The One Culture? A Conversation about Science
edited by Jay A. Labinger and Harry Collins
University of Chicago Press, 2001
Cloth: 978-0-226-46722-1 | Paper: 978-0-226-46723-8 | Electronic: 978-0-226-46724-5
DOI: 10.7208/chicago/9780226467245.001.0001
ABOUT THIS BOOKAUTHOR BIOGRAPHYTABLE OF CONTENTS

ABOUT THIS BOOK

So far the "Science Wars" have generated far more heat than light. Combatants from one or the other of what C. P. Snow famously called "the two cultures" (science versus the arts and humanities) have launched bitter attacks but have seldom engaged in constructive dialogue about the central issues. In The One Culture?, Jay A. Labinger and Harry Collins have gathered together some of the world's foremost scientists and sociologists of science to exchange opinions and ideas rather than insults. The contributors find surprising areas of broad agreement in a genuine conversation about science, its legitimacy and authority as a means of understanding the world, and whether science studies undermines the practice and findings of science and scientists.

The One Culture? is organized into three parts. The first consists of position papers written by scientists and sociologists of science, which were distributed to all the participants. The second presents commentaries on these papers, drawing out and discussing their central themes and arguments. In the third section, participants respond to these critiques, offering defenses, clarifications, and modifications of their positions.

Who can legitimately speak about science? What is the proper role of scientific knowledge? How should scientists interact with the rest of society in decision making? Because science occupies such a central position in the world today, such questions are vitally important. Although there are no simple solutions, The One Culture? does show the reader exactly what is at stake in the Science Wars, and provides a valuable framework for how to go about seeking the answers we so urgently need.

Contributors include:
Constance K. Barsky, Jean Bricmont, Harry Collins, Peter Dear, Jane
Gregory, Jay A. Labinger, Michael Lynch, N. David Mermin, Steve
Miller, Trevor Pinch, Peter R. Saulson, Steven Shapin, Alan Sokal,
Steven Weinberg, Kenneth G. Wilson

AUTHOR BIOGRAPHY

Jay A. Labinger is a research chemist and administrator of the Beckman Institute at the California Institute of Technology.

Harry Collins is a Distinguished Research Professor of Sociology and director of the Centre for the Study of Knowledge, Expertise, and Science at Cardiff University.

TABLE OF CONTENTS

Preface

Introduction

Part One: Positions

- Trevor Pinch
DOI: 10.7208/chicago/9780226467245.003.0002
[science studies, science wars, science and technology, natural science, symbolic legitimation, material resources]
The “science wars” refers to a debate raging within the academy and without over the status of fields like science and technology studies and cultural studies of science and technology. The debate has largely been initiated by natural scientists who have written books and made public statements critical of science studies and what they take to be some of its central ideas. The idea of the two cultures was made famous by C. P. Snow in the 1950s. It has come to refer to the separate self-contained cultures of the humanities and of the natural sciences. The two cultures were held to be largely ignorant of each other. What is perhaps surprising is why the two cultures have been able to coexist peacefully for so long. Aside from one or two skirmishes, such as that between Wittgenstein and Turing, the science wars are more the exception than the rule. It is important to note that radical disjunctions in our cultural sensibilities and practices need not necessarily lead to clashes. As long as each culture can flourish, receive ample material resources and symbolic legitimation, there need be little reason for dissent. (pages 13 - 26)
This chapter is available at:
    https://academic.oup.com/chica...

- Jean Bricmont, Alan Sokal
DOI: 10.7208/chicago/9780226467245.003.0003
[philosophy of science, radical relativism, epistemology, Strong Programme, sociology of science, war and peace]
This chapter aims to disentangle various confusions caused by fashionable ideas in the contemporary philosophy of science. Roughly speaking, the main thesis is that those ideas contain a kernel of truth that can be understood properly when those ideas are carefully formulated, but then they give no support to radical relativism. However, before getting to work, the chapter attempts to avoid several possible misunderstandings and to emphasize some points that are noncontroversial. First, the chapter clears the epistemological ground. Next, it formulates objections to the “Strong Programme.” And finally, it indicates some possible areas of collaboration between scientists and sociologists of science. (pages 27 - 47)
This chapter is available at:
    https://academic.oup.com/chica...

