65 Ginev, D.J., 55 Glaser, B.G., 40 von Glasersfeld, 46, 47 Glass, R.J., 61 Glennan, S., 75, 167 Glymour, C., 33 Godfrey-Smith, P., 73 Goldberg, D., 140 Good, R.G., 5, 54, 77–79, 101, 104, 171 Gordin, M.D., 138 Gould, S.J., 51, 167 Gould, S.L., 74 Gower, B., 1 Gray, H., 127 Greca, I.M., 115 Grosser, M., 147 Guerra-Ramos, M.T., 45 H Hacking, I., 99 Halliday, D., 130 Hamrah, S.Z., 6 Hanson, N.R., 73, 78, 148, 150 Haraway, D., 36 Harding, S., 36 Hattiangadi, J.N., 1, 87, 168 Hecht, E., 130 Heering, P., 99, 169 Heffron, J.M., 66 Hegel, G.W.F., 27 Heilbron, J.L., 103, 124, 164 Henry, C., 35 Herschel, J., 25 Hertz, H., 128 Hiberty, P.C., 162 Hildebrand, G.M., 36 Hill, J., 139

Name Index Hodson, D., 6, 61, 78, 100, 156, 163 Hoffmann, R., 9, 37, 38, 48, 101, 162, 167, 169 Holton, G., 16, 17, 33, 67, 68, 88, 102, 103, 115, 159, 164, 166 Holtzclaw, H., 111 Höttecke, D., 99, 169 Hoyningen-Huene, P., 169 I Irzik, G., 65, 168 J Jackson, D.F., 53 Jacobs, S., 62 Jammer, M., 128 Jansson, J., 167 Jenkins, E., 100 Joesten, M., 127 Johnson, K.W., 127 K Kadvany, J., 2 Kalman, C.S., 24, 25, 45, 49, 51, 53, 54, 56, 60, 61, 63, 109, 114, 118, 119, 127, 156, 163, 167, 168 Kalman, J.R., 156, 171 Karam, R., 57–59, 169 Keller, E.F., 36 Kelly, G.A., 13, 46, 85 Kelvin, L., 138 Kidd, I.J., 9, 32, 34, 37, 156, 171 Kitchener, R.F., 112 Kitcher, P., 34, 112 Klassen, S., 129, 142 Koertge, N., 68, 76, 164 Koestler, A., 13, 87 Kohler, R.E., 145–147 Kotz, J.C., 131, 133, 134 Kousathana, M., 43, 162 Kragh, H., 149, 154 Kuhn, T.S., 2, 5, 6, 10, 14, 20, 43, 45, 47, 56, 58, 64, 66, 68, 74–76, 78, 83–85, 87, 89, 90, 92–94, 98–100, 102, 106, 163, 167, 170 L Laats, A., 52 Lakatos, I., 2, 6, 8, 10, 12, 13, 16, 17, 19, 20, 23, 28, 31, 43, 45, 47, 48, 57–59, 64, 66, 68, 76, 83–85, 89, 93, 95, 99, 102,

217 110, 112–115, 125, 126, 145, 146, 148, 149, 157–159, 167–170 Laloë, F., 162 Lamb, D., 107 Lamont, J., 55 Latour, B., 11, 72 Laudan, L., 57, 74, 92, 93, 110 Laudan, R., 92, 93, 110 Lave, J., 68 Lavoisier, A., 66–68 Le, A.-P., 77 Lederman, N.G., 57, 65, 74, 76 Lewis, G.N., 9, 19, 20, 43, 44, 109, 144–147, 160, 162 Lincoln, Y.S., 41, 71, 81, 97 Lippincott, W.T., 139 Longino, H.E., 36 Lorentz, H.A., 114–116, 125, 126, 159 Losee, J., 56 Loving, C.C., 5, 76, 167 Lowry, T.M., 43 Lucas, K.B., 77 Luiggi, M., 141–143 Lynch, M., 72 M Machamer, P.K., 32, 99 Mackenzie, J., 102, 104, 171 Marcano, C., 131 Margenau, H., 125 Marsden, E., 123 Martin, M., 75 Matthews, M.R., 6, 39, 50, 78, 91, 97, 100, 112, 163 Maxwell, J.C., 16, 75, 111–113, 125, 157 Mayo, D.G., 86, 110 Mayr, E., 51 Maza, A., 141 McCarthy, C.L., 106, 107, 172 McComas, W.F., 74 McDonald, V., 88 McGinn, M.K., 72, 73 McMillan, B., 129 McMullin, E., 112 McMurry, J., 140, 162 McRobbie, C.J., 77 Meadows, L., 53 Medawar, P.B., 95 Medicus, H.A., 131, 132 Mendeleev, D., 18–20, 109, 134–144 Metz, D., 129 Michelson, A.A., 20, 109, 114–116, 129, 157, 158 Michelson-Morley, E.W., 16, 20

