Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

Enter the email address you signed up with and we'll email you a reset link.

  • We're Hiring!
  • Help Center

paper cover thumbnail

Review of Scientific Realism: Selected Essays of Mario Bunge, by Martin Mahner

Profile image of Peter Denton

Essays in Philosophy

Related Papers

Peter Denton

scientific realism essay

[email protected] Boyd

Anjan Chakravartty

Scientific realism is the view that our best scientific theories give approximately true descriptions of both observable and unobservable aspects of a mind-independent world. Debates between realists and their critics are at the very heart of the philosophy of science. Anjan Chakravartty traces the contemporary evolution of realism by examining the most promising strategies adopted by its proponents in response to the forceful challenges of antirealist sceptics, resulting in a positive proposal for scientific realism today. He examines the core principles of the realist position, and sheds light on topics including the varieties of metaphysical commitment required, and the nature of the conflict between realism and its empiricist rivals. By illuminating the connections between realist interpretations of scientific knowledge and the metaphysical foundations supporting them, his book offers a compelling vision of how realism can provide an internally consistent and coherent account of...

James Ladyman

Gerhard Schurz

Curtis J Forbes

Volume 9 of Spontaneous Generations: A Journal for the History and Philosophy of Science presents an eclectic but surprisingly harmonious collection of invited and peer-reviewed papers, organized under the title “The Future of the Scientific Realism Debate: Contemporary Issues Concerning Scientific Realism.” Curtis Forbes's editor's introduction starts off the collection by tracing some of the broader themes that unite the pieces. It is followed by a dialogue between Bas van Fraassen and Anjan Chakravartty entitled "What is Scientific Realism?", which sets the tone for the twenty one essays that follow: all quite stimulating, sometimes perplexed, and often freely speculative. In order of appearance, there are contributions from Jeff Foss, Hasok Chang, Theodore Arabatzis, Harry Collins, Arthur Fine, Joseph Rouse, Alan Musgrave, Howard Sankey, Stathis Psillos, P. Kyle Stanford, Jamie Shaw, James Ladyman, Robin Hendry, Pete Vickers, Mario Alai, Kerry McKenzie, K. Brad Wray, Tim Lyons, Paul Teller, Nancy Cartwright, as well as Cliff Hooker and Giles Hooker. The journal's usual distinction between full length articles, focused discussion pieces, and shorter opinion statements was suspended to give invited contributors maximal freedom to speak their mind. Some chose to share a new and carefully articulated, lengthy argument. Others chose to briefly clarify some previous work of theirs, or offer a preface to something forthcoming. Still others raise new problems for the would-be scientific realist, or their opponents, or boldly proffer some new terms the debate might, will, or should be based on moving forward. The result is a kind of snapshot that records some ways these specific philosophers of science are currently thinking about the state, history, and future of the debate over scientific realism. Much of it is reflective and opinionated, so the reader deserves some clarification: many of the ideas found here deserve to be fleshed out more fully elsewhere, and in most cases the authors have either referenced prior publications where they have done so, or mentioned their plans to do so soon. This special issue has given various parties to the scientific realism debate the opportunity to describe where they presently stand, and where they think the discussion should be going. The reader, however, will likely need to consult separate works to fully understand why each author stands where they do, and wait to see where the debate over scientific realism goes moving forward. As a whole, therefore, this collection is probably best treated as a touchstone or launching point for further study and discussion of many contemporary issues concerning scientific realism. Enjoy.

New Approaches to Scientific Realism,

Wenceslao J. Gonzalez

Scientific realism plays a central role in the philosophico-methodological discussions on research. There are two are the main directions in the contributions made to scientific realism: the “internal” line and the “external” path. Following the first line, there are new visions of realism focused on central aspects of science: semantic, logic, epistemological, methodological, ontological, axiological, and ethical components. When the route follows the second path, realism in science is seen as interrelated with realism in technology and as connected to a philosophical approach to society. Altogether there is now a plethora of characterizations of scientific realism, and this paper presents the main contemporary versions of scientific realism. The analysis of the central tenets of the recent views on scientific realism contributes to present this book, which offers novelty to the ongoing debate on scientific realism.

THE ROUTLEDGE COMPANION TO PHILOSOPHY OF SCIENCE

HARRIS ( K O N S T A N T I N O S ) NOTTAS , CO AUTHORS DTMS ET AL D T M S AS WELL AS MANY OTHER ANAGRAMS

A WORTHWHILE REVIEW 648 PAGES TO BE ENJOYED ON A QUIET NIGHT OR TWO The Routledge Companion to Philosophy of Science is an outstanding guide to the major themes, movements, debates and topics in philosophy of science. Fifty-five entries by a team of renowned international contributors are organized into four parts: • Historical and Philosophical Context • Debates • Concepts • Individual Sciences The Companion begins with a critical examination of how philosophy of science has been involved in a mutually fruitful interaction with philosophical theories in areas such as metaphysics, epistemology, and the philosophy of language, and reassesses the major schools of philosophy of science in the twentieth century. The second part explores the development of current debates among philosophers and scientists on issues such as confirmation, explanation, realism, scientific method, and the ethics of science. Part three discusses controversial concepts such as causation, prediction, unification, observation, and probability that lie at the heart of many disputes about science and scientific theories. The final part addresses some of the main philosophical problems that arise within eight branches of science: biology, chemistry, cognitive science, economics, mathematics, physics, psychology, and the social sciences. The Routledge Companion to Philosophy of Science is essential reading for anyone interested in philosophy of science and the connections between philosophy and the natural and social sciences. Stathis Psillos is an Associate Professor of Philosophy of Science at the University of Athens, Greece. He is the author of Scientific Realism: How Science Tracks Truth (Routledge), Causation and Explanation and Philosophy of Science A–Z. Martin Curd is an Associate Professor of Philosophy at Purdue University, USA. He is co-editor (with Jan Cover) of Philosophy of Science: The Central Issues. PART I Historical and philosophical context 1 1 The epistemology of science after Quine 3 PAUL A. ROTH 2 The history of philosophy and the philosophy of science 15 JOANNE WAUGH AND ROGER ARIEW 3 Metaphysics 26 Stephen Mumford 4 Philosophy of language 36 Rod Bertolet 5 The role of logic in philosophy of science 47 Diderik Batens 6 Critical rationalism 58 Gürol Irzik 7 The historical turn in the philosophy of science 67 ALEXANDER BIRD 8 Logical empiricism 78 Thomas Uebel 9 Pragmatism and science 91 Robert Almeder Part II Debates 101 10 Bayesianism 103 Colin Howson 11 Confirmation 115 ALAN HÁJEK AND JAMES M. JOYCE 12 Empiricism 129 Elliott Sober vi 13 Essentialism and natural kinds 139 BRIAN ELLIS 14 Ethics of science 149 David B. Resnik 15 Experiment 159 Theodore Arabatzis 16 Explanation 171 JAMES Woodward 17 The feminist approach to the philosophy of science 182 CASSANDRA L. PINNICK 18 Inference to the best explanation 193 Peter Lipton 19 Laws of nature 203 Marc Lange 20 Naturalism 213 RONALD N. GIERE 21 Realism/anti-realism 224 Michael Devitt 22 Relativism about science 236 Maria Baghramian 23 Scientific method 248 Howard Sankey 24 Social studies of science 259 ROBERT NOLA 25 The structure of theories 269 Steven French 26 Theory-change in science 281 John Worrall 27 Underdetermination 292 Igor Douven 28 Values in science 302 GERALD DOPPELT vii Part III Concepts 315 29 Causation 317 Christopher Hitchcock 30 Determinism 327 Barry Loewer 31 Evidence 337 Peter Achinstein 32 Function 349 D. M. WALSH 33 Idealization 358 James Ladyman 34 Measurement 367 HASOK CHANG AND NANCY CARTWRIGHT 35 Mechanisms 376 Stuart Glennan 36 Models 385 DEMETRIS PORTIDES 37 Observation 396 ANDRé KUKLA 38 Prediction 405 MALCOLM FORSTER 39 Probability 414 Maria Carla Galavotti 40 Reduction 425 Sahotra Sarkar 41 Representation in science 435 PAUL TELLER 42 Scientific discovery 442 Thomas Nickles 43 Space and time 452 OLIVER POOLEY 44 Symmetry 468 Margaret Morrison viii 45 Truthlikeness 478 Graham Oddie 46 Unification 489 TODD JONES 47 The virtues of a good theory 498 Ernan McMullin Part IV Individual sciences 509 48 Biology 511 Alexander Rosenberg 49 Chemistry 520 Robin FINDLAY Hendry 50 Cognitive science 531 Paul Thagard 51 Economics 543 USKALI MÄKI 52 Mathematics 555 PETER CLARK 53 Physics 567 SIMON SAUNDERS 54 Psychology 581 RICHARD SAMUELS 55 Social sciences 594 HAROLD KINCAID Index 605 THEN AS A COMPLEMENTARY READING https://www.academia.edu/2807083/In_defence_of_scientism

David Harker

Miguel ESPINOZA

RELATED TOPICS

  •   We're Hiring!
  •   Help Center
  • Find new research papers in:
  • Health Sciences
  • Earth Sciences
  • Cognitive Science
  • Mathematics
  • Computer Science
  • Academia ©2024

SEP home page

  • Table of Contents
  • Random Entry
  • Chronological
  • Editorial Information
  • About the SEP
  • Editorial Board
  • How to Cite the SEP
  • Special Characters
  • Advanced Tools
  • Support the SEP
  • PDFs for SEP Friends
  • Make a Donation
  • SEPIA for Libraries
  • Entry Contents

Bibliography

Academic tools.

  • Friends PDF Preview
  • Author and Citation Info
  • Back to Top

Realism and Theory Change in Science

Scientific theories seem to have an expiry date. If we look at the history of science, a number of theories that once were dominant and widely accepted are currently taught in history of science courses. Will this be the fate of current scientific theories? Is there a pattern of radical theory-change as science grows? Are theories abandoned en bloc ? Or are there patterns of retention in theory-change? That is, are some parts of theories more likely to survive than other parts? And what are the implications of all this for the scientific image of the world?

These kinds of question have played a major role in the scientific realism debate. The challenge to scientific realism is supposed to come directly from the history of science. The history of science, it is claimed, is at odds with scientific realism’s epistemic optimism. It is full of theories which were shown to be false and abandoned, despite their empirical successes. Hence, it is claimed, realists cannot be warrantedly optimistic about the (approximate) truth of currently empirically successful theories. If we take the historical evidence seriously, it is claimed, current theories too will, sooner or later, be abandoned and take their place in future history-of-science courses. This anti-realist line of argument has become known as ‘the pessimistic induction’ (aka pessimistic meta-induction)—henceforth PI . Without denying that theories change over time, scientific realists have tried to block this line of argument by showing either that it is fallacious or that there is substantive continuity in theory-change which warrants the realist’s optimism that current science is on the right track.

This entry discusses the origin and current state of the historical challenge to realism and the various realist reactions to it. The first part focuses on the first enactment of arguments based on historical pessimism, as these appeared in the so-called ‘bankruptcy of science controversy’ in the end of the nineteenth century.

The second part deals with the historical challenge to scientific realism as this is currently formulated and the various lines of defense of the claim that scientific knowledge grows despite theory-change.

1.1 The Bankruptcy-of-science Debate

1.2 duhem on continuity, 1.3 poincaré’s relationism, 1.4 boltzmann against historical pessimism, 2.1 the ‘disastrous meta-induction’, 2.2 the principle of no privilege, 2.3 getting nearer to the truth, 2.4 the plethora of false theories, 2.5 the divide et impera strategy, 2.6 criticisms of divide et impera, 2.7 structural realism, 2.8 induction or deduction, 2.9 a new induction, other internet resources, related entries, 1. the history of the historical challenge.

The issue of theory-change in science was debated in the context of the ‘bankruptcy of science’ controversy that was raging in Paris in the last decade of the nineteenth century and the first decade of the twentieth. A claim of growing popular reputation among various public intellectuals, spearheaded by Ferdinand Brunetière and Leo Tolstoy, was that scientific theories are ephemeral; and this was supposed to prove that science has at best no more than predictive value with no legitimate claim to showing what the world is like—especially in its unobservable aspects. In light of a growing interest in the history of science among scientists and philosophers, it was pointed out that science has had a poor track record: it has gone through many radical theory-changes in the past; hence, there is reason to believe that what is currently accepted will be overturned in the future.

In his essay “The Non-Acting”, published in French in August 1893, the Russian novelist Tolstoy (1828–1910) noted:

Lastly, does not each year produce its new scientific discoveries, which after astonishing the boobies of the whole world and bringing fame and fortune to the inventors, are eventually admitted to be ridiculous mistakes even by those who promulgated them? (…) Unless then our century forms an exception (which is a supposition we have no right to make), it needs no great boldness to conclude by analogy that among the kinds of knowledge occupying the attention of our learned men and called science, there must necessarily be some which will be regarded by our descendants much as we now regard the rhetoric of the ancients and the scholasticism of the Middle Ages. (1904: 105)

A few years earlier, in 1889, Ferdinand Brunetière (1849–1906), Professor at the École Normale Supérieure and editor of the prestigious journal Revue des Deux Mondes , noted in his review of Paul Bourget’s play ‘ Le Disciple ’:

We differ from animals in recognizing that humans have to be first (i.e., they have value). The laws of nature, the ‘struggle for life’ or ‘natural selection’, do not show what we have in common. Are these the only laws? Do we know whether perhaps tomorrow they will not join in the depths of oblivion the Cartesian vortices or the ‘quiddities’ of scholasticism? (1889: 222, author’s translation)

This history-fed pessimism about science, which seemed to capture the public mood, led to a spirited reaction by the scientific community. In an anonymous article that appeared in Revue Scientifique , a prestigious semi-popular scientific journal, in August 17 1889, the following questions were raised: Is the history of science the history of human error? Will what theories affirm today be affirmed in a century or two? The reply was:

We will say to savants, philosophers and physicists, physicians, chemists, astronomers or geologists: Go forward boldly, without looking behind you, without caring for the consequences, reasonable or absurd, that can be drawn from your work. Seek the truth, without the worry of its applications. (Anonymous 1889: 215, author’s translation)

A few years later, in 1895, Brunetière strikes back with an article titled ‘ Après Une Visite Au Vatican ’, published in Revue des Deux Mondes , by claiming that science is bankrupt:

Science has failed to deliver on its promise to change ‘the face of the world’. (...) Even if this is not a total bankruptcy, it is certainly a partial bankruptcy, enough to shake off the credit from science. (1895: 98, 103)

The eminent scientist Charles Richet (1850–1935), Professor of Physiology at the Collège de France, Editor of Revue Scientifique and Nobel Laureate for Medicine in 1913, replied with an article titled ‘La Science a-t-elle fait banqueroute?’ ( Revue Scientifique , 12 January 1895), which appeared in the section: Histoire des Sciences . In this, he did three things. Firstly, he noted that science can never understand the ‘why’ (‘ le pourquoi ’) of things, especially when it comes to the infinitely small and the infinitely large. Science “attends only to the phenomena. The intimate nature of things escapes from us” (1895: 34). Secondly, he stressed that “science has not promised anything”, let alone the discovery of the essence of things. Thirdly, he added that despite the fact that science has made no promises, it has changed the world, citing various scientific, industrial and technological successes (from the invention of printing and the microscope to the railways, the electric battery, the composition of the air, and the nature of fermentation).

Turning Brunetière’s argument on its head, Richet formulated what might be called an ‘optimistic induction’ based on the then recent history of scientific successes. To those who claim that science has failed in the past, his reply is that history shows that it is unreasonable to claim for any scientific question that we will always fail to answer it. Far from warranting epistemic pessimism, the history of science is a source of cognitive optimism. Richet referred to a few remarkable cases, the most striking of which is the case of Jean Louis Prevost and Jean Baptiste Dumas, who had written in 1823:

The pointlessness of our attempts to isolate the colouring matter of the blood gives us almost the certainty that one will never be able to find it. (1823, 246, author’s translation)

Forty years after their bold statement, Richet exclaimed, this coloured matter (haemoglobin) had been isolated, analysed and studied.

Richet’s reply to the historical challenge suggested lowering the epistemic bar for science: science describes the phenomena and does not go beyond them to their (unobservable) causes. This attitude was echoed in the reply to the ‘bankruptcy charge’ issued by the eminent chemist and politician of the French Third Republic, Marcelin Berthelot (1827–1907) in his pamphlet Science et Morale in 1897. He was firm in his claim that the alleged bankruptcy of science is an illusion of the non-scientific mind. Like Richet, he also argued that science has not pretended to have penetrated into the essence of things: “under the words ‘essence’, ‘the nature of things’, we hide the idols of our own imagination” (1897: 18, author’s translation). Science, he noted, has as its starting point the study of facts and aims to establish general relations, that is, ‘scientific laws’, on their basis. If science does not aim for more, we cannot claim that it is bankrupt; we cannot accuse it for “affirmations it did not make, or hopes it has not aroused”. [ 1 ]

Berthelot, who objected to atomism, captured a broad positivist trend in French science at the end of the nineteenth century, according to which science cannot offer knowledge of anything other than the phenomena. In light of this view, the history-fed pessimism is misguided precisely because there has been substantial continuity at the level of the description of the phenomena, even if explanatory theories have come and gone.

This kind of attitude was captured by Pierre Duhem’s (1906) distinction between two parts of a scientific theory: the representative part, which classifies a set of experimental laws; and the explanatory part, which “takes hold of the reality underlying the phenomena” (1906 [1954: 32]). Duhem understood the representative part of a theory as comprising the empirical laws and the mathematical formalism, which is used to represent, systematize and correlate these laws, while he thought that the explanatory part relates to the construction of physical (and in particular, mechanical) models and explanatory hypotheses about the nature of physical processes which purport to reveal underlying unobservable causes of the phenomena. For him, the explanatory part is parasitic on the representative. To support this view, he turned to the history of science, especially the history of optical theories and of mechanics. He argued that when a theory is abandoned because it fails to cover new experimental facts and laws, its representative part is retained, partially or fully, in its successor theory, while the attempted explanations offered by the theory get abandoned. He spoke of the “constant breaking-out of explanations which arise to be quelled” (1906 [1954: 33]).

Though Duhem embedded this claim for continuity in theory-change in an instrumentalist account of scientific theories, he also took it that science aims at a natural classification of the phenomena, where a classification (that is the representation of the phenomena within a mathematical system) is natural if the relations it establishes among the phenomena gathered by experiments “correspond to real relations among things” (1906 [1954: 26–27]). Hence, scientific knowledge does go beyond the phenomena but in doing so, that is, in tending to be a natural classification, it can extend only up to relations among “hidden realities whose essence cannot be grasped” (1906 [1954: 297]). A clear mark of the naturalness of a classification is when it issues in novel predictions (1906 [1954: 28]). Hence, successful novel predictions issued by a theory are a mark for the theory getting some aspects of reality right, viz. real relations among unobservable entities. [ 2 ]

This kind of relationism became a popular middle way between positivism and what may be called full-blown realism. Duhem himself, justly, traced it back to his contemporary Henri Poincaré. He noted with approval that Poincaré “felt a sort of revolt” against the proposition that “theoretical physics is a mere collection of recipes” and he “loudly proclaimed that a physical theory gives us something else than the mere knowledge of the facts, that it makes us discover the real relations among things ([1906] 2007: 446; improved translation from the French original by Marie Guegeun and the author).

In his address to the 1900 International Congress of Physics in Paris, Poincaré made a definitive intervention in the bankruptcy-of-science debate and its history-fed pessimism. He described the challenge thus:

The people of world [ les gens du monde ] are struck to see how ephemeral scientific theories are. After some years of prosperity, they see them successively abandoned; they see ruins accumulated on ruins; they predict that the theories in fashion today will quickly succumb in their turn, and they conclude that they are absolutely futile. This is what they call the bankruptcy of science (1900: 14, author’s translation).

The view of ‘the people of the world’ is not right:

Their scepticism is superficial; they understand none of the aim and the role of scientific theories; otherwise they would understand that ruins can still be good for something.

But unlike the positivist trend around him, Poincaré took it that scientific theories offer knowledge of the relational structure of the world behind the phenomena. In the Introduction to La Science et l’Hypothése in 1902, he made clear what he took to be the right answer to the historical challenge:

Without doubt, at first, the theories seem to us fragile, and the history of science proves to us how ephemeral they are; yet they do not entirely perish, and of each of them something remains. It is this something we must seek to unravel, since there and there alone is the true reality. (1902: 26, author’s translation)

Poincaré argued that what survives in theory-change are relations among physical magnitudes, expressed by mathematical equations within theories. His prime example was the reproduction of Fresnel’s laws concerning the relations of amplitudes of reflected rays vis-à-vis the amplitude of incident rays in the interface of two media within Maxwell’s theory of electromagnetism, although in this transition, the interpretation of these laws changed dramatically, from an ether-based account to an electromagnetic-field-based account. For Poincaré

These equations express relations, and if the equations remain true it is because these relations preserve their reality. They teach us, before and after, that there is such and such a relation between some thing and some other thing; only this something we used to call motion , we now call it electric current . But these names were only images substituted for the real objects which nature will eternally hide from us. The true relations between these real objects are the only reality we can attain to, and the only condition is that the same relations exist between these objects as between the images by which we are forced to replace them. If these relations are known, what does it matter if we deem it convenient to replace one image by another? (1900: 15, author’s translation.

In recent literature, Poincaré’s line of thought has come to be known as structural realism , though it may be best if we describe it as ‘relationism’. In the Introduction to La Science et l’Hypothése , he noted that

the things themselves are not what it [science] can reach, as the naive dogmatists think, but only the relations between things. Apart from these relations there is no knowable reality. (1902: 25, author’s translation)

It should be stressed that Poincaré does not deny that there is reality outside relations; but he does deny that this reality is knowable . Note also that Poincaré does not use the expression ‘things in themselves’ ( choses en soi ) but the expression ‘things themselves’ ( chose elles-memes ). Elsewhere he talks about the “nature of things” or “real objects”. It is quite clear that he wanted to draw a distinction between how things are—what their nature is—and how they are related to each other (and to us qua knowers). A plausible way to draw this distinction is to differentiate between the intrinsic and perhaps fully qualitative properties of things—what he plausibly calls ‘nature’ of things—and their relations. The former are unknowable, whereas the latter are knowable. [ 3 ]

So, Poincaré and Duhem initiated a strategy for dealing with theory-change in science which pointed to substantial continuities among successive theories. For them, the continuity is, by and large, relational (and in this sense mathematical ). Hence, mathematically-convergent scientific theories reveal the relational structure of the world.