- Michael Lynch
DOI: 10.7208/chicago/9780226467245.003.0004
[science wars, science peace process, war analogy, verbal debates, peace talk, talks]
The polarization associated with the “science wars” is unacceptable to many, and they would like to promote a “science peace process.” Before initiating such a peace process, however, it may be worth considering whether war is going on in the first place. If the “war” analogy is inappropriate in this instance, then there should be no need to declare peace. A peace process is an effort to start “talks” aiming to resolve, or at least interrupt, a chronic and violent conflict. The idea is that the combatants should talk instead of trying to destroy one another. Contrary to the usual kind of war, the science wars have been verbal debates between members of different academic fields. Until recently, many of the debaters had little to say to one another. As arguments often do, these particular debates frequently become heated and denunciatory, but it seems unnecessary to suspend a state of war in order to initiate peace talks. The state of this war consists of talk, and more talk, so it may make more sense to devise more interesting and respectful ways of talking and writing. (pages 48 - 60)
This chapter is available at:
    https://academic.oup.com/chica...

- Jane Gregory, Steve Miller
DOI: 10.7208/chicago/9780226467245.003.0005
[science wars, postwar era, educational movement, science broadcasting, science publishing, Stephen Hawking]
The meaning of the phrase “public understanding of science” has evolved and diversified in the postwar era: what used to refer to the little-understood and barely interesting phenomenon of the conception of sciences among laypeople now serves a variety of purposes. It provides a label for normative and operational definitions of what the public understands about science, as well as for policy in the area and for the social and educational movement the idea has spawned. Alongside the increasing interest in the more programmatic aspects of this latest surge in public understanding of science activity developed a boom in popular science broadcasting and publishing, epitomized perhaps by the success of Stephen Hawking's A Brief History of Time. But this boom also brought a definition of science into the public domain that was broader than the scientific establishment had envisaged when they urged the media to carry more science. (pages 61 - 72)
This chapter is available at:
    https://academic.oup.com/chica...

- Peter R. Saulson
DOI: 10.7208/chicago/9780226467245.003.0006
[physical scientist, scientific work, natural world, social structure, science studies, formula]
There may always be some unavoidable tensions between communities, based on where the different curiosities lie. Physical scientists tend mostly to be curious about the subject of their scientific work, much less so about the social structure of the world in which they carry out their scientific work (except in the form of gossip, which is universal). To the extent that we are curious about how science works, it tends to be related to trying to understand reasons for the remarkable success science has had over the past few centuries in building an ever deeper account of the natural world. In search of explanations for this success, we tend to look for a simple overarching principle, such as the formula that science succeeds by looking for the simplest possible explanation. (pages 73 - 82)
This chapter is available at:
    https://academic.oup.com/chica...

- N. David Mermin
DOI: 10.7208/chicago/9780226467245.003.0007
[academic exchanges, relativity, folklore physicists, contemporary knowledge, sociologists, David Bloor]
This chapter reexamines exchanges with sociologists on relativity, with a view to identifying avoidable ways in which they misunderstood each other, seeing what residual disagreements remain, and noting the many similarities with a more recent exchange had with Barry Barnes and David Bloor about their textbook Scientific Knowledge. The sociologist or historian clearly must distinguish what was known at the time under study from what is currently known. It is all too easy for late-twentieth-century knowledge to induce anachronistic misinterpretations of how knowledge developed in the early twentieth century. All the historic folklore physicists tell each other suffers from such distortion. But preventing such contamination by prohibiting any use of or reference to contemporary knowledge has its own risks. (pages 83 - 98)
This chapter is available at:
    https://academic.oup.com/chica...

- Steven Shapin
DOI: 10.7208/chicago/9780226467245.003.0008
[metascience, sociologists of science, natural scientists, modern culture, rationality, institutions of science]
The immediate occasion for the science wars seems to be a series of claims about science made by some sociologists, cultural historians, and fuzzy-minded philosophers. As a matter of convenience, the chapter refers to propositions about science as “metascience,” and, because it is very important to be clear about what is at issue, the chapter lists here just a few of the more contentious and provocative metascientific claims. For many readers, even listing such statements is unnecessary: they will already be thoroughly familiar with sentiments like these associated with the writings of sociologists of science and academic fellow travelers, as they will be equally familiar with the outraged reactions to them expressed by a number of natural scientists, convinced that such claims are motivated mainly or solely by hostility to science, or that they proceed from ignorance of science, or both. Science and rationality are said to be besieged by barbarians at the gate, and, unless such assertions are exposed for the rubbish they are, the institution of science, and its justified standing in modern culture, will be at risk. (pages 99 - 115)
This chapter is available at:
    https://academic.oup.com/chica...