Name Index

218 Mill, J.S., 25 Miller, D.C., 16, 26, 114, 115, 158 Millikan, R.A., 15–18, 33, 52, 67, 116–120, 128–132, 146, 158, 159, 166, 169 De Milt, C., 134 Moore, J.W., 139 Morley, E.W., 11, 109, 114, 115, 157 Moseley, H.G.J., 135, 137, 142 Motterlini, M., 2, 4, 23, 25, 27, 28, 30, 31, 76, 110, 148 Mugaloglu, E.Z., 50, 51, 54, 160, 171 Munevar, G., 35, 107 Musgrave, A., 28, 48, 66 Myers, J.Y., 60, 160 N Nanda, M., 170 Ndongko, T.M., 82, 83 Netterville, J., 127 Neuhauss, R., 165 Newton, I., 4, 5, 10, 13, 15, 21 Niaz, M., 1, 13, 33, 44, 46, 48, 52, 53, 58, 60, 65, 67, 74, 83, 84, 90, 92, 93, 99, 101, 103, 104, 106, 107, 112, 113, 119–122, 129, 130, 132, 134, 136, 139, 141–144, 146, 154, 160, 161, 164, 166, 169, 171 Nichols, S.E., 74 Nielsen, W., 49, 166 Niño, L.M., 43, 44 Nola, R., 3, 63–65, 122, 168 Novak, J.D., 120, 142 O Oberheim, E., 32, 35 O’Loughlin, M., 46 O’Neill, D.K., 77, 168 Ospina, J., 130 Ostwald, W., 16, 86, 88, 110, 157 P Pais, A., 114 Pardue, H., 146, 147 Park, H., 49, 166 Patton, M., 53 Pauli, W., 19 Pearson, K., 14 Pera, M., 99 Perl, M.L., 19, 150–154 Perrin, J., 86, 110, 111 Petrucci, R., 139

Piaget, J., 46, 47, 73, 165 Pinch, T., 35 Polanyi, M., 3, 11, 13, 14, 21, 42, 61, 62, 87, 99, 103, 104, 148, 168, 169 Polman, J.L., 77, 168 Pope, M.L., 85 Popper, K.R., 2, 3, 5, 6, 8, 10, 12, 14, 17, 23, 27, 43, 45, 47, 48, 57–59, 62, 64, 66, 73, 74, 76, 77, 83, 85, 89, 90, 92, 94–96, 99, 102, 114, 126, 163, 167–169 Preston, J., 1, 3, 99 Priestley, J., 66–68 Psimopoulos, M., 1 Purcell, K., 133, 134 Q Quale, A., 47, 48, 156, 167 R Rampal, A., 92, 167 Ramsey, W., 19, 137, 138 Reines, F., 150 Reisch, G., 75, 167 Resnick, R., 130 Rey, J., 66 Rivas, M., 119–122 Robinson, W., 111 Robottom, I., 73, 165 Rodebush, W.H., 145 Rodríguez, M.A., 119, 134, 139, 141 Rorty, R., 33, 34 Roth, W.-M., 72, 73, 77, 78, 102 Rowbottom, D.P., 59, 60, 160 Roychoudhury, A., 74, 78 Rucker, W.A., 138 Rudolph, J.L., 100 Ruse, M., 51 Rutherford, E., 17, 18, 25, 33, 46, 52, 117, 119, 120, 123–127, 132, 135, 137, 141, 146, 158, 159, 171 S Scheffler, I., 104 Schrödinger, E., 131 Schulz, R.M., 14, 100, 101 Schwab, J.J., 99 Schwartz, R.S., 76 Sears, F.W., 115 Segal, B., 133, 134, 147