This relational answer to historical pessimism was motivated, at least partly, by the widespread scepticism towards the atomic theory of matter. Atomism posited the existence of unobservable entities—the atoms—to account for a host of observable phenomena (from chemical bonding to Brownian motion). A trend among scientists opposed to the explanation of the visible in terms of the invisible was what Ludwig Boltzmann called “phenomenologists” (which included the early Max Planck), according to whom the aim of science was to “write down for every group of phenomena the equations by means of which their behavior could be quantitatively calculated” (Boltzmann 1901: 249). The theoretical hypotheses from which the equations might have been deduced were taken to be the scaffolding that was discarded after the equations were arrived at. For phenomenologists, then, hypotheses are not unnecessary or useless—rather they have only a heuristic value: they lead to stable (differential) equations and that’s it .

According to Boltzmann, a motivation for this phenomenological attitude was the “historical principle”, viz., that hypotheses are essentially insecure because they tend to be abandoned and replaced by others, “totally different” ones. As he put it:

frequently opinions which are held in the highest esteem have been supplanted within a very short space of time by totally different theories; nay, even as St. Remigius the heathens, so now they [the phenomenologists] exhorted the theoretical physicists to consign to the flames the idols that but a moment previously they had worshipped (1901: 252–253).

Like Poincaré, Boltzmann’s answer to historical pessimism was that despite the presence of “revolutions” in science, there is enough continuity in theory change to warrant the claim that some “achievements may possibly remain the possession of science for all time” (1901: 253). But unlike Poincaré, Boltzmann did not restrict the criterion of invariance-in-theory-change to relations only: The answer to the historical challenge is to look for patterns of continuity in theory change . In fact, as Boltzmann noted, if the historical principle is correct at all, it cuts also against the equations of the phenomenologists. For unless these very equations remain invariant through theory-change, there should be no warrant for taking them to be accurate descriptions of worldly relations (cf. 1901: 253). Besides, Boltzmann noted, the very construction of the differential equations of the phenomenologists requires commitment to substantive atomistic assumptions. Hence, the phenomenologists are not merely disingenuous when they jettison the atomistic assumptions after the relevant differential equations have been arrived at. Their move is self-undermining. In light of the historical principle, the success of the mathematical equations would lead to their defeat, since the very theory that led to this success would fall foul of the historical principle: it would have to be abandoned.

The history-based pessimism (and the relevant debate) came to an end by the triumph of atomism in the first decade of the twentieth century. Due to the work of Albert Einstein and the French physicist Jean Perrin on the atomic explanation of Brownian motion, one after the other of major scientists who were initially sceptical about the atomic conception of matter came to accept atomism. [ 4 ] The French philosopher André Lalande captured this point in his 1913 (pp. 366–367) thus:

M. Perrin, professor of physics at the Sorbonne, has described in Les Atomes , with his usual lucidity and vigour, the recent experiments (in which he has taken so considerable a part) which prove conclusively that the atoms are physical realities and not symbolical conceptions as people have for a long time been fond of calling them. By giving precise and concordant measures for their weights and dimensions, it is proved that bodies actually exist which, though invisible, are analogous at all points to those which we see and touch. An old philosophical question thus receives a positive solution.

Be that as it may, what this brief account of the history of the historical challenge to realism reveals are the two major lines of defense of realism at play. Both lines of defense are based on the presence of substantial continuity in theory-change in the history of science. This continuity suggests that the disruption of the scientific image of the world, as theories change, is less radical than is assumed by the historical challenge to realism. But the two lines of defense (the Poincaré-Duhem and the Boltzmann one) disagree over what is retained when theories change. The Poincaré-Duhem line of defense focuses on mathematical equations (which express relations) and claims that only relations among unobservable things are knowable, whereas the Boltzmann line of defense focuses on whatever theoretical elements (including entities like atoms) are retained while theories change; hence, it does not limit scientific knowledge to the knowledge of relations only. Both lines have resurfaced in the current debate.

2. Scientific Realism and the Pessimistic Induction

Capitalizing on the work of Richard Boyd, the early Hilary Putnam took scientific realism to involve three theses:

  • Theoretical terms refer to unobservable entities;
  • Theories are (approximately) true; and
  • There is referential continuity in theory change.

Putnam argued that the failure of the third thesis would lead to a disastrous “meta-induction”:

just as no term used in the science of more than fifty (or whatever) years ago referred, so it will turn out that no term used now (except maybe observation terms, if there are such) refers (1978: 25) (emphasis in the original).

An answer to this ‘disastrous’ history-fed argument was the development of a causal theory of reference, which allows for referential continuity in theory-change. This theory was first suggested by Saul Kripke (1972) as an alternative to the then dominant descriptive theories of reference of proper names and was extended by Putnam (1973, 1975) to cover natural kind terms and theoretical terms. According to the causal theory, the reference of a theoretical term t is fixed during an introducing event in which an entity or a physical magnitude is posited as the cause of various observable phenomena. The term t , then, refers to the posited entity. Though some kind of descriptions of the posited entity will be associated with t , they do not play a role in reference fixing. The referent has been fixed existentially : it is the entity causally responsible for certain effects.

The causal theory of reference makes it possible that the same term featuring in different theories refers to the same worldly entity. If, for instance, the referent of the term ‘electricity’ is fixed existentially, all different theories of electricity refer to, and dispute over, the same ‘existentially given’ magnitude, viz. electricity ; better, the causal agent of salient electrical effects. Hence, the causal theory makes available a way to compare past and present theories and to claim that the successor theory is more truthlike than its predecessors since it says truer things of the same entities. It turns out, however, that the causal theory faces a number of conceptual problems, most notable of which is that it makes referential success inevitable insofar as the phenomena which lead to the introduction of a new theoretical term do have a cause (see Psillos 1999: chapter 11 for a discussion). Philosophers of science have tried to put forward a causal-descriptive theory of reference which makes referential continuity possible whilst allowing room for causal descriptions in fixing the reference of a theoretical term. [ 5 ]

An analogous history-fed pessimistic argument can be based on the so-called “principle of no privilege”, which was advanced by Mary Hesse in her 1976. According to this principle:

our own scientific theories are held to be as much subject to radical conceptual change as past theories are seen to be. (1976: 266)

This principle can be used for the derivation of the strong conclusion that all theories are false . As Hesse put it:

Every scientific system implies a conceptual classification of the world into an ontology of fundamental entities and properties—it is an attempt to answer the question “What is the world really made of?” But it is exactly these ontologies that are most subject to radical change throughout the history of science. Therefore in the spirit of the principle of no privilege, it seems that we must say either that all these ontologies are true, ie: we must give a realistic interpretation of all of them or we must say they are all false. But they cannot all be true in the same world, because they contain conflicting answers to the question “What is the world made of?” Therefore they must all be false. (1976: 266)

This argument engages the history of theory-change in science in a substantial way. As Hesse admitted, the Principle of No Privilege arises “from accepting the induction from the history of science” (1976: 271). Hesse’s argument starts with the historical premise that, as science grows over time, there has been a recognizable pattern of change in the ‘ontology of fundamental entities and properties’ posited by scientific theories. Assuming, then, the Principle of No Privilege, it is argued that current theories too will be subjected to a radical change in the ontology of the entities and properties they posit. Hence, current theories are as false as the past ones.

The problem with this kind of argument is that the historical premise should be borne out by the actual history of theory-change in science. It’s not enough to say that scientific theories change over time; these changes should be such that the newer theories are incompatible with the past ones. Or, to use Hesse’s idiom, it should be shown that past and current scientific ‘ontologies’ are incompatible with each other. Showing incompatibility between the claims made by current theory T and a past theory T ′ requires a theory of reference of theoretical terms which does not allow that terms featuring in different theories can nonetheless refer to the same entity in the world. Hence, it is question-begging to adopt a theory of reference which makes it inevitable that there is radical-reference variance in theory-change.

Referential stability, as noted already, makes possible the claim that past and present ontologies are compatible, even if there have been changes in what current theories say of the posited entities. The “revolutionary induction from the history of science about theory change” (Hesse 1976: 268) can be blocked by pointing to a pattern of substantial continuity in theory change.

Can a history-fed argument be used in defence of realism? William Newton-Smith (1981) was perhaps the first in the recent debate to answer positively this question. Scientific realism is committed to the two following theses:

  • theories are true or false in virtue of how the world is, and
  • the point of the scientific enterprise is to discover explanatory truths about the world.

According to Newton-Smith, (2) is under threat “if we reflect on the fact that all physical theories in the past have had their heyday and have eventually been rejected as false”. And he added:

Indeed, there is inductive support for a pessimistic induction: any theory will be discovered to be false within, say 200 years of being propounded. We may think of some of our current theories as being true. But modesty requires us to assume that they are not so. For what is so special about the present? We have good inductive grounds for concluding that current theories—even our most favourite ones—will come to be seen to be false. Indeed the evidence might even be held to support the conclusion that no theory that will ever be discovered by the human race is strictly speaking true. So how can it be rational to pursue that which we have evidence for thinking can never be reached? (1981: 14)

The key answer to this question is that even if truth cannot be reached, it is enough for the defense of realism to posit “an interim goal for the scientific enterprise”, viz., “the goal of getting nearer the truth”. If this is the goal, the “sting” of the preceding induction “is removed”. Accepting PI “is compatible with maintaining that current theories, while strictly speaking false, are getting nearer the truth” (1981: 14).

But aren’t all false theories equally false? The standard realist answer is based on what Newton-Smith called “the animal farm move” (1981: 184), viz., that though all theories are false, some are truer than others. Hence, what was needed to be defended was the thesis that if a theory \(T_2\) has greater verisimilitude than a theory \(T_1\), \(T_2\) is likely to have greater observational success than \(T_1\). The key argument was based on the “undeniable fact” that newer theories have yielded better predictions about the world than older ones (cf. Newton-Smith 1981: 196). But if the ‘greater verisimilitude’ thesis is correct (that is, if theories “are increasing in truth-content without increasing in falsity-content”), then the increase in predictive power would be explained and rendered expectable. This increase in predictive power “would be totally mystifying (…) if it were not for the fact that theories are capturing more and more truth about the world” (1981: 196).

The key point, then, is that the defense of realism against the historical induction requires showing that there is, indeed, a privilege that current theories enjoy over past ones, which is strong enough to block transferring, on inductive grounds, features of past theories to current ones. For most realists, the privilege current theories enjoy over past ones is not that they are true while the past theories are false. Rather, the privilege is that they are more truthlike than past theories because they have had more predictive power than past theories. The privilege is underpinned by an explanatory argument: the increasing truthlikeness of current theories best explains their increasing predictive and empirical success.

But there is a way to see the historical challenge to realism which makes it have as its target precisely to undercut the explanatory link between empirical success and truthlikeness . This was brought under sharp relief in the subsequent debates.

The most famous history-based argument against realism, issued by Larry Laudan (1981), was meant to show how the explanatory link between success and truthlikeness is undermined by taking seriously the history of science. It should be noted that Laudan’s argument has been subjected to several diverging interpretations, which will be the focus of section 2.8. For the time being let’s stick to a particularly popular one, according to which Laudan argues inductively from the falsity of past theories to the falsity of current ones. This argument may be put thus:

Laudan substantiated (L) by means of what he has called “the historical gambit”: the following list—which “could be extended ad nauseam ”—gives theories which were once empirically successful and fruitful, yet just false.

Laudan’s list of successful-yet-false theories

  • the crystalline spheres of ancient and medieval astronomy
  • the humoral theory of medicine
  • the effluvial theory of static electricity
  • catastrophist geology, with its commitment to a universal (Noachian) deluge
  • the phlogiston theory of chemistry
  • the caloric theory of heat
  • the vibratory theory of heat
  • the vital force theory of physiology
  • the theory of circular inertia
  • theories of spontaneous generation
  • the contact-action gravitational ether of Fatio and LeSage
  • the optical ether
  • the electromagnetic ether

This is a list of a dozen of cases, but Laudan boldly noted the famous 6 to 1 ratio:

I daresay that for every highly successful theory in the past of science which we now believe to be a genuinely referring theory, one could find half a dozen once successful theories which we now regard as substantially non-referring. (1981: 35)

If we are to take seriously this “plethora” of theories that were both successful and false, it appears that (L) is meant to be a genuinely inductive argument.

There has been a plethora of theories ( ratio 6 to 1 ) which were successful and yet not truthlike.

Therefore, it is highly probable that current theories will not be truthlike (despite their success).

An argument such as (I) has obvious flaws. Two are the most important. The first is that the basis for induction is hard to assess. This does not just concern the 6:1 ratio, of which one may ask: where does it come from? It also concerns the issue of how we individuate and count theories as well as how we judge success and referential failure. Unless we are clear on all these issues in advance of the inductive argument, we cannot even start putting together the inductive evidence for its conclusion (cf. Mizrahi 2013).

The second flaw of (I) is that the conclusion is too strong. It is supposed to be that there is rational warrant for the judgment that current theories are not truthlike. The flaw with this kind of sweeping generalization is precisely that it totally disregards the fresh strong evidence there is for current theories—it renders current evidence totally irrelevant to the issue of their probability of being true. Surely this is unwarranted. Not only because it disregards potentially important differences in the quality and quantity of evidence there is for current theories (differences that would justify treating current theories as more supported by available evidence than past theories were by the then available evidence); but also because it makes a mockery of looking for evidence for scientific theories! If I know that X is more likely than Y and that this relation cannot change by doing Z , there is no point in doing Z .

The second flaw of (I) becomes (even more) apparent when one takes a closer look at the successful-yet-false theories in Laudan’s list. Would anyone be willing to insist that, say, the humoral theory of medicine, the vital force theory of physiology or the theory of crystalline spheres are on a par with our current scientific theories with the same domain of application? The difference between their respective evidence is undoubtedly enormous. Nevertheless, it would rather be mistaken to restrict our attention to those theories comprising Laudan’s own list. Indeed, subsequent scholars have provided new lists of cases where admittedly false theories had been used in the derivation of impressive empirical predictions. Most notably, Timothy Lyons (2002: 70–72) and Peter Vickers (2013: 191–194) suggest the following (partly overlapping) lists:

Lyons’s list

  • Caloric Theory
  • Phlogiston Theory
  • Rankine’s 19 th Century Vortex Theory
  • Newtonian Mechanics
  • Fermat’s Principle of Least Time
  • Fresnel’s Wave Theory of Light and of the Optical Ether
  • Maxwell’s Ether Theory
  • Dalton’s Atomic Theory
  • Kekulé’s Theory of Benzene Molecule
  • Mendeleev’s Periodic Law
  • Bohr’s Theory of the Atom
  • Dirac’s Relativistic Wave Equation
  • The Original (pre-inflationary) Big Bang Theory

Vickers’s list

  • Fresnel’s theory of light and the luminiferous ether
  • Rankine’s vortex theory of thermodynamics
  • Dirac and the positron
  • Teleomechanism and gill slits
  • Reduction division in the formation of sex cells
  • The Titius-Bode law
  • Kepler’s predictions concerning the rotation of the sun
  • Kirchhoff’s theory of diffraction
  • Bohr’s prediction of the spectral lines of ionized helium
  • Sommerfeld’s prediction of the hydrogen fine structure
  • Velikovsky and Venus
  • Steady state cosmology
  • The achromatic telescope
  • The momentum of light
  • S-matrix theory
  • Variation of electron mass with velocity
  • Taking the thermodynamic limit

These lists summarize much of the work done by historians of science and historically informed philosophers of science. They are meant to present cases of empirical successes that were (supposedly) brought about by false theoretical hypotheses, hence offering a fresh source of historical challenges to realism. At first sight, the cases provided look substantially different from the majority of Laudan’s examples (viz. the successes, in at least some of them, are more impressive). Yet, it remains to be seen whether they are more troublesome for scientific realism.

If we think of the pessimistic argument not as inductive but as a warrant-remover argument and if we also think that the fate of (past) theories should have a bearing on what we are warranted in accepting now, we should think of its structure differently. It has been argued by Psillos (1999: chapter 5) that we should think of the pessimistic argument as a kind of reductio. Argument (L) above aimed to “discredit the claim that there is an explanatory connection between empirical success and truth-likeness” which would warrant the realist view that current successful theories are truthlike. If we view the historical challenge this way, viz., as a potential warrant-remover argument, the past record of science does play a role in it, since it is meant to offer this warrant-remover.

Psillos’s (1996) reconstruction of Laudan’s argument was as follows:

Argument (P): (A) Currently successful theories are truthlike. (B) If currently successful theories are truthlike, then past theories are not. (C) These characteristically false past theories were, nonetheless, empirically successful. (The ‘historical gambit’) Hence, empirical success is not connected with truthlikeness and truthlikeness cannot explain success: the realist’s potential warrant for (A) is defeated.

Premise (B) of argument (P) is critical. It is meant to capture radical discontinuity in theory-change, which was put thus (stated in the material mode):

Past theories are deemed not to have been truth-like because the entities they posited are no longer believed to exist and/or because the laws and mechanisms they postulated are not part of our current theoretical description of the world. (Psillos 1999: 97).

In this setting, the ‘historical gambit’ (C) makes perfect sense. Unless there are past successful theories which are warrantedly deemed not to be truthlike, premise (B) cannot be sustained and the warrant-removing reductio of (A) fails. If (C) can be substantiated, success cannot be used to warrant the claim that current theories are true. The realists’ explanatory link between truthlikeness and empirical success is undercut. (C) can be substantiated only by examining past successful theories and their fate. History of science is thereby essentially engaged.

The realist response has come to be known as the divide et impera strategy to refute the pessimistic argument. The focus of this strategy was on rebutting the claim that the truth of current theories implies that past theories cannot be deemed truthlike. To defend realism, realists needed to be selective in their commitments. This selectivity was developed by Kitcher (1993) and (independently) by Psillos (1994).

One way to be selective is to draw a distinction between working posits of a theory (viz., those theoretical posits that occur substantially in the explanatory schemata of the theory) and presuppositional posits (putative entities that apparently have to exist if the instances of the explanatory schemata of the theory are to be true) (cf. Kitcher 1993: 149). Another way is to draw a distinction between the theoretical claims that essentially or ineliminably contribute to the generation of successes of a theory and those claims that are ‘idle’ components that have had no contribution to the theory’s success (cf. Psillos 1994, 1996). The underlying thought is that the empirical successes of a theory do not indiscriminably support all theoretical claims of the theory, but rather the empirical support is differentially distributed among the various claims of the theory according to the contribution they make to the generation of the successes. Generally, Kitcher (1993) and Psillos (1996, 1999) have argued that there are ways to distinguish between the ‘good’ and the ‘bad’ parts of past abandoned theories and to show that the ‘good’ parts—those that enjoyed evidential support, were not idle components and the like—were retained in subsequent theories.

It is worth-noting that, methodologically, the divide et impera strategy recommended that the historical challenge to realism can only be met by looking at the actual successes of past successful theories and by showing that those parts of past theories (e.g., the caloric theory of heat or the optical ether theories) that were fuelling theory successes were retained in subsequent theories and those theoretical terms which were central in the relevant past theories were referential. In fact, Vickers has recently made the methodological suggestion that if one’s sole aim is to cope with a historical challenge, then it is sufficient that one shows that the abandoned hypotheses were not essential for the relevant theory’s empirical success, without at the same time taking side on which are the essential theoretical hypotheses. As Vickers claims, in order to respond to a PI-style challenge, “all the realist needs to do is show that the specific assumptions identified by the antirealist do not merit realist commitment. And she can do this without saying anything about how to identify the posits which do merit realist commitment” (2017: 3224). Besides, according to Vickers’s conception of the dialectic of the PI-debate, the onus of proof lies with the antirealist: the antirealist has to reconstruct the derivation of a prediction, identify the assumptions that merit realist commitments and then show that at least one of them is not truthlike by our current lights; and then all the realists need to show is that the specific assumptions were inessential. In sum, Vickers argues that “the project of responding to the historical challenge” and “the project of explaining what realists should commit to” have to be kept distinct (2017: 3222).

At any rate, either employed in the identification of the trustworthy theoretical parts or in the (mere) handling of a historical challenge, the divide et impera move suggests that there has been enough theoretical continuity in theory-change to warrant the realist claim that science is ‘on the right track.

The realist move from substantive continuity in theory-change to truthlikeness has been challenged on grounds that there is no entitlement to move from whatever preservation in theoretical constituents there is in theory-change to these constituents’ being truthlike (Chang 2003: 910–12; Stanford 2006). Against this point it has been argued that the realist strategy proceeds in two steps (cf. Psillos 2009: 72). The first is to make the claim of continuity (or convergence) plausible, viz., to show that there is continuity in theory-change: substantive theoretical claims that featured in past theories and played a key role in their successes (especially novel predictions) have been incorporated in subsequent theories and continue to play an important role in making them empirically successful. But this first step does not establish that the convergence is to the truth . For this claim to be made plausible a second argument is needed, viz., that the emergence of this evolving-but-convergent network of theoretical assertions is best explained by the assumption that it is, by and large, truthlike. So there is, after all, entitlement to move from convergence to truthlikeness, insofar as truthlikeness is the best explanation of this convergence.

Another critical point was that the divide et impera strategy cannot offer independent support to realism since it is tailor-made to suit realism: it is the fact that the very same present theory is used both to identify which parts of past theories were empirically successful and which parts were (approximately) true that accounts for the realists’ wrong impression that these parts coincide (Stanford 2006). He says:

With this strategy of analysis, an impressive retrospective convergence between our judgements of the sources of a past theory’s success and the things it ‘got right’ about the world is virtually guaranteed: it is the very fact that some features of a past theory survive in our present account of nature that leads the realist both to regard them as true and to believe that they were the sources of the rejected theory’s success or effectiveness. So the apparent convergence of truth and the sources of success in past theories is easily explained by the simple fact that both kinds of retrospective judgements have a common source in our present beliefs about nature. (2006: 166)

It has been claimed by Psillos (2009) that the foregoing objection is misguided. The problem is this. There are the theories scientists currently endorse and there are the theories that had been endorsed in the past. Some (but not all) of them were empirically successful (perhaps for long periods of time). They were empirically successful irrespective of the fact that, subsequently, they came to be replaced by others. This replacement was a contingent matter that had to do with the fact that the world did not fully co-operate with the then extant theories: some of their predictions failed; or the theories became overly ad hoc or complicated in their attempt to accommodate anomalies, or what have you. The replacement of theories by others does not cancel out the fact that the replaced theories were empirically successful. Even if scientists had somehow failed to come up with new theories, the old theories would not have ceased to be successful. So success is one thing, replacement is another.