- Steven Weinberg
DOI: 10.7208/chicago/9780226467245.003.0009
[scientific knowledge, historical knowledge, physics, quantum mechanics, theory of gravitation, space and time]
This chapter discusses the uses that historical and scientific knowledge have for each other, but first it wants to take up what may be a more unusual topic: the dangers that history poses for physics, and physics for history. The danger in history for the work of physics is that, in contemplating the great work of the past—great heroic revolutions like relativity, quantum mechanics, and so on—we develop such respect for them that we become unable to reassess their place in a final physical theory. General relativity provides a good example. As developed by Einstein in 1915, general relativity appears almost logically inevitable. One of its fundamental principles, the equivalence of gravitation and inertia, says that there is no difference between gravity and the effects of inertia such as centrifugal force. This principle of equivalence can be reformulated as the principle that gravity is just an effect of the curvature of space and time—a beautiful principle from which Einstein's theory of gravitation follows almost uniquely. (pages 116 - 127)
This chapter is available at:
    https://academic.oup.com/chica...

- Peter Dear
DOI: 10.7208/chicago/9780226467245.003.0010
[science studies, epistemography, philosophy of science, sociology of science, intellectual endeavors, genealogy]
The ambiguities of the catch-all label “science studies” are not accidental: such crypto-disciplinary names tend to be hobbled from the start by the need to draw in as many constituencies as possible, thereby running the risk of yoking together quite distinct intellectual endeavors. If (no doubt, contestably) we date science studies from the later 1970s, we immediately perceive a genealogy. Before science studies, there was a fairly well-established specialty called the history and philosophy of science. There was also a subfield of sociology known as the sociology of science. Science studies has now swallowed up large amounts of what once fell into those earlier categories, without always thoroughly digesting them. But the parts that were left behind tell their own stories. (pages 128 - 141)
This chapter is available at:
    https://academic.oup.com/chica...

- Kenneth G. Wilson, Constance K. Barsky
DOI: 10.7208/chicago/9780226467245.003.0011
[social construction, sociology of science, Thomas Kuhn, scientific research, truth, science studies]
This chapter discusses the issue of cultural specificity and social construction in detail. What really matters is to identify root causes for the eruption of controversy about social construction. The chapter proposes that the controversy traces back to an internal problem of the sociology of science, that the sociologists were misdirected by their reading and interpretation of Thomas Kuhn's The Structure of Scientific Revolutions (1996). It is only now becoming possible, in hindsight, to recognize the shortcomings in Kuhn's book that caused the misdirection. The chapter suggests a new interpretation of Kuhn's opus that could open up a new direction for future research in the sociology of science. It begins, however, with a question regarding the version of “truth” that scientists use, a question that has been raised by sociologists of science and that underlies a legitimate but limited role for social construction in science. (pages 142 - 155)
This chapter is available at:
    https://academic.oup.com/chica...

- Harry Collins
DOI: 10.7208/chicago/9780226467245.003.0012
[sociology of science, religion, socialization, cultural polarization, social psychology, political interests]
The opening poem in this chapter describes a Martian's view of earthly things seen for the first time; it is a stranger's view. There has to be something strange about the sociologist's view of science, too. It is the job of the sociologist to estrange him or herself so that those things that are taken for granted in the native society—those things that seem just common sense—become things that require explanation. Accounts of the everyday life of a society from an estranged perspective can seem threatening. Consider how the sociologist might talk of religious beliefs. Imagine the sociologist trying to understand how it is that south of a certain line in a certain country it is generally believed that a certain kind of religious ceremony can turn wine into blood, whereas north of that line it is generally believed that no such thing occurs. The sociologist's account would be in terms of the histories of the respective beliefs; the processes of socialization through which the young are inducted into the beliefs; the separate social networks which reinforce the separate beliefs; the divergent conceptual structures in which the separate beliefs are embedded; the day-to-day actions in which the conceptual structures are expressed; the economic and political interests which drive the upholders of these beliefs apart; and, perhaps, the social psychology of cultural polarization. (pages 156 - 166)
This chapter is available at:
    https://academic.oup.com/chica...