Name Index Segre, M., 31 Sensevy, G., 55 Serway, R.A., 115 Shaik, S., 162 Shankland, R.S., 114, 115 Shapere, D., 137 Shapin, S., 77, 165 Siegel, H., 52, 54, 60, 99, 104, 106, 160, 164 Sisler, H., 140 Smith, M.U., 52–54, 74, 88 Solomon, J., 46 Sorgner, H., 29, 168 Stahl, G., 66 Strauss, A.L., 40 Swartz, R., 95, 96, 169 T Taber, K.S., 99, 100, 164 Taylor, P., 46 Theocharis, T., 1 Thomason, N., 32 Thomson, J.J., 17, 18, 33, 46, 52, 84, 119, 120, 123–125, 128, 132, 135, 137, 144–146, 158, 171 Tiberghein, A., 55 Tippins, D.J., 74 Tolvanen, S., 167 Travers, M.W., 138 Treichel, P.M., 131 Tro, N., 44, 162 Tsaparlis, G., 43, 162

219 U Umland, J.B., 111, 118 V van Spronsen, J., 135, 137 Van Strien, M., 162 Verhoek, F., 139 Vesterinen, V.-M., 167 W Wartofsky, M.W., 137 Waters, C.K., 46 Waters, M., 77 Weisberg, M., 137 Wenger, E., 68 Wheaton, B.R., 128 Wilson, D., 33, 103, 114, 124, 164, 171 Winchester, I., 13, 14, 85, 89–92, 161, 163 Windschitl, M., 77 Wood, T., 53 Woodhouse, H., 82, 83 Woodruff, E., 49, 166 Worrall, J., 58, 93, 94, 161 Wright, E., 78 Wynne, B., 29 Z Ziman, J., 137 Zumdahl, S.S., 113

Subject Index

A Absolute truths, 85, 156 Acid-base equilibria, 9, 20, 42–44, 162 Actor network theory (ATN), 11, 72 African and modern medicine, 82–83 Against method (AM), 76, 110 Alpha particle scattering experiments, 17, 20, 33, 109, 123–125, 158, 164, 171 Alternative approaches to growth of knowledge, 13, 82–85 Alternative hypotheses, 36, 139, 149, 160 Alternative literary forms, 11, 72–73 Anarchistic methodology, 10, 42, 45 Anomalies, 23, 24, 75, 170 Anything goes, 1, 3, 6–11, 13, 19, 25, 27, 28, 32, 35, 40, 41, 45, 47–50, 61, 65, 69, 85, 155, 156, 163, 167, 172 Arrhenius model, 9, 42, 43, 162 Atomic model, 25, 46, 83, 103, 119, 124, 125, 140, 158 Atwood’s machine, 5, 58 Auxiliary hypotheses, 23, 45, 114, 148, 149 Azande beliefs, 2 B Bending of light in the 1919 eclipse experiments, 33 Betrayal of reason, 4 Bohr model of orbits, 44 Bohr’s atomic theory, 24 Brønsted-Lowry model, 9, 43, 162 Brownian motion, 3, 13, 16, 20, 36, 86, 109–111, 157

C Certainty, 36, 51, 54, 104 Chemistry textbooks, 16–19, 43, 44, 66, 77, 101, 110, 113, 118, 120, 126, 127, 130–134, 137–139, 141, 144, 146, 162, 188–194 Clash between comprehensive theories, 68 Coexistence of the old and the new paradigms, 12, 84 Competing paradigms, 6 Competing programs, 6 Competitive cross-validation, 37 Composition of air, 11, 66 Conceptual change theory, 53 Conceptual understanding, 40, 49, 51, 99, 113, 141, 143, 147 Conjectures, 5, 138, 148 Constructive alternativism, 82, 85–86 Constructivism, 10, 42, 46–48, 78, 100, 102, 122 Context of discovery, 6, 95 Contingency, 35 Copenhagen and Bohmian interpretations of quantum mechanics, 20, 35 Copernicanism, 8, 25, 31, 32 Copernican Revolution, 24, 106 Corrosive skepticism, 30 Counterinduction, 7–11, 13, 15–20, 24, 25, 31, 32, 36, 41, 46, 49, 51, 55, 61, 63, 64, 68, 73, 86, 88, 96, 103, 107, 109–154, 157–160, 163, 164, 168, 172 Covalent bonding, 19, 20, 44, 109, 144–147, 160, 162 Creationism, 10, 52, 62, 88