Hence, it is one thing to inquire into what features of some past theories accounted for their success and quite another to ask whether these features were such that they were retained in subsequent theories of the same domain. These are two independent issues and they can be dealt with (both conceptually and historically) independently. One should start with some past theories and—bracketing the question of their replacement—try to identify, on independent grounds, the sources of their empirical success; that is, to identify those theoretical constituents of the theories that fuelled their successes. When a past theory has been, as it were, anatomised, we can then ask the independent question of whether there is any sense in which the sources of success of a past theory that the anatomy has identified are present in our current theories. It’s not, then, the case that the current theory is the common source for the identification of the successful parts of a past theory and of its truthlike parts.

The transition from Newton’s theory of gravity to Einstein’s illustrates this point. Einstein took it for granted that Newton’s theory of gravity (aided by perturbation theory) could account for 531 arc-second per century of the perturbation of Mercury’s perihelion. Not only were the empirical successes of Newton’s theory identified independently of the successor theory, but also some key theoretical components of Newton’s theory—the law of attraction and the claim that the gravitational effects from the planets on each other were a significant cause of the deviations from their predicted orbits—were taken to be broadly correct and explanatory (of at least part) of the successes. Einstein could clearly identify the sources of successes of Newton’s theory independently of his own alternative theory and it is precisely for this reason that he insisted that he had to recover Newton’s law of attraction (a key source of the Newtonian success) as a limiting case of his own GTR. He could then show that his new theory could do both: it could recover the (independently identified) sources of successes of Newton’s theory (in the form of the law of attraction) and account for its failures by identifying further causal factors (the curvature of space-time) that account for the discrepancies between the predicted orbits of planets (by Newton’s theory of gravity) and the observed trajectories. [ 6 ]

Apart from Stanford’s case against the divide et impera move, the latter has become the target of criticism—among others—by Timothy Lyons. [ 7 ] Lyons (2006) focuses his critique on Psillos’s criterion for the conditions under which a hypothesis indispensably contributes to the derivation of novel predictions. In his (1999: 100) Psillos says:

Suppose that \(H\) together with another set of hypotheses \(H'\) (and some auxiliaries A ) entail a prediction \(P\). \(H\) indispensably contributes to the generation of \(P\) if \(H'\) and A alone cannot yield \(P\) and no other available hypothesis \(H^*\) which is consistent with \(H'\) and A can replace \(H\) without loss in the relevant derivation of \(P\).

Lyons interprets this passage—as well as Psillos’s subsequent claim that \(H^*\) must satisfy some “natural epistemic constraints”, such as being “independently motivated, non ad hoc, potentially explanatory etc.” (ibid.)—as providing the following criterion for the essential role of hypothesis \(H\) in the derivation of prediction \(P\):

For \(H\) to be essential [for the derivation of \(P\)]:

Thus construed, Lyons criticizes Psillos’ criterion for essentiality, as being “superfluous, unmotivated, and therefore inappropriate” (2006: 541). Briefly put, his point is that condition 3 “unacceptably overshoots” the realist’s goal, since the absence of an alternative \(H^*\) has “no bearing whatsoever on whether \(H\) itself contributed to, was deployed in, the derivation of a given prediction” (2006: 540). Besides, Lyons states that condition 3 is so vague that it is “simply inapplicable” (2006: 542). According to Mario Alai’s (2021) summary of Lyons’s point, condition 3 doesn’t specify ( a ) when the alternative hypothesis \(H^*\) must or must not be available, ( b ) what ‘potentially explanatory etc.’ means and ( c ) whether \(H'\) and \(A\) must be essential too. In addition; ( d ) it doesn’t state whether \(H^*\) is allowed to lead to losses of other confirmed predictions and ( e ) whether \(H^*\) should be consistent with those elements of \(H'\) and \(A\), which, though they are ‘essential’ for other predictions, they are dispensable when it comes to the derivation of the prediction under scrutiny. Based on these points, Lyons suggests that even if realists might hold onto conditions 1 and 2 above, condition 3 has to be abandoned, thereby isolating “the deployment realist’s fundamental insight”, viz., that credit should be attributed to those posits that actually —as opposed to essentially —have been deployed in the derivation of empirical predictions (2006,543).

In reply to Lyons, Peter Vickers (2017) and Alai (2021) have defended the divide et impera move against the PI by suggesting the following refinement of condition 3 (let’s call it 3′):

According to Vickers, when realists are presented with an instance of a (seemingly) success-inducing-yet-false hypothesis, all they need to do is to show that the specific hypothesis does not satisfy the above condition. It should be noted, however, that this, in essence, is the strategy recommended by Psillos in his 1994, where he aimed to show, using specific cases, that various assumptions such as that heat is a material substance in the case of the caloric theory of heat, do not merit realist commitment, because there are weaker assumptions that fuel the derivation of successful predictions.

Alai claims that substituting condition 3′ for condition 3 is an improvement of the divide et impera move, for not only does condition 3′ perform the task that Psillos had in mind, but also escapes from Lyons’ criticisms (2021: 188). To begin with, condition 3′ is said not to suffer from the (alleged) vagueness of condition 3, for according to Alai: ( a ) there is no question about when the alternative \(H^*\) is available ; ( b ) there is no need to specify what ‘explanatory’ means; and ( c ) it is not required that \(H'\) and \(A\) are also essential. In addition, ( d ) condition 3′ allows that \(H^*\) may lead to losses of other confirmed predictions and ( e ), since 3′ excludes only hypotheses \(H^*\) which are entailed by \(H\), \(H^*\) are ipso facto consistent with \(H'\) and \(A\).

Now, it is rather evident that condition 3′ is neither superfluous nor unmotivated, since as Alai (2021, 188) stressed it is motivated by a plausible epistemic principle associated with the Occam’s razor:

in abductions we can assume only what is essential, i.e., the weakest hypothesis sufficient to explain a given effect; but if a hypotheses [sic], although deployed, was not essential in deriving [the novel prediction at hand], it is not essential in explaining its derivation either; therefore deployment realists need not (and must not) be committed to its truth.

In sum, contra Lyons, condition 3΄ is both epistemologically motivated as well as indispensable for the proper application of the divide et impera move.

The ‘Vickers-Alai’ refinement of the divide et impera move has not been uncontested. It has been criticized on principled grounds, as well as for not being sufficient in dealing with PI-style challenges. For instance, Dean Peters (2014) argues inter alia that Vickers’ criterion for essentiality cannot account for the unificatory aspect of scientific theorizing, whereas Florian Boge (2021) and Dana Tulodziecki (2021) have provided new historical counterexamples—within the field of nuclear physics and phychometry, and the 19 th century miasma theory of disease, respectively—that cannot be handled, or so it is argued, by the ‘Vickers-Alai’ criterion.

It should also be noted that, according to Vickers himself, the employment of condition 3′ in dealing with PI seems to bring scientific realism dangerously close to structural realism. As has already been said, Vickers’s recipe for handling a PI-style challenge is roughly the following: take the (false) hypothesis \(H\) that, according to the anti-realist, is employed in the derivation of a prediction \(P\), identify an (uncontested) \(H^*\) which is entailed by \(H\) and show that \(H^*\) is enough for the derivation of \(P\). This recipe goes a long way in disarming Lyon’s objection. And yet, Vickers notes, an even weaker hypothesis \(H^{**}\) is available, viz., that for the prediction of P only the mathematical structure of \(H^*\) is required. But then, “only the very abstract ‘structure’ truly merits realist commitment, as structural realists like to claim” (2017: 3227). If we take ‘structure’ to be identified with the Ramsey sentence of a given theory (see the next section), then Vickers’ concern is, at least prima facie , a plausible one. For the Ramsey sentence of a theory is obviously entailed by the latter and, as is well known, any theory and its Ramsey sentence have exactly the same observational consequences. Hence, it seems that the employment of condition 3′ forces realists to restrict their commitment solely towards the Ramsey sentences of their favoured theories. In reply, however, it should be stressed that though Vickers’s concern is prima facie warranted, it is far from conclusive. In fact, after raising his concern, Vickers doesn’t further explore it, whereas Alai (2021: 211–212) has argued that from the mere application of condition 3′ “it doesn’t follow that every hypothesis is dispensable in favor of its Ramsey sentence”.

An instance of the divide et impera strategy is structural realism. This view has been associated with John Worrall (1989), who revived the relationist account of theory-change that emerged in the beginning of the twentieth century. In opposition to scientific realism, structural realism restricts the cognitive content of scientific theories to their mathematical structure together with their empirical consequences. But, in opposition to instrumentalism, structural realism suggests that the mathematical structure of a theory represents the structure of the world (real relations between things). Against PI, structural realism contends that there is continuity in theory-change, but this continuity is (again) at the level of mathematical structure. Hence, the ‘carried over’ mathematical structure of the theory correctly represents the structure of the world and this best explains the predictive success of a theory. [ 8 ]

Structural realism was independently developed in the 1970s by Grover Maxwell (1970a, 1970b) in an attempt to show that the Ramsey-sentence approach to theories need not lead to instrumentalism. Ramsey-sentences go back to a seminal idea by Frank Ramsey (1929). To get the Ramsey-sentence \(^{R}T\) of a (finitely axiomatisable) theory T we conjoin the axioms of T in a single sentence, replace all theoretical predicates with distinct variables \(u_i\), and bind these variables by placing an equal number of existential quantifiers \(\exists u_i\) in front of the resulting formula. Suppose that the theory T is represented as T (\(t_1\),…, \(t_n\); \(o_1\),…, \(o_m\)), where T is a purely logical \(m+n\)-predicate. The Ramsey-sentence \(^{R}T\) of T is:

The Ramsey-sentence \(^{R}T\) that replaces theory T has exactly the same observational consequences as T ; it can play the same role as T in reasoning; it is truth-evaluable if there are entities that satisfy it; but since it dispenses altogether with theoretical vocabulary and refers to whatever entities satisfy it only by means of quantifiers, it was taken to remove the issue of the reference of theoretical terms/predicates. ‘Structural realism’ was suggested to be the view that: i) scientific theories issue in existential commitments to unobservable entities and ii) all non-observational knowledge of unobservables is structural knowledge , i.e., knowledge not of their first-order (or intrinsic) properties, but rather of their higher-order (or structural) properties. The key idea here was that a Ramsey-sentence satisfies both conditions (i) and (ii). So we might say that, if true, the Ramsey-sentence \(^{R}T\) gives us knowledge of the structure of the world: there is a certain structure which satisfies the Ramsey-sentence and the structure of the world (or of the relevant worldly domain) is isomorphic to this structure.

Though initially Worrall’s version of structural realism was different from Maxwell’s, being focused on—and motivated by— Poincaré’s argument for structural continuity in theory-change, in later work Worrall came to adopt the Ramsey-sentence version of structural realism (see appendix IV of Zahar 2001).

A key problem with Ramsey-sentence realism is that though a Ramsey-sentence of a theory may be empirically inadequate, and hence false, if it is empirically adequate (if, that is, the structure of observable phenomena is embedded in one of its models), then it is bound to be true. For, as Max Newman (1928) first noted in relation to Russell’s (1927) structuralism, given some cardinality constraints, it is guaranteed that there is an interpretation of the variables of \(^{R}T\) in the theory’s intended domain. [ 9 ]

More recently, David Papineau (2010) has argued that if we identify the theory with its Ramsey-sentence, it can be argued that past theories are approximately true if there are entities which satisfy, or nearly satisfy, their Ramsey-sentences. The advantage of this move, according to Papineau, is that the issue of referential failure is bypassed when assessing theories for approximate truth, since the Ramsey sentence replaces the theoretical terms with existentially bound variables. But as Papineau (2010: 381) admits, the force of the historical challenge to realism is not thereby thwarted. For it may well be the case that the Ramsey-sentences of most past theories are not satisfied (not even nearly so). [ 10 ]

In the more recent literature, there has been considerable debate as to how exactly we should understand PI. There are those, like Anjan Chakravartty who take it that PI is an Induction. He says:

PI can … be described as a two-step worry. First, there is an assertion to the effect that the history of science contains an impressive graveyard of theories that were previously believed [to be true], but subsequently judged to be false … Second, there is an induction on the basis of this assertion, whose conclusion is that current theories are likely future occupants of the same graveyard. (2008: 152)

Yet, it is plausible to think that qua an inductive argument, history-based pessimism is bound to fail. The key point here is that the sampling of theories which constitute the inductive evidence is neither random nor otherwise representative of theories in general.

It has been argued that, seen as an inductive argument, PI is fallacious: it commits the base-rate fallacy (cf. Lewis 2001). If in the past there have been many more false theories than true ones, (if, in other words, truth has been rare), it cannot be concluded that there is no connection between success and truth. Take S to stand for Success and not- S to stand for failure. Analogously, take T to stand for truth of theory T and not- T for falsity of theory T . Assume also that the probability that a theory is unsuccessful given that it is true is zero \((\textrm{Prob}({\textrm{not-}S}\mid T)=0)\) and that the probability that a theory is successful given that it is false is 0.05 \((\textrm{Prob}(S\mid {\textrm{not-}T})=0.05)\). Assume that is, that there is a very high True Positives (successful but true) rate and a small False Positives (successful but false theories) rate. We may then ask the question: How likely is it that a theory is true, given that it is successful? That is, what is the posterior probability \(\textrm{Prob}(T\mid S)\)?

This answer is indeterminate if we don’t take into account the base-rate of truth, viz., the incidence rate of truth in the population of theories. If the base rate is very low (let’s assume that only 1 in 50 theories have been true), then it is unlikely that T is true given success. \(\textrm{Prob}(T\mid S)\) would be around 0.3. But this does not imply something about the connection between success and truth. It is still the case that the false positives are low and that the true positives high. The low probability is due to the fact that truth is rare (or that falsity is much more frequent). For \(\textrm{Prob}(T\mid S)\) to be high, it must be the case that \(\textrm{Prob}(T)\) is not too small. But if \(\textrm{Prob}(T)\) is low, it can dominate over a high likelihood of true positives and lead to a very low posterior probability \(\textrm{Prob}(T\mid S)\). Similarly, the probability that a theory is false given that it is successful (i.e., \(\textrm{Prob}({\textrm{not-}T}\mid S))\) may be high simply because there are a lot more false theories than true ones. As Peter Lewis put it:

At a given time in the past, it may well be that false theories vastly outnumber true theories. In that case, even if only a small proportion of false theories are successful, and even if a large proportion of true theories are successful, the successful false theories may outnumber the successful true theories. So the fact that successful false theories outnumber successful true theories at some time does nothing to undermine the reliability of success as a test for truth at that time, let alone at other times (2001: 376–7).

Seen in this light, PI does not discredit the reliability of success as a test for truth of a theory; it merely points to the fact that truth is scarce among past theories. [ 11 ]

Challenging the inductive credentials of PI has acquired a life of its own. A standard objection (cf. Mizrahi 2013) is that theories are not uniform enough to allow an inductive generalization of the form “seen one, seen them all”. That is, theories are diverse enough over time, structure and content not to allow us to take a few of them—not picked randomly—as representative of all and to project the characteristics shared by those picked to all theories in general. In particular, the list that Laudan produced is not a random sample of theories. They are all before the twentieth century and all have been chosen solely on the basis that they had had some successes (irrespective of how robust these successes were). An argument of the form:

X % of past successful theories are false

Therefore, X % of all successful theories are false

would be a weak inductive argument because

it fails to provide grounds for projecting the property of the observed members of the reference class to unobserved members of the reference class. (Mizrahi 2013: 3219)

Things would be different, if we had a random sampling of theories. Mizrahi (2013: 3221–3222) collected 124 instances of ‘theory’ from various sources and picked at random 40 of them. These 40 were then divided into three groups: accepted theories, abandoned theories and debated theories. Of those 40 theories, 15% were abandoned and 12% debated. Mizrahi then notes that these randomly selected data cannot justify an inductively drawn conclusion that most successful theories are false. On the contrary, an optimistic induction would be more warranted:

72% of sampled theories are accepted theories (i.e., considered true).

Therefore, 72% of all theories are accepted theories (i.e., considered true).

Mizrahi has come back to the issue of random sampling and has attempted to show that the empirical evidence is against PI:

If the history of science were a graveyard of dead theories and abandoned posits, then random samples of scientific theories and theoretical posits would contain significantly more dead theories and abandoned posits than live theories and accepted posits. It is not the case that random samples of scientific theories and theoretical posits contain significantly more dead theories and abandoned posits than live theories and accepted posits. Therefore, It is not the case that the history of science is a graveyard of dead theories and abandoned posits. (2016: 267)

A similar argument has been defended by Park (2011). We may call it, the explosion argument: Most key theoretical terms of successful theories of the twentieth century refer “in the light of current theories”. But then, “most central terms of successful past theories refer”, the reason being that there are far more twentieth century theories than theories in total. This is because “the body of scientific knowledge exploded in the twentieth century with far more human and technological resources” (2011: 79).

Let’s call this broad way to challenge the inductive credentials of the pessimistic argument ‘the Privilege-for-current-theories strategy’. This has been adopted by Michael Devitt (2007) too, though restricted to entities. Devitt, who takes realism to be a position concerning the existence of unobservables, noted that the right question to ask is this: ‘What is the “success ratio” of past theories?’, where the “success ratio” is “the ratio of the determinately existents to the determinately nonexistents + indeterminates”. Asserting a privilege for current science, he claims that “we are now much better at finding out about unobservables”. According to him, then, it is “fairly indubitable” that the historical record shows “improvement over time in our success ratio for unobservables’.

In a similar fashion but focusing on current theories, Doppelt (2007) claims that realists should confine their commitment to the approximate truth of current best theories, where best theories are those that are both most successful and well established. The asymmetry between current best theories and past ones is such that the success of current theories is of a different kind than the success of past theories. The difference, Doppelt assumes, is so big that the success of current theories can only be explained by assuming that they are approximately true, whereas the explanation of the success of past theories does not require this commitment.

If this is right, there is sufficient qualitative distance between past theories and current best ones to block

any pessimistic induction from the successful-but-false superseded theories to the likelihood that our most successful and well-established current theories are also probably false. (Doppelt 2007: 110).

The key difference, Doppelt argues, is that

our best current theories enjoy a singular degree of empirical confirmation impossible for their predecessors, given their ignorance of so many kinds of phenomena and dimensions of nature discovered by our best current theories.

This singular degree of empirical confirmation amounts to raising the standards of empirical success to a level unreachable by past theories (cf. 2007: 112).

The advocate of PI can argue that past ‘best theories’ also raised and met the standards of empirical success, which inductively supports the conclusion that current best theories will be superseded by others which will meet even higher standards of success. Doppelt’s reply is that this new version of PI “should not be given a free pass as though it were on a par with the original pessimistic induction” the reason being that “in the history of the sciences, there is greater continuity in standards of empirical success than in the theories taken to realize them”. Hence, the standards of empirical success change slower than theories. Hence, it is not very likely that current standards of empirical success will change any time soon.

It has been argued, however, that Doppelt cannot explain the novel predictive success of past theories without arguing that they had truthlike constituents (cf. Alai 2017). Besides, as Alai puts it, “current best theories explain the (empirical) success of discarded ones only to the extent that they show that the latter were partly true” (2017: 3282).

The ‘Privilege-for-current-theories strategy’ has been supported by Ludwig Fahrbach (2011). The key point of this strategy is that the history of science does not offer a representative sample of the totality of theories that should be used to feed the historical pessimism of PI. In order to substantiate this, Fahrbach suggested, based on extensive bibliometric data, that over the last three centuries the number of papers published by scientists as well as the number of scientists themselves have grown exponentially , with a doubling rate of 15–20 years. Hence, he claims, the past theories that feed the historical premise of PI were produced during the time of the first 5% of all scientific work ever done by scientists. As such the sample is totally unrepresentative of theories in total; and hence the pessimistic conclusion, viz., that current theories are likely to be false and abandoned in due course, is inductively unwarranted. Moreover, Fahrbach argues, the vast majority of theories enunciated in the last 50–80 years, (which constitute the vast majority of scientific work ever produced) are still with us. Hence, as he puts it,

(t)he anti-realist will have a hard time finding even one or two convincing examples of similarly successful theories that were accepted in the last 50–80 years for some time, but later abandoned. (2011: 152)

Since there have been practically no changes “among our best (i.e., most successful) theories”, Fahrbach suggests

an optimistic meta-induction to the effect that they will remain stable in the future, i.e., all their empirical consequences which scientists will ever have occasion to compare with results from observation at any time in the future are true. (2011: 153)

The conclusion is that the PI is unsound: “its conclusion that many of our current best scientific theories will fail empirically in the future cannot be drawn” (2011: 153).

A key assumption of the foregoing argument is that there is a strong connection between the amount of scientific work (as measured by the number of journal articles) and the degree of success of the best scientific theories. But this can be contested on the grounds that it’s a lot easier to publish currently than it was in the seventeenth century and that current research is more tightly connected to the defense of a single theoretical paradigm than before. This might well be a sign of maturity of current science but, as it stands, it does not show that the present theoretical paradigm is not subject to radical change. Florian Müller (2015) put the point in terms of decreasing marginal revenues. The correlation between increased scientific work and scientific progress, which is assumed by Fahrbach may not be strong enough:

It seems more plausible to expect decreasing marginal revenues of scientific work since it usually takes much less time to establish very basic results than to make progress in a very advanced state of science. (Müller 2015: 404)

The ‘Privilege-for-current-theories strategy’ can be further challenged on the grounds that it requires some “fundamental difference between the theories we currently accept, and the once successful theories we have since rejected” (Wray 2013: 4325). As Brad Wray (2013) has argued Fahrbach’s strategy is doomed to fail because the argument from exponential growth could be repeated at former periods too, thereby undermining itself. Imagine that we are back in 1950 and we look at the period between 1890 and 1950. We could then argue, along Farhbach’s lines, that the pre-1890 theories (which were false and abandoned) were an unrepresentative sample of all theories and that the recent theories (1890–1950) are by far the most theories until 1950 and that, since most of them have not been abandoned (by 1950), they are likely to remain impervious to theory-change. Or imagine that we are further back in 1890 and look at the theories of the period 1830–1890. We could run the same argument about those theories, viz, that they are likely to survive theory change. But if we look at the historical pattern , they did not survive; nor did the theories between 1890–1950. By the same token, we should not expect current theories to survive theory-change.