- Jay A. Labinger
DOI: 10.7208/chicago/9780226467245.003.0013
[professional scientists, positive engagements, science studies, scientific ranks, scientific enterprise, science education]
It is doubtful whether this situation is alterable to any great extent. Most professional scientists have neither the inclination nor incentives to divert any appreciable fraction of their time and energies from their main pursuits. But should one try? Some want scientists to become more aware of science studies so that they can better defend against the dangers posed by the “antiscience brigades.” In contrast, calls for positive engagements have been rather rare even on the part of science studies practitioners, who one might think would have a strong interest in enlisting allies from the scientific ranks. This chapter argues that a defensive stance is not warranted and then tries to present a case for the potential benefits that could be realized if (at least some) scientists were to take a more serious interest in science studies. (pages 167 - 176)
This chapter is available at:
    https://academic.oup.com/chica...

Part Two: Commentaries

- Jean Bricmont, Alan Sokal
DOI: 10.7208/chicago/9780226467245.003.0014
[methodological relativism, antiscience, Strong Programme, scientific beliefs, sociology, contemporary research]
Sociologists frequently admit that they don't have the background to evaluate whether the claims made by scientists (particularly concerning contemporary research) are rationally justified or not, but then they assert that they are not obliged to make any such evaluation: they are concerned with social phenomena, not with physical or biological ones, and so are perfectly justified in ignoring this latter aspect. That would perhaps be fine if their aims were more modest than those of the “Strong Programme”: if, for example, they claimed only to recount some of the factors affecting the acceptance of scientific beliefs, without purporting to judge their relative importance. But in that case they ought not claim to give a causal account of the acceptance of scientific beliefs, when important parts of the cause—usually the dominant parts, in our view—are excluded a priori from consideration. (pages 179 - 183)
This chapter is available at:
    https://academic.oup.com/chica...

- Harry Collins
DOI: 10.7208/chicago/9780226467245.003.0015
[epistemological relativism, ontological relativism, methodological relativism, social-scientist investigator, philosophical relativism, sociology of science, scientific knowledge]
Epistemological relativism implies that one social group's way of justifying its knowledge is as good as another's and that there is no external vantage point from which to judge between them; all that can be known can be known only from the point of view of one social group or another. Ontological relativism seems to be the view that within social groups, reality itself is different. We can call any combination of epistemological and ontological relativism “philosophical relativism.” Methodological relativism says nothing direct about reality or the justification of knowledge. Methodological relativism is an attitude of mind recommended to the social-scientist investigator: the sociologist or historian should act as though the beliefs about reality of any competing groups being investigated are not caused by the reality itself. Philosophical relativism does not make any difference to the practice of the sociology of scientific knowledge. (pages 184 - 195)
This chapter is available at:
    https://academic.oup.com/chica...

- Peter Dear
DOI: 10.7208/chicago/9780226467245.003.0016
[sociologists of science, descriptive accuracy, open-endedness, historical story of science, relativity of theories, Whiggish fallacy, Weinberg]
When historians and sociologists of science emphasize the contingency of historical outcomes by making the argument that things could always have come out differently, the point they are making is as much one of methodology as of descriptive accuracy. The methodological importance of stressing a supposed open-endedness in the historical story of science being told is that it counters the Whiggish fallacy to which Weinberg alludes. However, it is worth remembering exactly what that fallacy is. Weinberg represents it as concerning the absoluteness or historical relativity of theories. (pages 196 - 200)
This chapter is available at:
    https://academic.oup.com/chica...

- Jane Gregory
DOI: 10.7208/chicago/9780226467245.003.0017
[science studies, public understanding, truth, Bricmont, Sokal, sponsorship]
In science, replications, peer review, and publication in Nature are usually good enough: the end product is usually well on its way to becoming what Bricmont and Sokal might call “reality” or “truth.” And proponents of the public understanding of science have argued that understanding the peer-review and publications processes is key to the public's ability to distinguish reliable from unreliable knowledge. But what happened to Benveniste would be baffling to anyone who understood those processes as generating truths. Bricmont and Sokal call for replications from “independent” scientists, implying that those who had successfully replicated the work had automatically lost their independence by virtue of their success; the marginalization of Benveniste himself over the last decade is surely deterrent enough to any scientist who might attempt a replication now. Bricmont and Sokal also urge the homeopaths to shoulder the burden of proving their claims, and yet when the homeopaths sponsored Benveniste's reputable, state-run lab to investigate their phenomenon, the fact of that sponsorship was used to discredit his results. (pages 201 - 205)
This chapter is available at:
    https://academic.oup.com/chica...