© Springer Nature Switzerland AG 2020 M. Niaz, Feyerabend’s Epistemological Anarchism, Contemporary Trends and Issues in Science Education 50, https://doi.org/10.1007/978-3-030-36859-3

221

222 Creativity, 46, 72, 73, 87, 89, 94, 120, 121, 135, 154, 167 Critical citizenry, 34 Critical thinking, 42, 48–49 Crucial experiments, 23, 122 Cumbrian sheep farmers, 29, 168 Current view may soon be voted out of office, 20, 59, 60, 122, 157, 160–162, 167, 172 Cutting-edge research, 6, 7 D Dappled world, 9, 32 Depthlessness, 33, 34 Dialectical tension, 46 Discarded theories from the history of science, 76 Diversity of methods, 42, 49, 157, 166–168, 173 Diversity of rival theories, 13, 82, 86 Dogmatism, 37, 41, 42, 59–61, 93, 161 Dualisms, 36, 107 Duhem-Quine thesis, 55 E Eclecticism, 6, 8, 32 Egalitarian romantic, 13, 87, 168 Einstein’s photoelectric equation, 17, 18, 53, 128–131, 159 Einstein’s special theory of relativity, 20, 91, 92, 115, 158 Elementary electrical charge, 15, 33, 67, 101, 103, 104, 116–118, 120, 158, 164, 166 Empirical evidence, 16, 19, 20, 35, 43, 52, 53, 115, 138, 149, 158, 160 English as a second language, 84 Enlightenment, 21, 34, 42, 50, 63, 87, 104, 165, 168, 171 Epistemic subject, 33, 36, 47, 99, 156 Epistemological anarchism, 1–69, 71–79, 81–107, 111, 122, 156–157, 160, 163, 169, 172, 173 Epistemological pluralism, 9, 20, 43, 162 Erosion of trust in science, 91 Euclidean geometry, 83 Evaluation, 7, 24, 30, 34, 39, 41, 60, 61, 65, 71–74, 81, 98, 102, 126, 142, 165 Evolution, 10, 21, 42, 48, 50–55, 62, 83, 104, 107, 160, 170, 171 Evolutionary theory, 10, 12, 20, 50, 53, 55, 62, 76

Subject Index Evolving nature of objectivity in the history of science, 13, 90 Explanatory power, 24, 27, 93 F Falsificationism, 23, 40, 76 Feminism, 36, 72, 74, 102 Feminist epistemology, 9, 36, 37 Feyerabend’s conception of ‘Being, 36 Final temple of science, 33, 122, 123 G Galilean idealization, 16, 112 Galilean revolution in astronomy, 6 Garbage dump of history, 84 General theory of relativity, 26 Genius in science, 13, 41, 82, 87, 157, 168 Geokinetic hypothesis, 30 Grammar of Science, 14, 90, 91 Grounded theory, 40 H Hard core, 23, 113 Harvard Project Physics, 88, 106 Hegelian historicism, 27 Heliocentric theory, 27 Heuristic inquiry approach, 53 Heuristic strategies, 40 History and philosophy of science (HPS), 2, 4, 7, 10, 14, 33, 43, 57, 58, 65, 67, 83, 88, 90–94, 97, 99, 101, 102, 105, 113, 141, 142, 161, 164 History of science, 6–8, 11, 13–16, 18, 19, 21, 24, 25, 32, 33, 36, 37, 46, 52, 53, 60, 74–76, 82, 88–91, 98–100, 103–106, 114, 124, 125, 134, 139, 142, 154, 164–166, 170–172 History of science becomes an inseparable part of the science itself, 11, 18, 21, 68, 134, 157, 163–164, 172 How science is done, 2, 6, 103 How science works, 8, 13, 23–38, 76, 84, 92, 93, 142, 154, 161 Hyperbolic flourishes, 19, 64, 110, 155–156 Hypothetico-deductivism, 49, 51, 65, 76 I Immutable truths, 13, 88, 161, 172 Incredulity to metanarratives, 33