Is there room for defending an epistemic privilege for current science? Two points are worth making. The first is that it’s hard to defend some kind of epistemic privilege of current science if the realist argument against PI stays only at a level of statistics (even assuming that there can be statistics over theories). If there is an epistemic privilege of current science in relation to past science, it is not a matter of quantity but of quality . The issue is not specifying how likely it is that an arbitrary current theory T be true, given the evidence of the past record of science. The issue, instead, is how a specific scientific theory—a real theory that describes and explains certain well-founded worldly phenomena—is supported by the evidence there is for it. If we look at the matter from this perspective, we should look at case-histories and not at the history of science at large. The evidence there is for specific theory T (e.g., the Darwinian synthesis or GTR etc.) need not be affected by past failures in the theoretical understanding of the world in general. The reason is that there is local epistemic privilege, that is, privilege over past relevant theories concerning first-order evidence and specific methods.

The second point is this. Wray’s argument against Fahrbach is, in effect, that there can be a temporal meta-(meta-)induction which undermines at each time t (or period Dt ) the privilege that scientific theories at t or Dt are supposed to have. So Wray’s point is this: at each time \(t_{i}\) (or period \(Dt_{i}\)), scientists claim that their theories are not subject to radical change at subsequent times; but if we look at the pattern of theory change over time, the history of science shows that there have been subsequent times \(t_{i}+1\) (or periods \({Dt}_{i}+1\)) such that the theories accepted at \(t_{i}\) were considered false and abandoned. Hence, he takes it that at no time \(t_{i}\) are scientists justified in accepting their theories as not being subject to radical change in the future. But this kind of argument is open to the following criticism. It assumes, as it were, unit-homogeneity, viz., that science at all times \(t_{i}\) (and all periods \({Dt}_{i}\)) is the same when it comes to how far it is from the truth. Only on this assumption can it be argued that at no time can scientists claim that their theories are not subject to radical change. For if there are senses in which subsequent theories are closer to the truth than their predecessors, it is not equally likely that they will be overturned as their predecessors were.

The point, then, is that though at each time \(t_{i}\) (or period \({Dt}_{i}\)) scientists might well claim that their theories are not subject to radical change at subsequent times, they are not equally justified in making this claim! There might well be times \(t_{i}\) (or periods \({Dt}_{i}\)) in which scientists are more justified in making the claims that their theories are not subject to radical change at subsequent times simply because they have reasons to believe that their theories are truer than their predecessors. To give an example: if Wray’s argument is right then Einstein’s GTR is as likely to be overthrown at future time \(t_{2100}\) as was Aristotle’s crystalline spheres theory in past time \(t_{-300}\). But this is odd. It totally ignores the fact that all available evidence renders GTR closer to the truth than the simply false Aristotelian theory. In other words, that GTR has substantial truth-content makes it less likely to be radically revised in the future.

An analogous point was made by Park (2016). He defined what he called Proportional Pessimism as the view that “as theories are discarded, the inductive rationale for concluding that the next theories will be discarded grows stronger” (2016: 835). This view entails that the more theories have been discarded before T is discarded, the more justified we are in thinking that T is likely to be discarded. However, it is also the case that based on their greater success, we are more justified to take newer theories to be more likely to be truthlike than older ones. We then reach a paradoxical situation: we are justified to take newer theories to be both more probable than older ones and more likely to be abandoned than older ones.

If an inductive rendering of historical pessimism fails, would a deductive rendering fare better? Could PI be considered at least as a valid deductive argument ? Wray (2015: 65) interprets the original argument by Laudan as being deductive. And he notes

as far as Laudan is concerned, a single successful theory that is false would falsify the realist claim that (all) successful theories are true; and a single successful theory that refers to a non-existent type of entity would falsify the realist claim that (all) successful theories have genuinely referring theoretical terms.

But if this is the intent of the argument, history plays no role in it. All that is needed is a single counterexample, past or present. This, it should be noted, is an endemic problem with all attempts to render PI as a deductive argument. Müller, for instance, notes that the fundamental problem raised by PI is “simply that successful theories can be false”. He adds:

Even just one counterexample (as long as it is not explained away) undermines the claim that truth is the best explanation for the success of theories as it calls into question the explanatory connection in general. (2015: 399)

Thus put, the history of past failures plays no role in PI. Any counterexample, even one concerning a current theory, will do.

How is it best to understand the realist theses that the history of science is supposed to undermine? Mizrahi (2013: 3224) notes that the realist claim is not meant to be a universal statement. As he puts it:

Success may be a reliable indicator of (approximate) truth, but this is compatible with some instances of successful theories that turn out not to be approximately true. In other words, that a theory is successful is a reason to believe that it is approximately true, but it is not a conclusive proof that the theory is approximately true.

The relation between success and (approximate) truth, in this sense, is more like the relation between flying and being a bird: flying characterizes birds even if kiwis do not fly. If this is so, then there is need for more than one counter-example for the realist thesis to be undermined.

A recent attempt to render PI as a deductive argument is by Timothy Lyons. He (2016b) takes realism to issue in the following meta-hypothesis : “our successful scientific theories are (approximately) true”. He then reconstructs PI thus:

This is supposed to be a deductive argument against the ‘meta hypothesis’. But in his argument the history of science plays no role. All that is needed for the argument above to be sound is a single instance of a successful theory that is not true. A single non-white swan is enough to falsify the hypothesis ‘All swans are white’—there is no point in arguing here: the more, the merrier! In a similar fashion, it doesn’t add much to argument (D) to claim

the quest to empirically increase the quantity of instances (…) is rather to secure the soundness of the modus tollens , to secure the truth of the pivotal second premise, the claim that there are counterinstances to the realist meta-hypothesis. (Lyons 2016b: 566)

In any case, a critical question is: can some false-but-rigorously-empirically-successful theories justifiably be deemed truthlike from the point of view of successor theories? This question is hard to answer without looking at actual cases in the history of science. The general point, made by Vickers (2017) is that it is not enough for the challenger of realism to identify some components of past theories which were contributing to their successes such that they were not retained in subsequent theories. The challenger of realism should show that false components “merit realist commitment”. If they do not, “ (…) that is enough to answer the historical challenge”.

More generally, the search for a generic form of the pessimistic X -duction (In-duction or De-duction) has yielded the following problem: If the argument is inductive, it is at best weak. If the argument is deductive, even if it is taken to be sound, it makes the role of the history of science irrelevant. [ 12 ]

Stanford (2006) has aimed to replace PI with what he calls the ‘new induction’ on the history of science, according to which past historical evidence of transient underdetermination of theories by evidence makes it likely that current theories will be supplanted by hitherto unknown (unconceived) ones, which nonetheless, are such that when they are formulated, they will be at least as well confirmed by the evidence as the current ones. But the new induction is effective, if at all, only in tandem with PI. For if there is continuity in our scientific image of the world, the hitherto unconceived theories that will replace the current ones won’t be the radical rivals they are portrayed to be. [ 13 ]

When it comes to the realist commitment to theories, the proper philosophical task is to ignore neither the first order scientific evidence that there is for a given theory nor the lessons that can be learned from the history of science. Rather, the task is to balance the first-order and the second order of evidence. The first-order evidence is typically associated with whatever scientists take into account when they form an epistemic attitude towards a theory. It can be broadly understood to include some of the theoretical virtues of the theory at hand—of the kind that typically go into plausibility judgments associated with assignment of prior probability to theories. The second-order evidence comes from the past record of scientific theories and/or from meta-theoretical (philosophical) considerations that have to do with the reliability of scientific methodology. It concerns not particular scientific theories, but science as a whole. This second-order evidence feeds claims such as those that motivate PI or the New Induction. Actually, this second-order evidence is multi-faceted—it is negative (showing limitations and shortcomings) as well as positive (showing how learning from experience can be improved).

  • Ainsworth, Peter M., 2009, “Newman’s Objection”, The British Journal for the Philosophy of Science , 60(1): 135–171. doi:10.1093/bjps/axn051
  • Alai, Mario, 2017, “Resisting the Historical Objections to Realism: Is Doppelt’s a Viable Solution?” Synthese , 194(9): 3267–3290. doi:10.1007/s11229-016-1087-z
  • –––, 2021, “The Historical Challenge to Realism and Essential Deployment”, in T.D. Lyons and P. Vickers (eds.), Contemporary Scientific Realism: The Challenge from the History of Science , Oxford: Oxford University Press, pp. 183–215.
  • Anonymous, 1889, “Causerie Bibliographique”, Revue Scientifique , No. 7, August 17 1889. [ Anonymous 1889: 215 available online ]
  • Berthelot, Marcelin, 1897, Science et Morale , Paris: Calmann Levy. [ Berthelot 1897 available online ]
  • Boge, Florian, J., 2021, “Incompatibility and the Pessimistic Induction: A Challenge for Selective Realism”, European Journal for Philosophy of Science , 11(2): 1–31. doi:10.1007/s13194-021-00367-4
  • Boltzmann, Ludwig, 1901, “The Recent Development of Method in Theoretical Physics”, The Monist , 11(2): 226–257. doi:10.5840/monist190111224
  • Brunetière, Ferdinand, 1889, “Revue Littéraire—A propos du Disciple de Paul Bourget”, Revue des Deux Mondes , 94: 214–226. [ Brunetière 1889 available available online ]
  • –––, 1895, “Après une visite au Vatican”, Revue des Deux Mondes , 127: 97–118. [ Brunetière 1895 available online ]
  • Chakravartty, Anjan, 2008, “What you don’t Know can’t Hurt you: Realism and the Unconceived”, Philosophical Studies , 137: 149–158. doi:10.1007/s11098-007-9173-1
  • Chang, Hasok, 2003, “Preservative Realism and its Discontents: Revisiting Caloric”, Philosophy of Science , 70(5): 902–912. doi:10.1086/377376
  • Cordero, Alberto, 2011, “Scientific Realism and the Divide et Impera Strategy: The Ether Saga Revisited”, Philosophy of Science , 78(5): 1120–1130. doi:10.1086/662566
  • Cruse, Pierre, 2005, “Ramsey-sentences, Structural Realism and Trivial Realisation”, Studies in History and Philosophy of Science , 36(3): 557–576. doi:10.1016/j.shpsa.2005.07.006
  • Cruse, Pierre & David Papineau, 2002, “Scientific Realism Without Reference”, in Michele Marsonet (ed.) The Problem of Realism , Aldershot: Ashgate.
  • Demopoulos, William, 2003, “On the Rational Reconstruction of Our Theoretical Knowledge”, The British Journal for the Philosophy of Science , 54(3): 371–403. doi:10.1093/bjps/54.3.371
  • Devitt, Michael, 2007, “Scientific Realism”, In Frank Jackson and Michael Smith, (eds) The Oxford Handbook of Contemporary Philosophy , Oxford: Oxford University Press.
  • –––, 2011, “Are Unconceived Alternatives a Problem for Scientific Realism?” Journal for General Philosophy of Science , 42(2): 285–293. doi:10.1007/s10838-011-9166-9
  • Doppelt, Gerald, 2007, “Reconstructing Scientific Realism to Rebut the Pessimistic Meta-induction”, Philosophy of Science , 74(1): 96–118. doi:10.1086/520685
  • Duhem, Pierre Maurice Marie, 1906 [1954], Théorie physique: son objet et sa structure , Paris. Translated from the 1914 second edition as The Aim and Structure of Physical Theory , Philip P. Wiener (trans.), Princeton, NJ: Princeton University Press, 1954.
  • –––, 1906 [2007], La Théorie Physique: son objet, sa structure , Paris: Vrin.
  • Fahrbach, Ludwig, 2011, “How the Growth of Science Ends Theory Change”, Synthese , 180(2): 139–155. doi:10.1007/s11229-009-9602-0
  • French, Steven, 2014, The Structure of the World: Metaphysics and Representation , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199684847.001.0001
  • Frost-Arnold, Greg, 2014, “Can the Pessimistic Induction be Saved from Semantic Anti-Realism about Scientific Theory?”, British Journal for the Philosophy of Science , 65(3): 521–548. doi:10.1093/bjps/axt013
  • Guegeun, Marie & Stathis Psillos, 2017, “Anti-Scepticism and Epistemic Humility in Pierre Duhem’s Philosophy of Science”, Transversal , 2: 54–73. doi:10.24117/2526-2270.2017.i2.06
  • Hesse, Mary B., 1976, “Truth and Growth of Knowledge”, in F. Suppe & P.D. Asquith (eds), PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association , volume 2, pp. 261–280, East Lansing: Philosophy of Science Association. doi:10.1086/psaprocbienmeetp.1976.2.192385
  • Kitcher, Philip, 1993, The Advancement of Science: Science Without Legend, Objectivity Without Illusions , Oxford: Oxford University Press. doi:10.1093/0195096533.001.0001
  • Kripke, Saul, 1972, “Naming and Necessity”, in Donald Davidson and Gilbert Harman (eds), Semantics of Natural Language , Dordrecht: Reidel pp. 253–355, 763–769.
  • Ladyman, James, 1998, “What is Structural Realism?”, Studies in History and Philosophy of Science , 29(3): 409–424. doi:10.1016/S0039-3681(98)80129-5
  • Ladyman, James & Don Ross, 2007, Every Thing Must Go: Metaphysics Naturalised , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199276196.001.0001
  • Lalande, André, 1913, “Philosophy in France in 1912”, The Philosophical Review , 22(4): 357–374. doi:10.2307/2178386
  • Laudan, Larry, 1981, “A Confutation of Convergent Realism”, Philosophy of Science , 48(1): 19–49. doi:10.1086/288975
  • Lewis, Peter J., 2001, “Why the Pessimistic Induction Is a Fallacy”, Synthese , 129(3): 371– 380. doi:10.1023/A:1013139410613
  • Lyons, Timothy D., 2002, “Scientific Realism and the Pessimistic Meta-Modus Tollens”, in S. Clarke and T.D. Lyons (eds.), Recent Themes in the Philosophy of Science: Scientific Realism and Commonsense , Dordrecht: Springer, pp. 63–90.
  • –––, 2006, “Scientific Realism and the Stratagema de Divide et Impera”, British Journal for the Philosophy of Science , 57(3): 537–560, doi:10.1093/bjps/axl021
  • –––, 2016a, “Structural Realism versus Deployment Realism: A Comparative Evaluation”, Studies in History and Philosophy of Science , 59: 95–105. doi:10.1016/j.shpsa.2016.06.006
  • –––, 2016b, “Scientific Realism”, in Paul Humphreys (ed.) The Oxford Handbook of Philosophy of Science , Oxford: Oxford University Press, pp. 564–584.
  • Magnus, P.D. & Craig Callender, 2004, “Realist Ennui and the Base Rate Fallacy”, Philosophy of Science , 71(3): 320–338. doi:10.1086/421536
  • Maxwell, Grover, 1970a, “Theories, Perception and Structural Realism”, in Robert Colodny (ed.) The Nature and Function of Scientific Theories: Essays in Contemporary Science and Philosophy , (University of Pittsburgh series in the philosophy of science, 4 ), Pittsburgh, PA: University of Pittsburgh Press, pp. 3–34.
  • –––, 1970b, “Structural Realism and the Meaning of Theoretical Terms”, in Michael Radner and Stephen Winokur (eds), Analyses of Theories and Methods of Physics and Psychology , (Minnesota Studies in the Philosophy of Science, 4), Minneapolis: University of Minnesota Press, pp. 181–192.
  • Mizrahi, Moti, 2013, “The Pessimistic Induction: A Bad Argument Gone Too Far”, Synthese , 190(15): 3209–3226. doi:10.1007/s11229-012-0138-3
  • –––, 2016, “The History of Science as a Graveyard of Theories: A Philosophers’ Myth?”, International Studies in Philosophy of Science , 30(3): 263–278. doi:10.1080/02698595.2017.1316113
  • Müller, Florian, 2015, “The Pessimistic Meta-induction: Obsolete Through Scientific Progress?”, International Studies in the Philosophy of Science , 29(4): 393–412. doi:10.1080/02698595.2015.1195144
  • Newman, M.H.A., 1928, “Mr. Russell’s ‘Causal Theory of Perception’”, Mind , 37(146): 137–148. doi:10.1093/mind/XXXVII.146.137
  • Newman, Mark, 2005, “Ramsey Sentence Realism as an Answer to the Pessimistic Meta-Induction”, Philosophy of Science , 72(5): 1373–1384. doi:10.1086/508975
  • Newton-Smith, W.H., 1981, The Rationality of Science , London: Routledge & Kegan Paul.
  • Papineau, David, 2010, “Realism, Ramsey , Sentences and the Pessimistic Meta-Induction”, Studies in History and Philosophy of Science , 41(4): 375–385. doi:10.1016/j.shpsa.2010.10.002
  • Park, Seungbae, 2011, “A Confutation of the Pessimistic Induction”, Journal for General Philosophy of Science , 42(1): 75–84. doi:10.1007/s10838-010-9130-0
  • –––, 2016, “Refutations of the Two Pessimistic Inductions”, Philosophia , 44(3): 835–844. doi:10.1007/s11406-016-9733-8
  • Paul, Harry W., 1968, “The Debate over the Bankruptcy of Science in 1895”, French Historical Studies , 5(3): 299–327. doi:10.2307/286043
  • Peters, Dean, 2014, “What Elements of Successful Scientific Theories Are the Correct Targets for ‘Selective’ Scientific Realism?”, Philosophy of Science , 81(3): 377–397. doi:10.1086/676537
  • Poincaré, Henri, 1900, “Sur les Rapports de la Physique Expérimentale et de la Physique Mathématique”, in Rapports Présentés au Congrès International de Physique , Vol.XCVI: 245–263.
  • –––, 1902, La Science et L’Hypothese , (1968 reprint) Paris: Flammarion.
  • Prevost, Jean Louis, and Dumas, Jean-Baptiste André, 1823, “Examen du Sang et de son Action dans les Divers Phènoménes de la Vie”, Journal de Physique, De Chimie et d’Histoire Naturelle , XCVI: 245–263. [ Prevost & Dumas 1823 available online ]
  • Psillos, Stathis, 1994, “A philosophical study of the transition from the caloric theory of heat to thermodynamics: Resisting the pessimistic meta-induction”, Studies in the History and Philosophy of Science , 25(2): 159–190. doi:10.1016/0039-3681(94)90026-4
  • –––, 1996, “Scientific Realism and the ‘Pessimistic Induction’”, Philosophy of Science , 63: S306–14. doi:10.1086/289965
  • –––, 1999, Scientific Realism: How Science Tracks Truth , London & New York: Routledge.
  • –––, 2009, Knowing the Structure of Nature: Essays on Realism and Explanation , London: Palgrave/MacMillan. doi:10.1057/9780230234666
  • –––, 2011, “Moving Molecules above the Scientific Horizon: On Perrin’s Case for Realism”, Journal for General Philosophy of Science , 42(2): 339–363. doi:10.1007/s10838-011-9165-x
  • –––, 2012, “Causal-descriptivism and the Reference of Theoretical Terms”, in Athanassios Raftopoulos & Peter Machamer (eds), Perception, Realism and the Problem of Reference , Cambridge University Press, pp. 212–238. doi:10.1017/CBO9780511979279.010
  • –––, 2014, “Conventions and Relations in Poincaré’s, Philosophy of Science ”, Methode-Analytic Perspectives , 3(4): 98–140.
  • Putnam, Hilary, 1973, “Explanation and Reference”, in Glenn Pearce & Patrick Maynard (eds), Conceptual Change , Dordrecht: Reidel, pp. 199–221. doi:10.1007/978-94-010-2548-5_11
  • –––, 1975, “The Meaning of ‘Meaning’”, in Keith Gunderson (ed.), Language, Mind and Knowledge , (Minnesota Studies in the Philosophy of Science, 7), Minneapolis: University of Minnesota Press, pp. 131–93.
  • –––, 1978, Meaning and the Moral Sciences , London; Boston: Routledge and Kegan Paul.
  • Ramsey, Frank Plumpton, 1929, “Theories”, in his The Foundations of Mathematics and Other Essays , R. B. Braithwaite (ed.), (1931) London: Routledge and Kegan Paul.
  • Richet, Charles, 1895, “ La Science a-t-elle fait banqueroute? ”, Revue Scientifique , No. 3, 12 January 1895, 33–39. [ Richet 1895 available online ]
  • Russell, Bertrand, 1927, The Analysis of Matter , London: George Allen & Unwin.
  • Ruttkamp-Bloem, Emma, 2013, “Re-enchanting Realism in Debate with Kyle Stanford”, Journal of General Philosophy of Science , 44(1): 201–224. doi:10.1007/s10838-013-9220-x
  • Saatsi, Juha, 2019, “Historical Inductions, Old and New”, Synthese , 196: 3979–3993. doi:10.1007/s11229-015-0855-5
  • Smith, George E., 2010, “Revisiting Accepted Science: The Indispensability of the History of Science”, The Monist , 93(4): 545–579. doi:10.5840/monist201093432
  • Stanford, P. Kyle, 2006, Exceeding Our Grasp: Science, History, and the Problem of Unconceived Alternatives , Oxford: Oxford University Press. doi:10.1093/0195174089.001.0001
  • Tolstoy, Leo, 1904, Essays & Letters , Aylmer Maud (trans.), New York: Funk and Wagnalls Company.
  • Tulodziecki, Dana, 2021 “Theoretical Continuity, Approximate Truth, and the Pessimistic Meta-Induction: Revisiting the Miasma Theory”, in T.D. Lyons and P. Vickers (eds.), Contemporary Scientific Realism: The Challenge from the History of Science , Oxford: Oxford University Press, pp. 11–32.
  • Vickers, Peter, 2013, “A Confrontation of Convergent Realism”, Philosophy of Science , 80(2): 189–211. doi:10.1086/670297
  • –––, 2017, “Understanding the Selective Realist Defence Against the PMI”, Synthese , 194(9): 3221–3232. doi:10.1007/s11229-016-1082-4
  • Worrall, John, 1989, “Structural Realism: The Best of Both Worlds?”, Dialectica , 43(1–2): 99–124. doi:10.1111/j.1746-8361.1989.tb00933.x
  • Wray, K. Brad, 2013, “The Pessimistic Induction and the Exponential Growth of Science Reassessed”, Synthese , 190(18): 4321–4330. doi:10.1007/s11229-013-0276-2
  • –––, 2015, “Pessimistic Inductions: Four Varieties”, International Studies in the Philosophy of Science , 29(1): 61–73. doi:10.1080/02698595.2015.1071551
  • Zahar, Elie, 2001, Poincaré’s Philosophy: From Conventionalism to Phenomenology , LaSalle IL: Open Court.
How to cite this entry . Preview the PDF version of this entry at the Friends of the SEP Society . Look up topics and thinkers related to this entry at the Internet Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers , with links to its database.
  • Scientific Realism and Antirealism , entry by Michael Liston in the Internet Encyclopedia of Philosophy .
  • Research Guide on Realism and anti-realism in the philosophy of Science , by Paul Dicken, History and Philoosphy of Science, Cambridge University.

-->Boltzmann, Ludwig --> | Duhem, Pierre | Poincaré, Henri | scientific realism | structural realism | truthlikeness

Acknowledgments

Many thanks to Ludwig Fahrbach, Stavros Ioannidis, Moti Mizrahi, Nathan Oseroff, Seungbae Park and Brad Wray for useful comments. Thanks are also due to the Editors of SEP and various anonymous reviewers for their encouragement and suggestions. Many thanks are also due to my student Kosmas Brousalis for his help with the revisions of the entry.

Copyright © 2022 by Stathis Psillos < psillos @ phs . uoa . gr >

  • Accessibility

Support SEP

Mirror sites.

View this site from another server:

  • Info about mirror sites

The Stanford Encyclopedia of Philosophy is copyright © 2023 by The Metaphysics Research Lab , Department of Philosophy, Stanford University

Library of Congress Catalog Data: ISSN 1095-5054

What is Scientific Realism?

Decades of debate about scientific realism notwithstanding, we find ourselves bemused by what different philosophers appear to think it is, exactly. Does it require any sort of belief in relation to scientific theories and, if so, what sort? Is it rather typified by a certain understanding of the rationality of such beliefs? In the following dialogue we explore these questions in hopes of clarifying some convictions about what scientific realism is, and what it could or should be. En route, we encounter some profoundly divergent conceptions of the nature of science and of philosophy.

  • Related Documents

Editor's Introduction

The debate over scientific realism, simply put, is a debate over what we can and should believe about reality once we've critically assessed all the available arguments and empirical evidence. Thinking earnestly about the merits of scientific realism as a philosophical thesis requires navigating contentious historiographical issues, being familiar with the technical details of various scientific theories, and addressing disparate philosophical problems spanning aesthetics, metaphysics, epistemology, and beyond. This issue of Spontaneous Generations: A Journal for the History and Philosophy of Science aims to make participating in the scientific realism debate easier for both newcomers and veterans, collecting over twenty invited and peer-reviewed papers under the title "The Future of the Scientific Realism Debate: Contemporary Issues Concerning Scientific Realism."

Scientific Realism versus Antirealism in Science Education

Scientific realists believe both what a scientific theory says about observables and unobservables. In contrast, scientific antirealists believe what a scientific theory says about observables, but not about unobservables. I argue that scientific realism is a more useful doctrine than scientific antirealism in science classrooms. If science teachers are antirealists, they are caught in Moore’s paradox when they help their students grasp the content of a scientific theory, and when they explain a phenomenon in terms of a scientific theory. Teachers ask questions to their students to check whether they have grasped the content of a scientific theory. If the students are antirealists, they are also caught in Moore’s paradox when they respond positively to their teachers’ questions, and when they explain a phenomenon in terms of a scientific theory. Finally, neither teachers nor students can understand phenomena in terms of scientific theories, if they are antirealists.

On the metaphysics of (epistemological) logical anti-exceptionalism

A recent logical anti-exceptionalist trend proposes that logical theories are revisable in the same manner as scientific theories, either on grounds of the method of theory selection or on what counts as evidence for this revision. Given this approximation of logic and science, the present essay analyzes the commitments of both these varieties and argues that, as it currently stands, this kind of anti-exceptionalism is committed to scientific realism, that is, to realism about some unobservable entities evoked in logical theories. The essay argues that anti-exceptionalism cannot be separated into metaphysical and epistemological varieties, and proposed rather to label anti-exceptionalists views either broadly in terms of theory revision, or narrowly in terms of logic’s affinity with science.

From the Evidence of History to the History of Evidence

This chapter looks into the transition from the Cartesian natural philosophy to the Newtonian one, and then to the Einsteinian science, making the following key point: though the shift from Descartes’s theory to Newton’s amounted to a wholesale rejection of Descartes’s theory, in the second shift, a great deal was retained; Newton’s theory of universal gravitation gave rise to a research program that informed and constrained Einstein’s theory. Newton’s theory was a lot more supported by the evidence than Descartes’s and this made it imperative for the successor theory to accommodate within it as much as possible of Newton’s theory: evidence for Newton’s theory became evidence for Einstein’s. This double case study motivates a rebranding of the “divide et impera” strategy against the pessimistic induction introduced in the book Scientific Realism, which shifts attention from the (crude) evidence of the history of science to the (refined) history of evidence for scientific theories.

1. The methods and fruits of science

This chapter briefly discusses central key topics in the philosophy of science that the remainder of the book draws upon. It begins by considering the scientific method. ‘Induction’—the idea that we construct scientific theories just by generalizing from observations—is a very poor match to real science. ‘Falsification’—Popper’s idea that we create a theory, test against observation, and discard it if it fails the test—is much more realistic, but still too simple: data only falsifies data given auxiliary assumptions that can themselves be doubted. The issues are illustrated through an example from modern astrophysics: dark matter. The chapter then explores how we can resolve issues of underdetermination, where two theories give the same predictions. Finally, it introduces ‘scientific realism’, the view that our best theories tell us things about the world that go beyond what is directly observable.

Scientific realism and antirealism

Traditionally, scientific realism asserts that the objects of scientific knowledge exist independently of the minds or acts of scientists and that scientific theories are true of that objective (mind-independent) world. The reference to knowledge points to the dual character of scientific realism. On the one hand it is a metaphysical (specifically, an ontological) doctrine, claiming the independent existence of certain entities. On the other hand it is an epistemological doctrine asserting that we can know what individuals exist and that we can find out the truth of the theories or laws that govern them. Opposed to scientific realism (hereafter just ‘realism’) are a variety of antirealisms, including phenomenalism and empiricism. Recently two others, instrumentalism and constructivism, have posed special challenges to realism. Instrumentalism regards the objects of knowledge pragmatically, as tools for various human purposes, and so takes reliability (or empirical adequacy) rather than truth as scientifically central. A version of this, fictionalism, contests the existence of many of the objects favoured by the realist and regards them as merely expedient means to useful ends. Constructivism maintains that scientific knowledge is socially constituted, that ‘facts’ are made by us. Thus it challenges the objectivity of knowledge, as the realist understands objectivity, and the independent existence that realism is after. Conventionalism, holding that the truths of science ultimately rest on man-made conventions, is allied to constructivism. Realism and antirealism propose competing interpretations of science as a whole. They even differ over what requires explanation, with realism demanding that more be explained and antirealism less.

The Role of Existential Quantification in Scientific Realism

AbstractScientific realism holds that the terms in our scientific theories refer and that we should believe in their existence. This presupposes a certain understanding of quantification, namely that it is ontologically committing, which I challenge in this paper. I argue that the ontological loading of the quantifiers is smuggled in through restricting the domains of quantification, without which it is clear to see that quantifiers are ontologically neutral. Once we remove domain restrictions, domains of quantification can include non-existent things, as they do in scientific theorizing. Scientific realism would therefore require redefining without presupposing a view of ontologically committing quantification.

Unconceived alternatives and another argument for instrumentalism

Selective skepticism in relation to fundamental scientific theories and criticism of the inference to the best explanation as an eliminative approach to substantiate hypotheses, enable K. Stanford to interpret and combine in his own way the classical arguments against the scientific realism – the arguments of the pessimistic meta-induction and that of the underdetermination of theory by data. Despite the fact that his justification of the instrumentalist interpretation of scientific knowledge is just another version of the argument «from error», K. Stanford’s book should be recommended to a scientific realism could be. Reflection on the book: Stanford K. Exceeding Our Grasp: Science, History, and the Problem of Unconceived Alternatives. Oxford University Press, 2006.

El contenido empírico del realismo científico

RESUMENUna forma común de entender el realismo científico (RC) en las últimas décadas ha sido plantearlo como una inferencia explicativa: RC es la mejor explicación del éxito predictivo-instrumental de la ciencia. Algunos de sus partidarios mantienen, además que es una hipótesis empíricamente constrastable. Intentaré argumentar, que, entendido así, RC no es empíricamente contrastable. En primer lugar, aunque el éxito predictivo-instrumental initerrumpido de una teoría T es una consecuencia observacional de la verdad de T, este hecho no hecho no constituye una evidencia empírica diferente del propio "explanandum". En segundo lugar, elaorar un registro histórico del éxito -no sólo empírico, sino teórico- obtenido mediante la postulación de entidades por consideraciones explicativas, confirmaría como mucho, y eso suponiendo que fuera posible, una cocincidencia entre una metodología determinada y unos resultados, pero no daría cuenta del vínculo explicativo entre éxito predictivo instrumental por un lado, y verdad y existencia, por otro. Por consiguiente, RC no es una hipótesis empírica en un sentido genuino; a fortiori, tampoco es una hipótesis científica. Esta conclusión, no obstante, no cierra el camino a un realismo científico de carácter local.PALABRAS CLAVEREALISMO CIENTÍFICO, OBSERVACIÓN, TEORÍA, INFERENCIA A LA MEJOR EXPLICACIÓNABSTRACTA common way of understanding scientific realism (SR) during the latest decades says that SR is the best explanation of the predictive success enjoyed by scientific theories. Some os this advocates claim, aslo, that SR is an empirically testable hypothesis. I will try to argue that, as an explanation of predictive sucess, SR is not empirically testable. Firstly, even though the uninterrupted preditive success of T is an observational consequence of T´s truth, this fact is not a kind of evidence distinguishable from the very explanandum. Secondly, a historical record of success obtained by postulating theoretical emities would confirm, at most, a correlation between some methodological norms and some particular results. But confirming such correlation is not the same as vindicating an explanotory link between truth and existence (the explanans), and predictive success (the explanandum). In sum, SR is not a genuine empirical hypothesis; a fortiori, it is neither a scientific hypothesis. Anyway, this conclusion does not forbid some kind of "local" scientific realism.KEYWORDSSCIENTIFIC REALISM, OBSERVATION, THEORY, INFERENCE TO THE BEST EXPLANATION

Scientific Realism

This chapter addresses scientific realism. After the heyday of empiricism in the interwar period and its immediate aftermath, many critical reactions to empiricism seemed to suggest scientific realism. It was widely agreed that scientific theories make references to things that cannot be directly observed (or at least seen), and thus emerged the issue of the status of non-observables. As scientific realism became increasingly dominant, new philosophical stances such as Bas C. van Fraassen’s constructive empiricism were often defined in opposition to it. Van Fraassen understands scientific realism as a claim that science aims to give us, in its theories, a literally true story of what the world is like; and acceptance of a scientific theory involves the belief that it is true. More in line with established forms of scientific realism, Ilkka Niiniluoto talks about verisimilitude, or truth-likeness. This concept is supposed to avoid the consequences of claiming to have access to the truth itself. The chapter then considers how the social sciences seem to pose difficulties for scientific realism.

Export Citation Format

Share document.

  • Advanced Search
  • All new items
  • Journal articles
  • Manuscripts
  • All Categories
  • Metaphysics and Epistemology
  • Epistemology
  • Metaphilosophy
  • Metaphysics
  • Philosophy of Action
  • Philosophy of Language
  • Philosophy of Mind
  • Philosophy of Religion
  • Value Theory
  • Applied Ethics
  • Meta-Ethics
  • Normative Ethics
  • Philosophy of Gender, Race, and Sexuality
  • Philosophy of Law
  • Social and Political Philosophy
  • Value Theory, Miscellaneous
  • Science, Logic, and Mathematics
  • Logic and Philosophy of Logic
  • Philosophy of Biology
  • Philosophy of Cognitive Science
  • Philosophy of Computing and Information
  • Philosophy of Mathematics
  • Philosophy of Physical Science
  • Philosophy of Social Science
  • Philosophy of Probability
  • General Philosophy of Science
  • Philosophy of Science, Misc
  • History of Western Philosophy
  • Ancient Greek and Roman Philosophy
  • Medieval and Renaissance Philosophy
  • 17th/18th Century Philosophy
  • 19th Century Philosophy
  • 20th Century Philosophy
  • History of Western Philosophy, Misc
  • Philosophical Traditions
  • African/Africana Philosophy
  • Asian Philosophy
  • Continental Philosophy
  • European Philosophy
  • Philosophy of the Americas
  • Philosophical Traditions, Miscellaneous
  • Philosophy, Misc
  • Philosophy, Introductions and Anthologies
  • Philosophy, General Works
  • Teaching Philosophy
  • Philosophy, Miscellaneous
  • Other Academic Areas
  • Natural Sciences
  • Social Sciences
  • Cognitive Sciences
  • Formal Sciences
  • Arts and Humanities
  • Professional Areas
  • Other Academic Areas, Misc
  • Submit a book or article
  • Upload a bibliography
  • Personal page tracking
  • Archives we track
  • Information for publishers
  • Introduction
  • Submitting to PhilPapers
  • Frequently Asked Questions
  • Subscriptions
  • Editor's Guide
  • The Categorization Project
  • For Publishers
  • For Archive Admins
  • PhilPapers Surveys
  • Bargain Finder
  • About PhilPapers
  • Create an account

Scientific realism: selected essays of Mario Bunge

Author's profile.

scientific realism essay

Reprint years

Call number, buy this book.

scientific realism essay

PhilArchive

External links.

scientific realism essay

  • Available at Amazon.com

Through your library

  • Sign in / register and customize your OpenURL resolver
  • Configure custom resolver

Similar books and articles

Citations of this work.

No citations found.

References found in this work

No references found.

Phiosophy Documentation Center

234CHAPTER 11 Scientific Realism and Chemistry

  • Published: July 2016
  • Cite Icon Cite
  • Permissions Icon Permissions

SCIENTIFIC REALISM Is a philosophical issue with relevance to all sciences, but there are some particularly interesting and distinctive ways in which it has manifested itself in chemistry. Paying proper attention to such aspects will deliver two types of benefits: First, it will aid the philosophical understanding of the nature of chemical knowledge; second, it will throw some fresh light on the realism debate in places where it has developed without much attention to chemical practices and chemical concepts. In the following discussion I will attempt to make a reasonably comprehensive survey of relevant literature, while also advancing some original points and viewpoints. Recall Bas van Fraassen’s now-classic formulation of the realism debate as an argument about whether we can know about unobservable entities featuring in scientific theories, and whether we should try to know about them (van Fraassen 1980). If this is how we understand realism, and if we take the long view of the history of science, chemistry is the most important science to consider in the realism debate. Until the development of atomic, nuclear, and elementary-particle physics starting in the early twentieth century, chemistry was the science in which debates about the epistemic and ontological status of unobservable theoretical entities took place with most ferocity and most relevance to practice. An interesting contrast is astronomy, in which the Copernican Revolution brought in a long and secure phase of realism about astronomical objects far out of reach of any human senses (including those that do not even register as tiny specks of light to our eyes). In contrast, the achievements of chemistry up to the early nineteenth century only deepened the sense of inaccessibility and unobservability concerning the putative fundamental entities postulated in chemical theories. Unobservability in relation to chemical theories is not only an issue about atomism, though surely the problem was clearly present with the atomistic particles imagined by a wide range of thinkers from Democritus and Leucippus of ancient times to Descartes and other early-modern mechanical philosophers.

Signed in as

Institutional accounts.

  • Google Scholar Indexing
  • GoogleCrawler [DO NOT DELETE]

Personal account

  • Sign in with email/username & password
  • Get email alerts
  • Save searches
  • Purchase content
  • Activate your purchase/trial code

Institutional access

  • Sign in with a library card Sign in with username/password Recommend to your librarian
  • Institutional account management
  • Get help with access

Access to content on Oxford Academic is often provided through institutional subscriptions and purchases. If you are a member of an institution with an active account, you may be able to access content in one of the following ways:

IP based access

Typically, access is provided across an institutional network to a range of IP addresses. This authentication occurs automatically, and it is not possible to sign out of an IP authenticated account.

Sign in through your institution

Choose this option to get remote access when outside your institution. Shibboleth/Open Athens technology is used to provide single sign-on between your institution’s website and Oxford Academic.

  • Click Sign in through your institution.
  • Select your institution from the list provided, which will take you to your institution's website to sign in.
  • When on the institution site, please use the credentials provided by your institution. Do not use an Oxford Academic personal account.
  • Following successful sign in, you will be returned to Oxford Academic.

If your institution is not listed or you cannot sign in to your institution’s website, please contact your librarian or administrator.

Sign in with a library card

Enter your library card number to sign in. If you cannot sign in, please contact your librarian.

Society Members

Society member access to a journal is achieved in one of the following ways:

Sign in through society site

Many societies offer single sign-on between the society website and Oxford Academic. If you see ‘Sign in through society site’ in the sign in pane within a journal:

  • Click Sign in through society site.
  • When on the society site, please use the credentials provided by that society. Do not use an Oxford Academic personal account.

If you do not have a society account or have forgotten your username or password, please contact your society.

Sign in using a personal account

Some societies use Oxford Academic personal accounts to provide access to their members. See below.

A personal account can be used to get email alerts, save searches, purchase content, and activate subscriptions.

Some societies use Oxford Academic personal accounts to provide access to their members.

Viewing your signed in accounts

Click the account icon in the top right to:

  • View your signed in personal account and access account management features.
  • View the institutional accounts that are providing access.

Signed in but can't access content

Oxford Academic is home to a wide variety of products. The institutional subscription may not cover the content that you are trying to access. If you believe you should have access to that content, please contact your librarian.

For librarians and administrators, your personal account also provides access to institutional account management. Here you will find options to view and activate subscriptions, manage institutional settings and access options, access usage statistics, and more.

Our books are available by subscription or purchase to libraries and institutions.

  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Rights and permissions
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

The Role of the ‘No Miracles’ Argument for Scientific Realism Essay

The most general explanation of the principles of scientific realism is based on the idea that science can provide the true description of the real world and details of the world’s processes.

To justify the appropriateness of the philosophy, the supporters of scientific realism are inclined to promote the ‘no miracles’ argument which is discussed as rather debatable among anti-realists (French & Saatsi 2011, p. 84). That is why, it is necessary to discuss the role of the argument for supporting scientific realism.

This significant role was accentuated by Hilary Putnam as the developer of the argument. According to Putnam’s discussion of the argument, if the scientific theory is not true, it is a miracle that this theory can generate effective and working predictions, and moreover, “the positive argument for realism is that it is the only philosophy that doesn’t make the success of science a miracle” (Putnam, cited in Norris 2002, p. 218).

On the one hand, the ‘no miracles’ argument is rather convincing to support scientific realism because it is the strongest argument based on the principle of the empirical success.

On the other hand, the ‘no miracles’ argument cannot be discussed as convincing to support the philosophy because it fails to explain why many successful scientific theories of the past are not discussed as true ones today.

As a result, to conclude about the appropriateness of the argument, it is important to focus on the ideas of such anti-realists as Bas Van Fraassen and Larry Laudan. Moreover, it is necessary to discuss the ‘no miracles’ argument from the point of scientific realists who support the argument and refer to Alan Musgrave and Ian Hacking’s views.

The ‘no miracles’ argument is also known as the ‘success’ or ‘ultimate’ argument. According to this argument, the scientific theories are successfully accepted and used because they are approximately true. If these theories are not true, their success is rather miraculous.

Thus, the success of the theories is grounded on their truth and empirical evidence, but not miracles (French & Saatsi 2011, p. 84-85). The argument provides the non-miraculous alternative to speak about the success of scientific theories, and the empirical success should be taken into account while discussing the effectiveness of theories (Clarke & Lyons 2002, p. xi).

While referring to the ideas of the above-mentioned philosophers of science, it is reasonable to start with the discussion of anti-realists’ ideas and their visions of the effectiveness of the ‘no miracles’ argument because in their works, Musgrave and Hacking are inclined to support the argument as important for scientific realism in response to Van Fraassen and Laudan’s discussions.

Scientific realism should be discussed from the perspective of its three dimensions, which are the commitment to the development of the mind-independent world; the specific semantic commitment to the literal interpretation of hypotheses, and the epistemological commitment to discussing scientific entities (Sankey 2012, p. 34).

Metaphysically, commitment to the mind-independent world means that scientists develop theories to explain the objective or external reality which is independent from their mind (Sankey 2012, p. 7).

The semantic commitment explains the scientists’ focus on discussing claims about different entities as literally true and valuable even if they are not observed. The epistemological commitment is the focus on theoretical statements as explaining and forming the true knowledge of the whole world (Sankey 2012, p. 36).

In his works, Van Fraassen refers to the ideas of constructive empiricism, discusses the dimensions of scientific realism and proposes the improved statement of the philosophy of scientific realism while discussing the effectiveness of the ‘no miracles’ argument as the ‘ultimate’ argument (Van Fraassen 2013, p. 1078-1079).

Thus, Van Fraassen is inclined to propose the more developed and accurate statement referred to scientific realism, “science aims to give us, in its theories, a literally true story of what the world is like; and acceptance of a scientific theory involves the belief that it is true.