Subject Index

223

Induction, 4, 10, 45, 58, 66, 95, 96, 140 Inductivism, 49, 59, 65, 76, 130 Inferring objectivity from empirical approaches, 73, 157, 165–166 Instrumentalism, 110 Intelligent design (ID), 10, 20, 21, 50, 51, 54, 160, 170 Interactional expertise, 29

Objectivity, 3, 5, 11, 13, 15, 21, 36, 40, 41, 51, 53, 73, 74, 92–94, 102–104, 107, 157, 165–166, 171, 172 Objectivity versus subjectivity, 82, 89, 90 Ocean of anomalies, 23, 170 Open-mindedness, 30 Owl of Minerva, 27 Oxygen theory, 11, 20, 68, 76, 164

L Lakatosian methodologist, 28 Law of conservation of mass, 67, 164 Lewis model, 9, 43, 145, 162 Local realism, 9, 32 Logic of scientific discovery, 4, 57 Logical positivism, 5, 9, 12, 20, 43, 78, 162

P Periodical table of expertises, 29 Permanent revolution, 32 Perspectival realism, 9, 14, 15, 32, 35, 92 Perspectival realist, 7, 9, 22, 35, 36, 92, 106, 107, 157, 171, 172 Perspectivism, 35, 36 Phlogiston theory, 11, 20, 21, 66, 68, 76, 163, 164 Piaget’s model of equilibration, 46 Planck’s constant, 17, 53, 125, 128, 159 Pluralism, 6, 7, 9, 14, 20, 21, 25, 32, 34, 37, 41, 43, 48, 64, 93, 95, 96, 103, 104, 106, 157, 161, 162, 166, 167, 169, 173 Pluralism of perspectives, 47 Pluralistic methodology, 8, 25, 33, 49 Pluralistic society, 28 Plurality of models, 46 Plurality of theories, 11, 45, 106, 107 Popperian fallibilism, 27 Positive heuristic, 23, 113 Postmodernism, 5, 7, 12, 15, 32–34, 50, 78, 98, 102–104 Practice of science, 7, 35, 37–38, 93, 161 Presuppositions of science teachers, 82, 90–94 Principia, 4, 5, 58, 87, 112 Progressive problem shifts, 24, 27 Proliferation of theories, 2, 5, 7, 11, 12, 41, 72, 73, 75–76, 167 Promiscuous realism, 9, 32 Protective belt, 23 Proton transfer model, 43, 44 Punctuated equilibrium, 74, 87, 167

M Masculine worldview, 11, 74 Methodological pluralism, 6, 7, 14, 21, 41, 48, 95, 96, 106, 157, 166, 167, 169, 173 Methodology of scientific research programs (MSRP), 4, 23, 26–28 Michelson-Morley experiment, 16, 20, 26, 45, 52, 109, 114–116, 157, 158 Milieu of the time, 52 Multitude of truths, 31 N Naïve falsificationism, 23, 40 Nature of science (NOS), 5, 9–12, 14, 15, 21, 42, 50, 52, 57, 60, 65, 72, 74–75, 77, 79, 90, 91, 98, 100–101, 104, 120–122, 131, 149, 157, 167, 168, 170, 171, 173 Negative heuristic, 23, 24, 113 New grew out of the old, 10, 21, 56, 157, 169, 170, 173 Newtonian mechanics, 1, 7, 55, 56, 94, 112, 170, 173 Newtonian method, 4, 5, 8, 10, 24, 42, 57–59, 89, 94, 136, 169, 173 Newtonian physics, 4, 60 Newtonian theory, 149 Newton’s gravitational theory, 24 Normal science, 2, 6, 11, 14, 20, 42, 59–61, 98–100, 160, 164 O Objective truth, 1 Objectivist realism, 92