This is the correct statement of scientific realism” (Van Fraassen 2013, p. 1062). Thus, the philosopher focuses on the importance of the science’s aim and on the role of the belief to speak about the theory as true or not. Moreover, according to Van Fraassen, scientific theories can be discussed as successful when they are literally construed.

To oppose the ‘no miracles’ argument, Van Fraassen chooses not only to develop the definition of scientific realism but also focuses on the discussion of the alternative vision, which is more effective than scientific realism to discuss the relations between the world and science.

Thus, Van Fraassen states that scientific realism in contrast to constructive empiricism cannot provide the real platform for discussing the successful theory and notes that “acceptance of a theory involves as belief only that it is empirically adequate” (Van Fraassen 2013, p. 1065).

From this point, constructive empiricism is more effective to speak about the success of scientific theories because it operates the idea of adequacy.

Pointing at the weaknesses of scientific realism, Van Fraassen develops the alternative explanation to the ‘no miracles’ argument, according to which it is even unnecessary to explain the nature of the scientific theories’ success because this success “is no miracle” (Van Fraassen 2013, p. 1080).

Van Fraassen is inclined to discuss the success of the definite scientific theories as the natural process, which is similar to survival because “only the successful theories survive – the ones which in fact latched on to actual regularities in nature” (Van Fraassen 2013, p. 1080).

On the one hand, Van Fraassen’s approach can be discussed as the indirect criticism of the ‘no miracles’ argument because the philosopher chooses to discuss the successful theories from the unusual perspective and with references to the Darwinist’s view.

On the other hand, Van Fraassen adds to the discussion of the role of the ‘no miracles’ argument for scientific realism because of improving the statement of scientific realism’s principles.

From this perspective, Van Fraassen cannot provide the effective arguments and evidences to state that the ‘no miracles’ argument is irrelevant because he avoids explaining of the reasons for the theories’ success.

One of the anti-realists’ responses to the ‘no miracles’ argument is the researches conducted by Laudan to state that many successful theories of the past are not supported today.

In his works, Laudan is inclined to build his argument against the ‘no miracles’ argument as the platform of scientific realism while evaluating the effectiveness of the argument in relation to meeting the dimensions of scientific realism (Laudan 2013, p. 1110).

Thus, Laudan focuses on the central question of the debates among realists and anti-realists, which is “whether the realist’s assertions about the interrelations between truth, reference and success are sound” (Laudan 2013, p. 1111).

While focusing on discussing the correlation between the mentioned notions, Laudan pays attention to the fact that the realists’ argument fail because it has weaknesses to explain the success of the theories from the point of semantic and epistemic perspectives as well as from the point of truth in the theoretical context.

To develop his argument, Laudan continues to focus on semantic and epistemic aspects as significant to conclude about the relevance of realists’ visions.

That is why, according to Laudan, even if the realist had “a semantically adequate characterisation of approximate or partial truth, and even if that semantics entailed that most of the consequences of an approximately true theory would be true”, this realist would have no any criterion “that would epistemically warrant the ascription of approximate truth to a theory” (Laudan 2013, p. 1119).

As a result, basing on the set dimensions and criteria, realists cannot provide explanations to the theories “which are not approximately true”, but often successful, as it is relevant while discussing the theories of the past (Laudan 2013, p. 1124).

Thus, according to Laudan, the aspects of the theories’ successfulness and effective empirical evidence associated with the ‘no miracles’ argument cannot serve to reflect the real world appropriately.

As a result, if the ‘no miracles’ argument fails to explain the success of the definite theories in the history of science, this argument also fails to support the idea of scientific realism effectively.

However, Laudan is rather focused to state that realists fail to explain the success of theories in the historical context, and this approach prevents him from looking at the problem from many perspectives (Hudson 2013, p. 203-204).

That is why, Laudan’s position is directly opposed to Musgrave’s one, and Musgrave provides many effective arguments to support his focus on the positive role of the ‘no miracles’ argument for scientific realism.

If Van Fraassen and Laudan as anti-realists are inclined to provide the sound evidence and support to state that the idea of scientific realism based on the platform of the ‘no miracles’ argument is not appropriate to discuss the true character of many successful scientific theories, Musgrave and Hacking focus on responding to the opponents while supporting the role of the ‘no miracles’ argument for scientific realism.

Thus, Musgrave builds his argument for scientific realism with its theoretical platform and against the anti-realists’ visions while evaluating Van Fraassen’s approach in detail.

According to Musgrave, Van Fraassen’s Darwinian explanation can be “accepted by realists and anti-realists alike”, but Musgrave also pays attention to the fact that “to say that only successful theories are allowed to survive is not to explain why any particular theory is successful” (Musgrave 2013, p. 1094).

As a result, Van Fraassen’s approach cannot be discussed as threatening to the idea of the ‘no miracles’ argument because the philosopher’s argument should be considered as the alternative vision of the problem.

According to the ‘no miracles’ argument promoted by Musgrave, successful scientific theories make claims which are true or approximately true because the entities mentioned in the claims and scientific theories really exist. As a result, the success of the theories is closely connected with the empirical success.

That is why, it is not a miracle that the scientific claims are associated with the empirical success because the entities mentioned in the theories exist, and the theories as well as their principles are true or approximately true (Musgrave 2013, p. 1087; Sankey 2012, p. 130).

Thus, according to Musgrave, anti-realism fails to provide the appropriate explanation to speak about the success of theories (Sankey 2012, p. 130). That is why, the ‘no miracles’ argument or the ‘ultimate’ argument serves perfectly to support scientific realism in order to explain the success of scientific theories.

While referring to Van Fraassen’s theory, it is also important to note that his constructive empiricism “is weaker than earlier anti-realist views in all kinds of ways, and correspondingly closer to realism” (Musgrave 2013, p. 1105).

As a result, Musgrave supports the idea that Van Fraassen in his discussions of the successful theories can be considered as closer to realists than anti-realists that is why scientific realism seems to be the most appropriate philosophy to discuss and explain the essence of the scientific theories’ success.

Furthermore, it is important to pay attention to the fact that Musgrave concentrates on the ‘no miracles’ argument as the best explanation to the theories’ success with references to the idea of predictive novelty.

Thus, anti-realists build their arguments against the effectiveness of scientific realism with references to the fact that the ‘ultimate’ argument cannot provide the explanation to the successfulness of the theories which are not true today (Sankey 2012, p. 130).

However, Musgrave points at the specific idea of the predictive novelty which means that it is reasonable to concentrate on hypotheses which can be useful to predict unknown facts rather than on hypotheses operating known details; thus, those theories are successful which serve to make novel, but credible predictions (Sankey 2012, p. 132).

This approach allows the further discussion of the approximately true theories focused on the novel knowledge. In this situation, Musgrave is inclined to use the epistemic terms in order to speak about the ‘ultimate’ or ‘no miracles’ argument, and this fact contributes to his discussion of scientific realism (Sankey 2012, p. 132).

From this point, Musgrave’s claims that Van Fraassen’s theory fails to explain the success of scientific theories, and Van Fraassen’s discussions are rather correlated with realists’ ones because of using the same terms and concentrating on the same relations between the ideas of truth and success (Musgrave 2013, p. 1094).

Furthermore, Musgrave supports the ‘ultimate’ argument while focusing on the concept of the novelty to explain the aspects which are predominantly discussed by such anti-realists as Laudan as the main ones to argue against the principles of scientific realism (Psillos 2005, p. 70-72).

As a result, Musgrave’s focus on the predictive novelty contributes to the support of the ‘no miracles’ argument, and his explanations can diminish the threatening effect of Van Fraassen’s alternative approach on the notions of scientific realism.

In this situation, Musgrave’s discussion is effective to support the unique role of the ‘no miracles’ argument for scientific realism, and this argument sounds as really ‘ultimate’ for the philosophy.

However, to state about the appropriateness of the ‘no miracles’ argument to support scientific realism, it is also important to refer to the ideas proposed by Hacking. On the one hand, the philosopher’s claims can be discussed as defending in relation to the ideas of scientific realism in general and ‘no miracles’ argument in particular.

On the other hand, Hacking chooses one more alternative position and focuses on the discussion of the role of experiments rather than theoretical arguments for stating about the success of scientific theories (Psillos 2005, p. 303).

Thus, Hacking pays attention to the fact that “discussions about scientific realism or anti-realism usually talk about theories, explanation and prediction”, however, it is necessary to focus on the fact that “only at the level of experimental practice is scientific realism unavoidable” (Hacking 2013, p. 1140).

According to Hacking, experiments and the scientists’ practical activities are more important than the discussions of theories.

Thus, to discuss whether the theories are true, it is necessary to conduct the experiments in order to conclude about the use and existence of entities. Hacking tries to draw the readers’ attention to the problem while stating, “think about practice, not theory” (Hacking 2013, p. 1153).

That is why, the philosopher presents the supporting facts to conclude that theoretical debates cannot provide scientists with more support and evidence to decide on the effectiveness of this or that theory and claim (Hudson 2013, p. 172-173).

From this point, Hacking’s approach cannot add significantly to the discussion of the effectiveness of the ‘ultimate’ argument for developing the principles of scientific realism because the philosopher shares the alternative position (Hacking 2013, p. 1142).

As a result, Hacking’s theory is discussed as closer to supporting scientific realism with its platform than to supporting anti-realists’ visions.

While focusing on the ideas of Bas Van Fraassen, Larry Laudan, Alan Musgrave, and Ian Hacking on the relevance of the ‘no miracles’ argument in relation to scientific realism, it is possible to share the vision of the argument’s supporters.

Thus, the ‘ultimate’ argument can be discussed as the convincing defence of scientific realism because it is the strongest argument to explain the scientific theories’ success while referring to all three dimensions of scientific realism.

In spite of the fact that anti-realists can provide significant arguments to oppose the principles of scientific realism, the ‘no miracles’ argument is the most detailed approach to explain the success of theories while basing on the important theoretical and empirical background (Okasha 2002, p. 66; Psillos 2005, p. 69).

In this case, the most convincing arguments to discuss the ‘ultimate’ argument as the perfect platform for scientific realism are provided by Musgrave who is able to explain all the weaknesses mentioned by anti-realists while operating the notion of the predictive novelty (Sankey 2012, p. 132).

The ‘no miracles’ argument, which states that claims of the successful theories are true and they reflect the real world’s processes seems to be convincing because facts and empirical evidences are traditionally discussed as the convincing arguments to support the theory.

As a result, the ‘no miracles’ argument is effective to explain the success of the scientific theories to support the principles of scientific realism. On the contrary, Van Fraassen’s alternative vision based on the theory of survival is not effective to explain the success of the theories, but only to determine it.

Furthermore, Laudan’s approach is also debatable because it can be criticised with references to Musgrave’s notion of the predictive novelty. In addition, Hacking’s approach can be interpreted from the point of empirical success to support the effectiveness of the ‘ultimate’ argument.

That is why, the ‘no miracles’ argument can be discussed as convincing because there are no opposite arguments which are as effective and detailed as the ‘ultimate’ argument to explain the processes of the real world with references to the theoretical notions and grounds.

The ‘no miracles’ argument also seems to be convincing as the defence of scientific realism while responding to the three dimensions of scientific realism because this philosophy involves the discussion of the argument’s semantic and epistemic nature along with the focus on the possible truth of scientists’ predictions.

Reference List

Clarke, S. & Lyons, T. 2002, Recent themes in the philosophy of science: scientific realism and commonsense , Springer, USA.

French, S. & Saatsi, J. 2011, The Continuum companion to the philosophy of science , Bloomsbury, UK.

Hacking, I. 2013, ‘Experimentation and scientific realism’, in M Curd, J Cover, & C Pincock (eds), Philosophy of science: the central issues, Norton & Company, USA, pp. 1140-1155.

Hudson, R. 2013, Seeing things: the philosophy of reliable observation , Oxford University Press, UK.

Laudan, L. 2013, ‘A confutation of convergent realism’, in M Curd, J Cover, & C Pincock (eds), Philosophy of science: the central issues, Norton & Company, USA, pp. 1108-1125.

Musgrave, A. 2013, ‘Realism versus constructive empiricism’, in M Curd, J Cover, & C Pincock (eds), Philosophy of science: the central issues, Norton & Company, USA, pp. 1083-1106.

Norris, C. 2002, Hilary Putnam: realism, reason and the uses of uncertainty , Manchester University Press, UK.

Okasha, S. 2002, Philosophy of science: a very short introduction , Oxford University Press, UK.

Psillos, S. 2005, Scientific realism: how science tracks truth , Routledge, USA.

Sankey, H. 2012, Scientific realism and the rationality of science , Ashgate Publishing, Ltd., Australia.

Van Fraassen, B. 2013, ‘Arguments concerning scientific realism’, in M Curd, J Cover, & C Pincock (eds), Philosophy of science: the central issues, Norton & Company, USA, pp. 1060-1081.

  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2020, February 13). The Role of the ‘No Miracles’ Argument for Scientific Realism. https://ivypanda.com/essays/the-role-of-the-no-miracles-argument-for-scientific-realism-essay/

"The Role of the ‘No Miracles’ Argument for Scientific Realism." IvyPanda , 13 Feb. 2020, ivypanda.com/essays/the-role-of-the-no-miracles-argument-for-scientific-realism-essay/.

IvyPanda . (2020) 'The Role of the ‘No Miracles’ Argument for Scientific Realism'. 13 February.

IvyPanda . 2020. "The Role of the ‘No Miracles’ Argument for Scientific Realism." February 13, 2020. https://ivypanda.com/essays/the-role-of-the-no-miracles-argument-for-scientific-realism-essay/.

1. IvyPanda . "The Role of the ‘No Miracles’ Argument for Scientific Realism." February 13, 2020. https://ivypanda.com/essays/the-role-of-the-no-miracles-argument-for-scientific-realism-essay/.

Bibliography

IvyPanda . "The Role of the ‘No Miracles’ Argument for Scientific Realism." February 13, 2020. https://ivypanda.com/essays/the-role-of-the-no-miracles-argument-for-scientific-realism-essay/.

  • "The Cuisine and Empire" by Rachel Laudan: Cooking in World History
  • Observation and Theory in Aristotle’s Scientific Practice
  • Does the Universe Have A Direction?
  • Conan Doyle’s Gender Conception
  • Research and Evaluation Theory in Project Management
  • Hume’s View of Miracles
  • Should we be realist or non-realist towards religious claims
  • David Hume on Miracles: Philosophical Critique
  • Philosophy Issues: David Hume and Miracles
  • A Miracle as an Extraordinary Happening Occurring in the Physical World
  • The Problem of Demarcation
  • Popper on Corroboration
  • Aligning Philosophies of Science with Research Approaches
  • Inductivist Method of Science
  • Objectivity of Science

An Essay for Educators: Epistemological Realism Really is Common Sense

  • Published: 15 June 2007
  • Volume 17 , pages 425–447, ( 2008 )

Cite this article

  • William W. Cobern 1 &
  • Cathleen C. Loving 2  

1096 Accesses

28 Citations

1 Altmetric

Explore all metrics

“What is truth?” Pontius Pilot asked Jesus of Nazareth. For many educators today this question seems quaintly passé. Rejection of “truth” goes hand-in-hand with the rejection of epistemological realism. Educational thought over the last decade has instead been dominated by empiricist, anti-realist, instrumentalist epistemologies of two types: first by psychological constructivism and later by social constructivism. Social constructivism subsequently has been pressed to its logical conclusion in the form of relativistic multiculturalism. Proponents of both psychological constructivism and social constructivism value knowledge for its utility and eschew as irrelevant speculation any notion that knowledge is actually about reality. The arguments are largely grounded in the discourse of science and science education where science is “western” science; neither universal nor about what is really real. The authors defended the notion of science as universal in a previous article. The present purpose is to offer a commonsense argument in defense of critical realism as an epistemology and the epistemically distinguished position of science (rather than privileged) within a framework of epistemological pluralism. The paper begins with a brief cultural survey of events during the thirty-year period from 1960–1990 that brought many educators to break with epistemological realism and concludes with comments on the pedagogical importance of realism. Understanding the cultural milieu of the past forty years is critical to understanding why traditional philosophical attacks on social constructivist ideas have proved impotent defenders of scientific realism.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

Similar content being viewed by others

scientific realism essay

Rivers and Fireworks: Social Constructivism in Education

scientific realism essay

Comparing radical, social and psychological constructivism in Australian higher education: a psycho-philosophical perspective

Penny Van Bergen & Mitch Parsell

Toward a New Philosophical Anthropology of Education: Fuller Considerations of Social Constructivism

Stephen Fleury & Jim Garrison

Traditionally realism refers to ontology. However, especially in education circles, realism is taken as an epistemology. Few anti-realist in the education community are ontological anti-realists––the issue is epistemology.

Our cultural survey is of necessity very brief. First, our argument is meant as an hypothesis to stimulate further study and discussion. Second, a longer treatment would be beyond the scope of the journal. Third, our focus is limited to American culture. Other countries and societies would undoubtedly tell the story differently.

It should be noted that logical positivism , in its doctrinaire form, was never a realist position. Early positivists like Carnap and Ayer rejected the idea that science aims to describe an independent reality, not because they thought it was false, but because they saw no way to confirm or disconfirm it by experience. Later (long before the 1960s), many former positivists abandoned this position in favor of a form of realism known as logical empiricism . The two positions have significant similarities but should not be confused (Salmon 2000 ).

There were other reasons for reforming science education. See Rudolph ( 2002 ) for a thorough discussion of economic and political pressures for science education reform prominent in the early Cold War period.

For an excellent discussion of the difference between the interests of science and public interest in science, see Eger ( 1989 ).

For examples of socially relevant science curriculum ideas of the period, see Baird ( 1937 ) or Zechiel ( 1937 ).

One indication that the critics failed in their efforts is that the Kromhout and Good title reappears thirteen years later in Gross et al. ( 1996 ). Indeed, in the eyes of many in science, the situation had only worsened as indicated by the two-word addition in the Gross et al title, The Flight From Science and Reason .

See < http://www.emory.edu/EDUCATION/mfp/Kuhnsnap.html > for a brief biographical sketch of Kuhn’s life and work. See also Science & Education vol.9 nos.1–2 for discussion of Kuhn’s impact on science educators.

We are not indicating a chronological order. For the most part, these were simultaneous events during the decade.

Although our focus is the United States, Kuhn’s book had more immediate impact in Great Britain during the late 1960s founding of the University of Edinburgh’s Strong Program in the sociology of science. This school of sociology was “in direct conflict with all philosophical theories that seek to distinguish logic or rationality from psychology or sociology” (Giere 1991 , p. 51). On the Continent, while no direct influence is claimed here between Kuhn’s science writings and the European literary “deconstructionists,” it is interesting to note some similar revolutionary writings. While Kuhn was revising the first edition of his magnum opus in an attempt to deal with criticisms of his myriad uses of “paradigms” in science communities, Jacque Derrida was, at about the same time, “deconstructing” literary texts in articles with titles like Ends of Man (Derrida 1969 ), The Purveyor of Truth (Derrida ( 1975 ), or his psychoanalysis of the “truth factor” ( 1975 ).

See RachelCarson.org (“a website devoted to the life and legacy of Rachel Carson”) at: http://www.rachelcarson.org/.

The education and social science literatures often overstate Kuhn’s influence in academic philosophy. As a counterbalance, consider that in Wesley Salmon's ( 1989 ) Four Decades of Scientific Explanation , Kuhn is mentioned only once in over 200 pages of meticulous historical survey.

The notion that Copernicus was an instrumentalist is an historical myth. “All of the evidence is that Copernicus was a robust realist and that it is Osiander, not Copernicus, who bears responsibility for the instrumentalism here. When Copernicus's disciple Georg Joachim Rheticus (author of the famous “Narratio Prima”) read the unsigned preface, he was furious and said that if he had positive proof that Osiander had inserted this he would personally give him such a thrashing that Osiander would never again interfere in the affairs of scientific men! Many good scientists who read further than the preface realized that Copernicus is an earnest realist: Maestlin and his famous pupil Kepler, Thomas Digges in England, etc.” (McGrew 2002 )

We quote Vico because Glasersfeld does; however, we do not necessarily agree with Glasersfeld’s interpretation of Vico’s work. For a different perspective on Vico, see Lilla ( 1993 ).

For a discussion on types of multiculturalism, see: Haack ( 1998 , Ch. 8).

It should be apparent that epistemological realism and ontological realism go hand in hand.

The inability to have direct access to reality is a key supposition for anti-realists. For an incisive rebuttal and defense of the theory of direct perception, see Nola ( 2003 ).

Along with the sociology of science, critical realism agrees that constructing goes on in science––that science is not about discovering “already categorized objects and relations.” The difference comes, however, in that scientists can legitimately claim “genuine similarities” between logical constructs and aspects of reality. Rather than “critical,” Giere ( 1999 ) refers to “perspectival” realism to emphasize that scientific theories never capture completely the “totality of reality” but provide us with only—perspectives “…science that is perspectival rather than absolute” (Giere 1999 , p. 79). Our use of “critical realism” is in this vein. For a philosophical introduction to critical realism, see Bhaskar ( 1989 ), Harré ( 1975 ), Putman ( 1987 ) or Salmon ( 1989 ). There are different varieties of critical realism such as Giere’s ( 1999 ) “constructive realism” but what they have in common is nicely described by Polkinghorne ( 1991 , p. 304): “epistemology models ontology.”

For more on Traditional Ecological Knowledge, see Snively and Corsiglia ( 2001 ).

Adams III HH (1986) African and African-American contributions to science and technology. In the Portland African-American Baseline Essays. Portland Public Schools, Portland, OR

Adas M (1989) Machines as the measure of man: Science, technology, and ideologies of western dominance. Cornell University Press, Ithaca, NY

Google Scholar  

Adler MJ (1974) Little errors in the beginning. The Thomist 38:27–48

Alexander D (2001) Science in search of God. Guardian Unlimited. <http://www.guardian.co.uk/Archive/Article/0,4273,4245372,00.html>

Aronowitz S (1988) Science as power: Discourse and ideology in modern society. University of Minnesota Press, Minneapolis, MN

Atwater MM (1994) Research on cultural diversity in the classroom. In: Gabel DL (ed) Handbook of research on science teaching and learning. MacMillan Publishing Company, New York, pp 558–576

Atwater MM (1995) The multicultural science classroom. The Science Teacher 62(2):21–23

Baird H (1937). Teaching the physical sciences from a functional point of view. Educational Method 16: 407–412

Bhaskar R (1989) Reclaiming reality: a critical introduction to contemporary philosophy. Verso, London

Bell D (1976) The cultural contradictions of capitalism. Basic Books, New York

Boswell J (1998) Life of Johnson (abridged and edited) -project Gutenberg’s etext of life of Johnson by [James] Boswell, Release #1564. In: CG Osgood (ed) Boswell’s Life of Johnson. P. O. Box 2782, Champaign, IL 61825: Project Gutenberg. http://digital.library.upenn.edu/webbin/gutbook/lookup?num=1564

Boyd RN (1983) On the current status of scientific realism. Erkenntnis 19:45–90

Article   Google Scholar  

Brickhouse NW (2001) Embodying science: A feminist perspective on learning. J Res Sci Teach 38(3):282–295

Bruner JS (1960) The process of education. Vintage Books, New York

Bunge M (1996) In praise of intolerance to charlatanism in academia. In: Gross PR, Levitt N, Lewis MW (eds) The flight from science and reason. New York Academy of Sciences, New York, pp 96–115

Bush V (1945) Science, the endless frontier; A report to the President. Government Printing Office, Washington, DC

Byrds (Musical group) (Composer) (1965) Greatest hits. Columbia, New York

Capra F (1977) The tao of physics: reflections on the cosmic dance. Saturday Rev 5:21–8

Carson R (1962) Silent spring. Fawcett, Greenwich, CN

Cobern WW (1997) Public understanding of science as seen by the scientific community: Do we need to re-conceptualize the challenge and to re-examine our own assumptions? In: Sjøberg S, Kallerud E (eds) Science, technology and citizenship: The public understanding of science and technology in science education and research policy. Norwegian Institute for Studies in Research and Higher Education, Oslo, Norway, pp 51–74

Cobern WW, Loving CC (2001) Defining ‘science’ in a multicultural world: Implications science education. Scie Edu 85(1):50–67

Collins H (1981) Stages in the empirical programme of relativism - introduction. Soc Stud Sci 11(1):3–10

Crouch S (1995) Melting pot blues. Am Enterp 6(2):51–55

Dawkins R (1986) The blind watchmaker: why the evidence of evolution reveals a universe without design. W.W. Norton & Company, New York

DeBoer GE (1991) A history of ideas in science education: Implications for practice. Teachers College Press, New York

Dedijer S (1962) Measuring the growth of science. Science 138(3542):781–788

Denzin NK, Lincoln YS (1994) Introduction: entering the field of qualitative research. In: Denzin NK, Lincoln YS (eds) Handbook of qualitative research. SAGE Publications, Thousand Oaks, CA, pp 1–17

Derrida J (1969) Ends of man. Philos Phenomenol Res 30(1):31–57

Derrida J (1975) The purveyor of truth. Yale Fr Stud 52:31–113

Derrida J (1975) The truth factor + separating the essence from the pretext in the psychoanalytic interpretation of literary texts, particularly the works of Edgar Allan Poe. Poetique 21:96–147

Donovan AL, Laudan L, Laudan R (1988) Scrutinizing science: empirical studies of scientific change. Johns Hopkins University Press, Baltimore

Duschl RA (1985) Science education and the philosophy of science: Twenty-five years of mutually exclusive development. Sch Sci Math 85(7):541–555

Eflin JT, Glennan S, Reisch G (1999) The nature of science: A perspective from the philosophy of science. J Res Sci Teach 36(1):107–116

Eger M (1989) The ‘interests’ of science and the problems of education. Synthese 81(1):81–106

Elkana Y (1970/2000) Science, philosophy of science and science teaching. Sci Edu 9(5):463–485

Fox-Genovese E (1999) Ideologies and realities. Orbis 43(4):531–539. WilsonSelectPlus_FT

Friberg SR (2000) Science and Religion: Paradigm Shifts, Silent Springs, and Eastern Wisdom. World Order 31(3):49–52

Garrison JW, Bentley ML (1990) Teaching scientific method: The logic of confirmation and falsification. Sch Sci Math 90(3):188–197

Geelan DR (1997) Epistemological Anarchy and the Many Forms of Constructivism. Sci Edu 6(1–2):15–28

Gergen K (1988) Feminist critique of science and the challenge of social epistemology. In: Gergen MM (ed) Feminist thought and the structure of knowledge. New York University Press, New York, pp 27–48

Giere RN (1988) Explaining Science: A Cognitive Approach. Chicago: University of Chicago Press

Giere RN (1991) Understanding scientific reasoning. New York: Holt, Rinehart and Winston

Giere RN (1999) Science without laws. The University of Chicago Press, Chicago

Glasersfeld EV (1988) Constructivism as a scientific method. Sci Reason Res Inst News Lett 3(2):8–9

Glasersfeld EV (1989) Cognition, construction of knowledge, and teaching. Synthese 80(1):121–140

Glasersfeld EV (2001) Learning as constructive activity. Paper presented at the annual meeting of the proceedings of the 5th annual meeting of the North American group of PME (1983). UMass Scientific Reasoning Research Institute, Amherst, MA

Glazer N, Moynihan DP (1979) Beyond the melting pot; the Negroes, Puerto Ricans, Jews, Italians, and Irish of New York City, 2nd edn. The M.I.T. Press, Cambridge, Massachusetts

Gollnick DM, Chinn PC (1986) Multicultural education in a pluralistic society. Merrill, Columbus, OH

Good RG (1991) Theoretical bases for science education research: Contextual realism in science and science education. Paper presented at the annual meeting of the National Association for Research in Science Teaching. The Abbey, Fontane, WI

Gross PR, Levitt N, Lewis MW (1996) The flight from science and reason. New York Academy of Sciences, New York

Guba EG, Lincoln YS (1994) Competing paradigms in qualitative research. In: Denzin NK, Lincoln YS (eds) Handbook of qualitative research. SAGE Publications, Thousand Oaks, CA, pp 105–117

Haack S (1998) Manifesto of a passionate moderate. The University of Chicago Press, Chicago, IL

Hanson NR (1958) Patterns of discovery: an inquiry into the conceptual foundations of science. Cambridge University Press, Cambridge, UK

Harré R (1975) Causal powers: a theory of natural necessity. Basil Blackwell, Oxford

Hofstein A, Yager RE (1982) Societal issues as organizers for science education in the ‘80s. Sch Sci Math 82(7):539–547

Holton G (2000) The Rise of Postmodernisms and the “End of Science”. J Hist Ideas 61(2):327–341

Hurd PD (1998) Scientific literacy: New minds for a changing world. Sci Edu 82(3):407–416

Hurd PD (2000) Science education for the 21 st century. Sch Sci Math 100(6):282–238

Johnson S (1969/1752) Elementa philosophica: containing chiefly Noetica, or things relating to the mind or understanding; and Ethica, or things relating to the moral behaviour. Kraus Reprint Co, New York

Khlentzos D (2000) Semantic challenges to realism. Stanford Encyclopedia of Philosophy

Kromhout R, Good R (1983) Beware of societal issues as organizers for science education. Sch Sci Math 83(8):647–650

Kuhn TS (1962/1970) The structure of scientific revolutions. University of Chicago Press, Chicago

Ladriere J (1977) The challenge presented to cultures by science and technology. UNESCO, Paris, France

Laudan L (1990) Science and relativism: some key controversies in the philosophy of science. University of Chicago Press, Chicago

Lear L (1998) Rachel Louise Carson [Web Page]. URL http://www.rachelcarson.org/biography.htm [2001, July 19]

Lilla M (1993) G. B. Vico: The antimodernist. The Wilson Quarterly XVII(3):32–39

Lillegard N (2001) Meaning, language, rules, social construction under philosophy of social science (Internet Encyclopedia of Philosophy) [Web Page]. URL http://www.utm.edu/research/iep/s/socscien.htm#Meaning, Language, Rules, Social Construction

Loving CC (1991) The scientific theory profile: A philosophy of science models for science teachers. Journal of Research in Science Teaching 28(9): 823–838

Loving CC (1997) From the summit of truth to its slippery slopes: science education’s journey through positivist-postmodern territory. Am Edu Res J 34(3):421–452

Loving CC, Cobern WW (2000) Invoking Thomas Kuhn: What citation analysis reveals for science education. Sci Edu 9(1/2):187–206

Loving CC, Ortiz de Montellano B (2000) Classical and reform curriculum analysis: How do multicultural science curricula fare? Can culturally relevant science be taught? Paper presented at the annual meeting of the American Educational Research Association, New Orleans, LA

Luft J (1998) Multicultural science education: An overview. J Sci Teach Edu 9(2):103–122

Lyons S (2001) Health care fraud: Medical ‘quackery’ booming in America. American Health Line, pp 1–2

Mahner M, Bunge M (1996) Is religious education compatible with science education? Sci Edu 5(2):101–123

Mallinson GG (1984) Leastwise, not much. Sch Sci Math 84(1):1–6

Matthews MR (1994) Science teaching: The role of history and philosophy of science. New York, Routledge

Matthiessen P (Environmentalist: Rachel Carson [Web Page]. URL http://www.time.com/time/time100/scientist/profile/carson.html [2001, July 19]

McDowell SA, Ray BD (issue ed) (2000) Home schooling. Peabody J Edu 75(1&2)

McGuire B (Composer) (1960–1969) Eve of destruction. Beverly Hills, CA, Dunhill

McGrew T (2002) Personal communication

Nadeau R, Desautels J (1984) Epistemology and the teaching of science. Science Council of Canada, Ottawa, Canada

Nola R (1997) Constructivism in science and in science education: a philosophical critique, Science & education 6(1–2):55–83. Reproduced in M.R. Matthews (ed), Constructivism in science education: A philosophical debate, Kluwer Academic Publishers, Dordrecht, 1998

Nola R (2003) ‘Naked before reality; Skinless before the absolute’: A Critique of the inaccessibility of reality argument in constructivism. Sci Edu 12(2):131–166

O’Neil, J. (1991) On the Portland Plan: a conversation with Matthew Prophet. Educational Leadership, 24–27

Ortiz de Montellano B (1996) Afrocentric pseudoscience: The miseducation of African Americans. In: Gross PR, Levitt N, Lewis MW (eds) The flight from science and reason. The New York Academy of Sciences, NY, pp 561–572

Peirce CS (1931) Collected papers of Charles Sanders Peirce. Harvard University Press, Cambridge, MA

Phillips DC, Burbules NC (2000) Postpositivism and educational research. Rowman & Littlefield Publishers, Lanham, MD

Polkinghorne JC (1991) God’s action in the world. Cross Currents 41(3):293–307

Powell CF (1972) The aims of science in our time. In: Burhop EHS, Lock WO, Memon MGK (eds) Collected papers of Cecil Frank Powell. American Elsevier Publishing Co, New York, NY, pp 434–446

Prather JP (1990) Tracing science teaching. National Science Teachers Association, Washington, DC

Putman H (1987) The many faces of realism. Open Court Press, LaSalle, IL

Randall JH Jr (1940) The making of the modern mind. Columbia University Press, New York

Reid T (1764/1997) An inquiry into the human mind: on the principles of common sense. University Park, PA, Pennsylvania State University Press

Reiff D (1993) Multiculturalism’s silent partner - It’s the newly globalized economy, stupid. Harper’s 287(1719):62–72

Rosenblatt B (2001 May) Say ‘Aaah’: Alternative-treatment trend raises insurance issues. Los Angeles Times, p. 2

Rostow WW (1971) The stages of economic growth: a non-communist manifesto. Cambridge University Press, Cambridge, UK

Roth W-M., Roychdhury A (1994) Physics students epistemologies and views about knowing and learning. J Res Sci Teach 31(1):5–30

Rudolph JL (2002) Scientists in the classroom: the cold war reconstruction of American science education. Palgave, New York

Salmon WC (1989) Four decades of scientific explanation. In: W. Salmon, Kitcher (eds), Scientific explanation. University of Minnesota Press. Notes, Minneapolis: Call Number: Q174.8 .S26 1989

Salmon WC (2000) Logical empiricism. In: Newton-Smith WH (ed) A companion to the philosophy of science. Blackwell Publishers, Malden, MA, pp 233–242

Sankey H (2001) Scientific realism: An elaboration and a defense. PhilSci Archive. <http://philsci-archive.pitt.edu/documents/disk0/00/00/03/04/index.html>

Schwab JJ (1962) The teaching of science as enquiry. The Inglis Lectureship. Harvard University Press, Cambridge, MA

Shanker A (1992) The dangers of a multicultural curriculum. The World & I 7(4):110–115

Siegel H (2001) Incommensurability, rationality and relativism: in science, culture and science education. Boston Stud Phil Sci 216:207–224

Sleeter CE, Grant CA (1987) An analysis of multicultural education in the United States. Harvard Edu Rev 57(4):421–444

Slezak P (1994) Sociology of scientific knowledge and scientific education: Part I. Sci Edu 3(3):265–294

Slezak P (1994) Sociology of scientific knowledge and science education. Part II: laboratory life under the microscope. Sci Edu 3(4):329–355

Snively G, Corsiglia J (2001) Discovering indigenous science: Implications for science education. Sci Edu 85(1):6–34

Snow CP (1964) Two cultures and the scientific revolution. Cambridge University Press, Cambridge, UK

Sokal A, Bricmont J (1998) Fashionable nonsense: postmodern philosophers’ abuse of Science. Picador, New York

Southerland SA (2000) Epistemic universalism and the shortcomings of curricular multicultural science education. Sci Edu 9(3):298–307

Staver JR (1998) Constructivism: Sound theory for explicating the practice of science and science teaching. J Res Sci Teach 35(5):501–520

Strauss A, Corbin J (1994) Grounded theory methodology. In: Denzin NK, Lincoln YS (eds) Handbook of qualitative research. SAGE Publications, Thousand Oaks, CA, pp 273–285

Suchting WA (1992) Constructivism deconstructed. Sci Edu 1(3):223–254

Suchting WA (1995) The nature of scientific thought. Sci Edu 4(1):1–22

Tempels P (1959) Bantu philosophy. C King (Translator) Presence Africaine, Paris, FR

TIME (2001 April) Alternative medicine: TIME profiles leaders in six fields. American Health Line, pp 1–2

Toulmin S (1960) The philosophy of science: an introduction. Harper & Row, New York

Yaffe G (2000) Thomas Reid. Stanford Encyclopedia of Philosophy. <http://plato.stanford.edu/entries/reid/>

Yager RE (1983) In defense of societal issues as organizers for school science. Sch Sci Math 83(8):651–653

Zacharias J (1959) Education for the age of science. President’s Science Advisory Committee, Washington, DC

Zechiel AN (1937) Recent trends in revision of science curricula. Educational Method XVI: 402–407

Download references

Author information

Authors and affiliations.

The Mallinson Institute for Science Education, Western Michigan University, Kalamazoo, MI, 49008, USA

William W. Cobern

Department of Teaching, Learning & Culture, Texas A&M University, College Station, Texas, 77843, USA

Cathleen C. Loving

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to William W. Cobern .

Rights and permissions

Reprints and permissions

About this article

Cobern, W.W., Loving, C.C. An Essay for Educators: Epistemological Realism Really is Common Sense. Sci & Educ 17 , 425–447 (2008). https://doi.org/10.1007/s11191-007-9095-5

Download citation

Received : 15 September 2006

Accepted : 26 May 2007

Published : 15 June 2007

Issue Date : April 2008

DOI : https://doi.org/10.1007/s11191-007-9095-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Science education
  • Multiculturalism
  • Cultural studies
  • Commonsense
  • Find a journal
  • Publish with us
  • Track your research

Your Article Library

Essay on realism.

scientific realism essay

ADVERTISEMENTS:

Read this essay to learn about Realism. After reading this Essay you will learn about: 1. Introduction to Realism 2. Fundamental Philosophical Ideas of Realism 3. Forms 4. Realism in Education 5. Curriculum 6. Evaluation

  • Essay on the Evaluation

Essay # 1. Introduction to Realism:

Emerged as a strong movement against extreme idealistic view of the world around.

Realism changed the contour of education in a systematic way. It viewed external world as a real world; not a world of fantasy.

It is not based upon perception of the individuals but is an objective reality based on reason and science.

The Realist trend in philosophical spectrum can be traced back to Aristotle who was interested in particular facts of life as against Plato who was interested in abstractions and generalities. Therefore, Aristotle is rightly called as the father of Realism. Saint Thomas Aquinas and Comenius infused realistic spirit in religion.

John Locke, Immanuel Kant, John Freiderich Herbart and William James affirmed that external world is a real world. In the 20th century, two sections of realist surfaced the area of philosophy. Six American professors led by Barton Perry and Montague are neo-realists. Another section spearheaded by Arthur Lovejoy, Johns Hopkins and George Santayana emerged are called as critical realists.

Essay # 2. Fundamental Philosophical Ideas of Realism :

(i) phenomenal world is true:.

Realists believe in the external world which is true as against the idealist world-a world d this life. It is a world of objects and not ideas. It is a pluralistic world. Ross has commented, “Realism simply affirms the existence of an external world and is therefore the antithesis of subjective idealism.”

There is an order and design of the external world in which man is a part and the world idealism by the laws of cause and effect relationships. As such there is no freedom of the will for man.

(ii) Opposes to Idealist Values :

In realism, there is no berth for imagination and speculation. Entities of God, soul and other world are nothing; they are mere figments of human imagination. Only objective world is real world which a man can know with the help of his mind. Realism does not believe in ideal values, would discover values in his immediate social life. The external world would provide the work for the discovery and realization of values.

(iii) Theory of Organism :

Realists believe that an organism is formed by conscious and unconscious things. Mind is regarded as the function of organism. Whitehead, a Neo-realist remarks “ The universe is a vibrating organism in the process of evolution. Change is the fundamental feature of this vibrating universe. The very essence of real actuality is process. Mind must be regarded as the function of the organism.”

(iv) Theory of Knowledge :

According to realists, the world around us is a reality; the real knowledge is the knowledge of the surrounding world. Senses are the gateways of knowledge of the external world. The impressions and sensations as a result of our communication with external world through our sense organs result in knowledge which is real.

The best method to acquire the knowledge of the external world is the experiment or the scientific method. One has to define the problem, observe all the facts and phenomena pertaining to the problem, formulate a hypothesis, test and verify it and accept the verified solution. Alfred North, Whitehead, and Bertrand Russel have stressed on the use of this scientific method.

(v) Stress on Present Applied Life :

According to realists, spiritual world is not real and cannot be realized. They believed in the present world-physical or material which can be realized. Man is a part and parcel of this material world. They put premium upon the molding and directing of human behaviour as conditioned by the physical and material facts of the present life, for this can promote happiness and welfare.

Therefore, metaphysics according to realism is that the external world is a reality-it is a world of objects and not ideas. Epistemology deals with the knowledge-knowledge of this external world through the senses and scientific method and enquiry. Axiology in it is that realists reject idealistic values, favour discovering values in the immediate social life.

Essay # 3. Forms of Realism :

There are four forms of realism, viz., humanistic realism, social realism, sense realism and neo-realism.

(i) Humanistic Realism :

The advocates of this form of realism are Irasmus, Rebelias and Milton. The supporters of the realism firmly believed that education should be realistic which can promote human welfare and success. They favoured the study of Greek & Roman literature for individual, social and spiritual development.

Irasmus (1446-1536) castigated narrow educational system and in its place. favoured broad and liberal education. Rebelias (1483-1553) also advocated liberal education, opposed theoretical knowledge and said that education should be such as to prepare the individual to face all the problems of life with courage and solve them successfully.

He suggested scientific and psychological methods and techniques. Milton (1608-1674) also stressed liberal and complete education. He, in this connection, writes, “I call therefore a complete and generous education that which fits a man to perform justly, skillfully and magnanimously all the offices both private and public of peace and war.”

He opposed mere academic education and insisted that education should give knowledge of things and objects. He prescribed language, literature and moral education is main subjects of study; and physiology, agriculture and sculpture as subsidiary subjects of study for children.

(ii) Social Realism :

Social realists opposed academic and bookish knowledge and advocated that education should promote working efficiency of men and women in the society. Education aims at making human life happier and successful. They suggested that curriculum should include History, Geography. Law, Diplomacy, Warfare, Arithmetic’s, Dancing, Gymnastics etc. for the development of social qualities.

Further, with a view to making education practical and useful, the realists stressed upon Travelling, Tour, observation and direct experience. Lord Montaigne (1533-1552) condemned cramming and favored learning by experience through tours and travels. He opposed knowledge for the sake of knowledge and strongly advocated practical and useful knowledge.

John Locke (1635-1704) advocated education through the mother tongue and lively method of teaching which stimulates motivation and interest in the children. As an individualist, he believed that the mind of a child is a clean slate on which only experiences write. He prescribed those subjects which are individually and socially useful in the curriculum.

(iii) Sense Realism :

Developed in the Seventeenth century sense realism upholds the truth that real knowledge comes through our senses. Further, sense realists believed all forms of knowledge spring from the external world. They viewed that education should provide plethora of opportunities to the children to observe and study natural phenomena and come in contact with external objects through the senses.

Therefore, true knowledge is gained by the child about natural objects, natural phenomena and laws through the exercises of senses. They favoured observation, scientific subjects, inductive method and useful education. Mulcaster (1530—1611) advocated physical and mental development aims of education.

Reacted against any forced impressions upon the mind of the child, he upheld use of psychological methods of teaching for the promotion of mental faculties-intelligence, memory and judgement.

Francis Bacon (1562-1623) writes, “The object of all knowledge is to give man power over nature.” He, thus, advocated inductive method of teaching-the child is free to observe and experiment by means of his senses and limbs. He emphasised science and observation of nature as the real methods to gain knowledge.

Ratke (1571-1625) said that senses are the gateways of knowledge and advocated the following maxims:

a. One thing at a time,

b. Follow nature,

c. Repetition,

d. Importance on mother-tongue,

e. No rote learning,

f. Sensory knowledge,

g. Knowledge through experience and uniformity of all things.

Comenius (1592-1671) advocated universal education and natural method of education. He said that knowledge comes not only through the senses but through man’s intelligence and divine inspiration. He favoured continuous teaching till learning is achieved and advocated mother-tongue to precede other subjects.