Q Quantum mechanics, 1, 20, 35, 56, 57, 84, 91, 92, 94, 106, 161, 162, 169, 170, 173 Quantum model, 44

224 R Radical constructivism, 46, 47, 100, 102 Rationalism, 2, 4, 14, 19, 26, 29, 82, 94, 95, 155 Rationality, 1, 3, 4, 25, 26, 28, 36, 63, 67, 73, 94, 165 Realism, 9, 14, 15, 32, 34, 35, 51, 92, 100, 110 Refutability, 12, 77 Relativism, 6, 7, 32–34, 47, 49, 65 Relativity theory, 56, 92, 106, 114, 149, 161 Research program, 4, 12, 23, 24, 27, 28, 45, 46, 60, 83, 84, 99, 112, 113, 125, 128, 145, 153 Rival research programs, 12, 23, 83, 84 S Schrödinger equation, 57 Science and religion, 42, 51, 62–63 Science curriculum, 10, 15, 36, 42, 52, 62, 66, 68, 98, 160–163 Science education, 2, 5–9, 11–15, 20, 26, 36, 39, 42, 46–48, 50, 51, 54–57, 59–61, 65, 69, 72–75, 77, 79, 81, 85, 88, 89, 91, 93, 99, 100, 102, 105, 106, 142, 155, 157, 160, 161, 164, 168, 170, 171 Science For All Americans, 5 Science textbooks, 7, 8, 15–17, 20, 76, 88, 90–92, 100, 106, 109–154, 157, 158, 164 Scientific expertise, 8, 21, 28–30, 42, 63–64, 157, 168, 173 Scientific method, 2, 3, 8, 9, 14, 15, 21, 25, 28, 35, 38, 40, 42, 44, 45, 65–68, 72, 73, 75–77, 82, 90, 94–96, 99–101, 117, 119–121, 157, 158, 165, 167–169, 173 Scientific perspectivism, 36 Scientific progress, 5, 11, 15, 17, 21, 32, 40, 47, 48, 50, 52, 57, 66, 89, 99, 102, 120, 122, 123, 131, 145, 146, 159, 165 Scientism, 9, 10, 21, 30, 41, 51, 59, 60, 62, 171 Scrabblers, 38 Sheepfarming after Chernobyl, 29 Situated learning, 11, 42, 68, 69 Skeletons in the Newtonian cupboard, 21, 58–59, 157, 169 Skepticism, 2, 3, 8, 30, 85, 171 Slippery slopes of postmodernism, 5, 78, 102

Subject Index Social constructivism, 46, 47 Solipsism, 48 Sophisticated falsificationism, 24 Special theory of relativity (STR), 16, 20, 26, 91, 92, 114–116, 157–158 Suspend belief, 68 Suspend our judgment, 67 T Tacit knowledge, 11, 21, 29, 42, 61–62, 103, 169, 173 Tau lepton, 19, 33, 110, 150–154 Teacher demonstrations, 12, 72, 77–78 Tentative nature of science, 12, 20, 43, 44, 49, 57, 74, 160, 167, 172 Theory-laden nature of scientific observations, 8, 26 Theory of electrolytic dissociation, 42 Thermodynamics, 10, 13, 55, 56, 86, 110, 167 Tight-rope walker, 34 Transgression of categorization, 9, 41, 48 Transgression of method, 3, 8 Transmission view of learning, 78 Trojan horse, 4 Tro’s scenario, 44 Truths, 1, 5, 13–15, 26, 31, 33, 36, 45, 48, 54, 62, 73, 85, 88, 90–95, 104, 106, 107, 156, 161, 165, 167, 172 Truths out there to be discovered, 91, 92 Truths to be memorized, 73 U Uncertainty, 12, 21, 46, 54, 77, 152–154, 168, 173 Unitarian view of science, 9, 41, 45 V Vedic Science, 63 Violating categories, 9, 37, 48, 167, 172 W Wave theory of light, 17, 18, 53, 129, 132, 159 Working hypotheses, 12, 13, 75, 76, 88, 143, 153, 161, 172 Worldviews, 11, 50, 66, 72, 74, 78–79, 164