(iv) Neo-Realism :

The positive contribution of neo-realism is its acceptance of the methods and results of modern development in physics. It believes that rules and procedures of science are changeable from time to time according to the conditions of prevailing circumstances.

Whitehead said that an organism is formed by the consciousness and the unconsciousness, the moveable and immovable thing. Education should give to child full-scale knowledge of an organism. Man should understand all values very clearly for getting full knowledge about organism. Bertrand Russell emphasized sensory development of the child.

He favoured analytical method and classification. He assigned no place to religion and supported physics to be included as one of the foremost subjects of study. Further, he opposed emotional strain in children as it leads to development of fatigue.

Essay # 4. Realism in Education :

Realism asserts that education is a preparation for life, for education equips the child by providing adequate training to face the crude realities of life with courage as he or she would perform various roles such as a citizen, a worker, a husband, a housewife, a member of the group, etc. As such, education concerns with problems of life of the child.

Chief Characteristics of Education :

The following are the chief characteristics of realistic education:

(i) Based on Science:

Realism emphasized scientific education. It favored the inclusion of scientific subjects in he curriculum and of natural education. Natural education is based on science which is real.

(ii) Thrust upon present Life of the Child :

The focal point of realistic education is the present life of the child. As it focuses upon the real and practical problems of the life, it aims at welfare and happiness of the child.

(iii) Emphasis on Experiment and Applied life :

It emphasizes experiments, experience and practical knowledge. Realistic education supports learning by doing and practical work for enabling the child to solve his or her immediate practical problems for leading a happy and successful life.

(iv) Opposes to Bookish Knowledge :

Realistic education strongly condemned all bookish knowledge, for it does not help the child to face the realities of life adequately. It does not enable the child to decipher the realities of external things and natural phenomena. The motto of realistic education is ”Not Words but Things.”

(v) Freedom of Child:

According to realists, child should be given full freedom to develop his self according to his innate tendencies. Further, they view that such freedom should promote self-discipline and self-control the foundation of self development.

(vi) Emphasis on Training of Senses:

Unlike idealists who impose knowledge from above, realists advocated self-learning through senses which need to be trained. Since, senses are the doors of knowledge, these needs to be adequately nurtured and trained.

(vii) Balance between Individuality and Sociability :

Realists give importance to individuality and sociability of the child equally. Bacon lucidly states that realistic education develops the individual on the one hand and tries to develop social trails on the other through the development of social consciousness and sense of service of the individual.

Aims of Education :

The following aims of education are articulated by the realists:

(i) Preparation for the Good life:

The chief aim of realistic education is to prepare the child to lead a happy and good life. Education enables the child to solve his problems of life adequately and successfully. Leading ‘good life’ takes four important things-self-preservation, self-determination, self-realization and self-integration.

(ii) Preparation for a Real Life of the Material World:

Realists believe that the external material world is the real world which one must know through the senses. The aim of education is to prepare a child for real life of material world.

(iii) Development of Physical and Mental Powers:

According to realists, another important aim of education is to enable the child to solve different life problems by using the faculty of mind: intelligence, discrimination and judgement.

(iv) Development of Senses:

Realists thought that development of senses is the sine-qua-non for realization of the material world. Therefore, the aim of education is to help the development of senses fully by providing varied experiences.

(v) Acquainting with External Nature and Social Environment:

It is an another aim of realistic education to help the child to know the nature and social environment for leading a successful life.

(vi) Imparting Vocational Knowledge and Skill :

According to realists, another important aim of education is to provide vocational knowledge, information, skill etc., to make the child vocationally efficient for meeting the problems of livelihood.

(vii) Development of Character :

Realistic education aims at development of character for leading a successful and balanced life.

(viii) Enabling the Child to Adjust with the Environment :

According to realists, education should aim at enabling the child to adapt adequately to the surroundings.

Essay # 5. Curriculum of Realism :

Realists wanted to include those subjects and activities which would prepare the children for actual day to day living. As such, they thought it proper to give primary place to nature, science and vocational subjects whereas secondary place to Arts, literature, biography, philosophy, psychology and morality.

Besides, they have laid stress upon teaching of mother- tongue as the foundation of all development. It is necessary for reading, writing and social interaction but not for literary purposes.

(i) Methods of Teaching :

Realists favoured principles of observation and experience as imparting knowledge of objects and external world can be given properly through the technique of observation and experience. Further, they encouraged use of audio-visual aids in education as they would develop sensory powers in the children.

Children would have “feel” of reality through them. Realists also encouraged the use of lectures, discussions and symposia. Socratic and inductive methods were also advocated. Memorization at early stage was also recommended.

Besides, learning by travelling was also suggested. The maxims of teaching are to proceed from easy to difficult, simple to complex, known to unknown, definite to indefinite, concrete to abstract and particular to general. In addition, realists give importance on the principle of correlation as they consider all knowledge as one unit.

(ii) Discipline :

Realists decry expressionistic discipline and advocate self-discipline to make good adjustment in the external environment. They, further, assert that virtues can be inculcated for withstanding realities of physical world. Children need to be disciplined to become a part of the world around in and to understand reality.

(iii) Teacher :

Under the realistic school, the teacher must be a scholar and his duty is to guide the children towards the hard core realities of life. He must expose them to the problems of life and the world around. The teacher should have full knowledge of the content and needs of the children.

He should present the content in a lucid and intelligible way by employing scientific and psychological methods is also the duty of the teacher to tell children about scientific discoveries, researches and inventions id he should inspire them to undertake close observation and experimentation for finding out new facts and principles.

Moreover, he himself should engage in research activities. Teachers, in order to be good and effective, should get training before making a foray into the field of teaching profession.

(iv) School :

Some realists’ view that school is essential as it looks like a mirror of society reflecting its real picture of state of affairs. It is the school which provides for the fullest development of the child in accordance with his needs and aspirations and it prepares the child for livelihood. According to Comenius, “The school should be like the lap of mother full of affection, love and sympathy. Schools are true foregoing places of men.”

Essay # 6. Evaluation of Realism :

Proper evaluation of realism can be made possible by throwing a light on its merits and demerits.

(i) Realism is a practical philosophy preaching one to come to term with reality. Education which is non-realistic cannot be useful to the humanity. Now, useless education has come to be considered as waste of time, energy and resources.

(ii) Scientific subjects have come to stay in our present curriculum due to the impact of realistic education.

(iii) In the domain of methods of teaching the impact of realistic education is ostensible. In modern education, inductive, heuristic, objective, experimentation and correlation methods have been fully acknowledged all over the globe.

(iv) In the area of discipline, realism is worth its name as it favours impressionistic and self-discipline which have been given emphasis in modern educational theory and practice in a number of countries in the globe.

(v) Realistic philosophy has changed the organisational climate of schools. Now, schools have been the centres of joyful activities, practical engagements and interesting experiments. Modern school is a vibrant school.

(i) Realism puts emphasis on facts and realities of life. It neglects ideals and values of life. Critics argue that denial of ideals and values often foments helplessness and pessimism which mar the growth and development of the individuals. This is really lop-sided philosophy.

(ii) Realism emphasizes scientific subjects at the cost of arts and literature. This affair also creates a state of imbalance in the curriculum. It hijacks ‘humanities’ as critics’ label.

(iii) Realism regards senses as the gateways of knowledge. But the question comes to us, how does illusion occur and how do we get faulty knowledge? It does not provide satisfactory answer.

(iv) Realism accepts the real needs and feelings of individual. It does not believe in imagination, emotion and sentiment which are parts and parcel of individual life.

(v) Although realism stresses upon physical world, it fails to provide answers to the following questions pertaining to physical world.

(i) Is the physical world absolute ?

(ii) Is there any limits of physical world ?

(iii) Is the physical world supreme or powerful?

(vi) Realism is often criticized for its undue emphasis on knowledge and it neglects the child. As the modern trend in education is paedocentric, realism is said to have put the clock behind the times by placing its supreme priority on knowledge.

In-spite of the criticisms, realism as a real philosophy stands to the tune of time and it permeates all aspects of education. It is recognized as one of the best philosophies which need to be browsed cautiously. It has its influence in modern educational theory and practice.

Related Articles:

  • Influence of Sense-Realism on Education | J. A. Comenius
  • Importance of Realism in Geography

Comments are closed.

web statistics

Advertisement

Supported by

A Novelist Comes Home to Bury Her Words, and Brings Them Back to Life

In Julia Alvarez’s “The Cemetery of Untold Stories,” a boneyard in the Dominican Republic becomes a rich wellspring for discarded narratives.

  • Share full article

scientific realism essay

By Luis Alberto Urrea

Luis Alberto Urrea’s most recent novel is “Good Night, Irene.”

THE CEMETERY OF UNTOLD STORIES , by Julia Alvarez

The best stories live within us long after the final word; characters and places continue beyond the lines on a page. Yet the image of all the unfinished, unsatisfying, impossible stories we leave in our wakes haunts the writer as well as the reader. And with more than 20 books published across a three-decade career, no one may be haunted more than Julia Alvarez.

The hero of Alvarez’s seventh novel, Alma Cruz, is a writer from the Dominican Republic who has come to the United States and created a literary life, beginning with critically acclaimed books about the motherland and evolving into a chronicler of life in the U.S.A. (Longtime readers of Alvarez’s work will recognize her own trajectory, from her early classics like “How the García Girls Lost Their Accents” and “In the Time of the Butterflies” through poetry, memoir, children’s books and more.)

The famous author has always wrestled shadow and sunlight, laughter and agony, into tales that sometimes felt like ghost stories. Readers knew to seek the truths behind the narrative — to find sorrow in the funniest scenes, or the unexpected outburst of joy in a somber one. Of course, I am speaking about Alma Cruz. (But also, Julia Alvarez.)

One day, Alma decides she has had enough with the fame game, the big career and its ups and downs. She comes to a lovely conclusion: It is time to return to the homeland she fled, and she will take all the drafts of her unfinished or unpublished books and lay them to rest there, giving each a proper burial. She buys a plot of land and begins to build a graveyard.

The locals become a fantastic choir of curious, suspicious, baffled neighbors: One rumor has it that “the place will be a resort, which would provide employment for maids, gardeners, waiters, cooks, watchmen”; another imagines “a grand house, complete with a swimming pool, a tennis court, a mini putting green.” Still another posits, “A baseball academy would be a dream come true for the tigueritos roaming the streets. Keep them out of trouble.” But when they realize Alma is building a graveyard, the outburst is comedic: a cemetery! Fear of zombies immediately clashes with the fear of homeless people defecating in mausoleums — and what kinds of jobs, by the way, are there in a boneyard? They have more reckoning to do when they realize Alma intends to put her stories in the ground, literally.

Word goes out that the great author has returned, and the locals flock to her like butterflies, everyone eager to share. A festival of storytelling breaks out in the tropics. Are the neighbors hoping to bury their own or are they giving life to tales untold? No matter. Rumors and gossip, histories and familial dramas swirl around Alma. “A little bird told me. Había una vez. Cuentan los viejos. Some scandal on the news, who is sleeping with whom, what fulano has done or said to fulana, a juicy chisme, a hot rumor. …” Amid the chatter, Alma’s own stories, the ones she has come home to bury, somehow find their place.

Soon Alma is meeting with architects, and more characters join in the cumbia of story — the dueling Perla and Filomena, who have not spoken for 30 years but keep each other’s phone numbers just in case. It is a shadowy feud, of course: “Way back, Filomena destroyed Perla’s peace of mind. The story has been buried so deep, it should have rotted into oblivion. But like Lazarus in la Biblia, it keeps coming back to life.”

Indeed. Here comes the fabulous Bienvenida (it means “Welcome”), with her tragic history that Alma cannot resist. She was once, you see, the wife of the dictator, Trujillo, a loyal and devoted first lady who is cast aside when she is unable to produce an heir. Eventually, El Jefe falls for the charms of another woman — “jealous and possessive, with a will equal to his own” — and Bienvenida’s “death knell comes when this mistress gives birth to a son.”

Men and boys, too, join in with their dramas and secrets, pride and regrets. As voices and stories are set free, it feels like a carnival, a festival. Alma’s first-person voice is jostled. We do not care; we are already in the warm sun and the sea wind and the cooking smells and the music of the dance.

As the book accelerates, the characters seem to become their own novelists. They rewrite their lives, they revise their histories, they reinvent their ongoing myths even as Alma is planning to bury her own stories in their troubled, sacred earth. Only an alchemist as wise and sure as Alvarez could swirl the elements of folklore and the flavor of magical realism around her modern prose and make it all sing.

The camino that “The Cemetery of Untold Stories” travels — from Vermont to the Dominican Republic, from literary fame to chosen retreat, from modern American writing to a profoundly Latin American tone — is lively, joyous, full of modern details and old tall tales. Any reader with roots and ancestors in other lands lives in a multiple-narrative story, one that we try to share with everyone, though we have to translate it. Yet we also go back to the ancestral home, and find ourselves translating our Yanqui life as well. Which story is the truest?

This often witty, occasionally somber and elegiac novel begins with a simple exhortation, in English: “Tell me a story.” It ends on a melancholy and evocative note. Spoiler alert: Another single line, this time in Spanish after the last page concludes, announces, “Este cuento se ha acabado.” (This story has ended.) A definitive slam of the door.

THE CEMETERY OF UNTOLD STORIES | By Julia Alvarez | Algonquin | 256 pp. | $28

Explore More in Books

Want to know about the best books to read and the latest news start here..

James McBride’s novel sold a million copies, and he isn’t sure how he feels about that, as he considers the critical and commercial success  of “The Heaven & Earth Grocery Store.”

How did gender become a scary word? Judith Butler, the theorist who got us talking about the subject , has answers.

You never know what’s going to go wrong in these graphic novels, where Circus tigers, giant spiders, shifting borders and motherhood all threaten to end life as we know it .

When the author Tommy Orange received an impassioned email from a teacher in the Bronx, he dropped everything to visit the students  who inspired it.

Do you want to be a better reader?   Here’s some helpful advice to show you how to get the most out of your literary endeavor .

Each week, top authors and critics join the Book Review’s podcast to talk about the latest news in the literary world. Listen here .

IMAGES

  1. A Critical Introduction to Scientific Realism: : Bloomsbury Critical

    scientific realism essay

  2. (PDF) Evidence, Explanation, and the Empirical Status of Scientific

    scientific realism essay

  3. (PDF) Towards a Practical Realist Philosophy of Science

    scientific realism essay

  4. New Approaches to Scientific Realism

    scientific realism essay

  5. Chapter 11 On the Current Status of Scientific Realism

    scientific realism essay

  6. 1 Scientific Realism

    scientific realism essay

VIDEO

  1. Interstellar BTS: Nolan's focus on realism and scientific accuracy 🌌

  2. Intro to realism essay

  3. Science & Truth

  4. Pragmatic Realism: Philosophy, History & Science

  5. Scientific or Critical Realism by John Locke

  6. Scientific Realism

COMMENTS

  1. Scientific Realism

    Debates about scientific realism are closely connected to almost everything else in the philosophy of science, for they concern the very nature of scientific knowledge. ... Essays on Realism and Explanation, London: Palgrave Macmillan. Putnam, Hilary, 1975a, Mathematics, Matter and Method, Cambridge: Cambridge University Press.

  2. Scientific Realism

    This is a file in the archives of the Stanford Encyclopedia of Philosophy. Scientific Realism. First published Wed Jun 12, 2002. It is easier to define scientific realism than it is to identify its role as a distinctly philosophical doctrine. Scientific realists hold that the characteristic product of successful scientific research is knowledge ...

  3. (PDF) What is Scientific Realism?

    When the term appears, as it does quite often, cursorily in an introduction. or comment, it is easy to gain the impression that a scienti c realist is a true. believer, someone who holds currently ...

  4. Scientific Realism and Antirealism

    5. Scientific Realism. In the 1970s, a particularly strong form of scientific realism (SR) was advocated by Putnam, Boyd, and others (Boyd 1973, 1983; Putnam 1962, 1975a, 1975b). When scientific realism is mentioned in the literature, usually some version of SR is intended. SR is often characterized in terms of two commitments (van Fraassen 1980):

  5. Scientific Realism

    Scientific realism is committed to the unobservable world enjoying this independence. We might call this the 'independence dimension' of realism. The very influential philosophers of science Thomas Kuhn and Paul Feyerabend think that scientific entities are not independent but are somehow 'constructed' by the theories we have of them ...

  6. Full article: Putting scientific realism into perspective

    Putting scientific realism into perspective. In this paper, we offer a brief overview of the debate between realism and anti-realism in the philosophy of science. On the background of that debate, we consider two recently developed approaches aimed at vindicating realist intuitions while acknowledging the limitations of scientific knowledge.

  7. (PDF) Review of Scientific Realism: Selected Essays of Mario Bunge, by

    Volume 9 of Spontaneous Generations: A Journal for the History and Philosophy of Science presents an eclectic but surprisingly harmonious collection of invited and peer-reviewed papers, organized under the title "The Future of the Scientific Realism Debate: Contemporary Issues Concerning Scientific Realism."

  8. Realism and Theory Change in Science

    The second part deals with the historical challenge to scientific realism as this is currently formulated and the various lines of defense of the claim that scientific knowledge grows despite theory-change. 1. The History of the Historical Challenge. 1.1 The Bankruptcy-of-science Debate. 1.2 Duhem on Continuity.

  9. What is Scientific Realism?

    The debate over scientific realism, simply put, is a debate over what we can and should believe about reality once we've critically assessed all the available arguments and empirical evidence. Thinking earnestly about the merits of scientific realism as a philosophical thesis requires navigating contentious historiographical issues, being ...

  10. Scientific Realism: Selected Essays of Mario Bunge

    Scientific Realism: Selected Essays of Mario Bunge. M. Bunge, M. Mahner. Published 1 August 2001. Philosophy. The prolific and indefatigable Mario Bunge has been writing books and articles on almost every facet of philosophy for nearly fifty years. At a time when much contemporary philosophy is mired in obfuscating jargon and devoid of purpose ...

  11. Scientific realism: selected essays of Mario Bunge

    Energy: Between Physics and Metaphysics -- 4. The Revival of Causality -- 5. Emergence and the Mind -- 6 SCIENTIFIC REALISM -- 6. The Status of Concepts -- 7. Popper's Unworldly World 3 --II. METHODOLOGY AND PHILOSOPHY OF SCIENCE -- 8. On Method in the Philosophy of Science -- 9. Induction in Science -- 10. The GST Challenge to the Classical ...

  12. [PDF] Scientific realism: what it is, the contemporary debate, and new

    First, I answer the controversial question 'What is scientific realism?' with extensive reference to the varied accounts of the position in the literature. Second, I provide an overview of the key developments in the debate concerning scientific realism over the past decade. Third, I provide a summary of the other contributions to this special issue.

  13. Scientific Realism

    Scientific realism maintains that by applying to science one can, first, come to know a great deal about it, Second, that this knowledge relates not just to peripheral matters but to essentials. Finally, this information is provided only through science. It holds that science provides a rational account of the salient and fundamental properties ...

  14. Evidence, Explanation, and Realism: Essays in Philosophy of Science

    The essays in this volume address three fundamental questions in the philosophy of science: What is required for some fact to be evidence for a scientific hypothesis? What does it mean to say that a scientist or a theory explains a phenomenon? Should scientific theories that postulate "unobservable" entities such as electrons be construed realistically...

  15. What is scientific realism?

    Scientific realism is therefore a sympathetic basis in the philosophy of social science for those philosophers and sociologists who are most concerned to put aside the positivist origins of both philosophy of science and sociology. Mario Bunge argues strongly that scientific realism is most suited to an appropriate methodology for the social ...

  16. Scientific realism

    Scientific realism involves two basic positions: First, it is a set of claims about the features of an ideal scientific theory; an ideal theory is the sort of theory science aims to produce. Second, it is the commitment that science will eventually produce theories very much like an ideal theory and that science has done pretty well thus far in ...

  17. Scientific Realism : Selected Essays of Mario Bunge

    Books. Scientific Realism: Selected Essays of Mario Bunge. Mario Bunge. Prometheus Books, 2001 - Philosophy - 438 pages. The prolific and indefatigable Mario Bunge has been writing books and articles on almost every facet of philosophy for nearly fifty years. At a time when much contemporary philosophy is mired in obfuscating jargon and devoid ...

  18. Scientific Realism and Chemistry

    scientific realism is a philosophical issue with relevance to all sciences, but there are some particularly interesting and distinctive ways in which it has manifested itself in chemistry. Paying proper attention to such aspects will deliver two types of benefits: First, it will aid the philosophical understanding of the nature of chemical knowledge; second, it will throw some fresh light on ...

  19. The Role of the 'No Miracles' Argument for Scientific Realism Essay

    The 'no miracles' argument also seems to be convincing as the defence of scientific realism while responding to the three dimensions of scientific realism because this philosophy involves the discussion of the argument's semantic and epistemic nature along with the focus on the possible truth of scientists' predictions.

  20. PDF Polanyians on Realism: an Introduction

    1. Introduction. The present issue of Tradition and Discovery on Polanyi's realism contains the papers that were written for the 1999 Annual Meeting of the Polanyi Society held in Boston, in conjunction with the AAR, on November 18-19. With the exception of Dale Cannon and John Puddefoot who had commitments elsewhere, the other contributors ...

  21. An Essay for Educators: Epistemological Realism Really is ...

    The authors defended the notion of science as universal in a previous article. The present purpose is to offer a commonsense argument in defense of critical realism as an epistemology and the epistemically distinguished position of science (rather than privileged) within a framework of epistemological pluralism.

  22. Scientific Realism Essay

    Scientific Realism Essay. 787 Words2 Pages. Scientific realism states that our knowledge of an object is acquired by the ideas created from our experience of it, not from direct perceptions. Our ideas are not the object itself but a representation of it. The theory states that the world is of mind-independent objects (people, animals, trees ...

  23. Essay on Realism

    Essay # 4. Realism in Education: Realism asserts that education is a preparation for life, for education equips the child by providing adequate training to face the crude realities of life with courage as he or she would perform various roles such as a citizen, a worker, a husband, a housewife, a member of the group, etc. ... Based on Science ...

  24. Book Review: 'The Cemetery of Untold Stories,' by Julia Alvarez

    The best stories live within us long after the final word; characters and places continue beyond the lines on a page. Yet the image of all the unfinished, unsatisfying, impossible stories we leave ...