- Jay A. Labinger
DOI: 10.7208/chicago/9780226467245.003.0018
[science wars, Bricmont, Sokal, scientific knowledge, experimental science, common sense]
The aspect this chapter is specifically concerned with is: to what degree is one justified in extrapolating from commonsense or “easy” cases to claims made with respect to science in general? For example, Bricmont and Sokal use the “obvious” question of whether or not it is raining as a starting point to address the role of reality in determining belief. They do recognize that there is an issue here: one can question whether “ordinary” and “scientific” knowledge should be treated the same way. Yes, they conclude: if reality constrains ordinary knowledge, it constrains scientific knowledge even more so because experimental science is carried out expressly to make that the case. Thus science would seem to be just an elaborate form of common sense. (pages 206 - 209)
This chapter is available at:
    https://academic.oup.com/chica...

- Michael Lynch
DOI: 10.7208/chicago/9780226467245.003.0019
[situated knowledge, science wars, epistemology, scientific work, scientific practice, methodological rules, technical work]
One of the prominent tendencies in the constructionist studies of science is a resolute insistence that science is work. Like other forms of work, scientific practice is viewed as an embodied and material labor process involving numerous, often obscure, parties. Connected with this emphasis is a stress upon the “local” or “situated” character of scientific and technical work—work that involves practical actions and reasoned judgments, which are not a matter of mechanically following methodological rules. Moments of creative struggle and opportunities for improvisation in a laboratory occur from top to bottom in a hierarchy of research directors, staff scientists, technicians, and civilian participants. The contingent products of this collective labor process (data, results, publications, discovery claims) are more than deliberately planned outcomes, as they can be sources of surprise and puzzlement. An understanding of the implications of this picture of scientific work may provide a basis for solidarity rather than epistemological infighting. (pages 210 - 215)
This chapter is available at:
    https://academic.oup.com/chica...

- N. David Mermin
DOI: 10.7208/chicago/9780226467245.003.0020
[human experience, body of knowledge, human activity, Harry Collins, quantum mechanics, linguistic practices]
No reliable body of knowledge can be undermined by viewing its acquisition as a collective human activity. Indeed the chapter goeso beyond Harry Collins and maintains that scientists can benefit from “estranging themselves from their own practices” in this way. What else was Einstein's recognition of the conventional character of simultaneity? Or Steven Weinberg's warning against “ill-placed loyalty” to “the great heroic ideas of the past”? Many of the interpretive confusions plaguing the foundations of quantum mechanics might reflect insufficient estrangement from unsound linguistic practices. (pages 216 - 220)
This chapter is available at:
    https://academic.oup.com/chica...

- Trevor Pinch
DOI: 10.7208/chicago/9780226467245.003.0021
[conversation, David Mermin, academic exchanges, David Bloor, social event, face-to-face meetings]
There is a well-known finding in the field of conversation analysis that participants in conversations display a proclivity to end in agreement. The sense of moving toward agreement, which one finds in the first round of contributions, is surely further indication that a real conversation is taking place. That a conversation is a social event should not be forgotten. It is surely no accident that the productive exchange (remarked upon by several contributors) between David Mermin and Collins and Pinch has been accompanied by many face-to-face meetings. Mermin himself notes the importance of his having spent time with David Bloor for his forming a charitable attitude toward Bloor's intentions. (pages 221 - 226)
This chapter is available at:
    https://academic.oup.com/chica...

- Peter R. Saulson
DOI: 10.7208/chicago/9780226467245.003.0022
[metaphysics, epistemology, politics, Alan Sokal, postmodernism, science studies]
Do people really care enough about metaphysics and epistemology to make them angry? Aside from a few philosophers, the chapter states that most do not. In particular, few scientists seem to pay any attention at all to the philosophical issues that surround science. Scientists do care about politics, some of them passionately. Alan Sokal explained in his Lingua Franca article that his intention in perpetrating his hoax was to help reclaim the left for the forces of reason from the mystics who hold it hostage today. Few other science warriors are so explicitly leftist; indeed a more common attitude would be that science studies, postmodernism, and what passes for the left on campus are all one big mess. Even so, few scientists care strongly enough about their politics to explain the anger behind the actions in the science wars. (pages 227 - 232)
This chapter is available at:
    https://academic.oup.com/chica...

- Shapin Steven
DOI: 10.7208/chicago/9780226467245.003.0023
[antiscience, university, natural science, science studies, states of mind, specialized work]
This chapter attempts to draw attention to the structural implications of a situation in which a group of academics are obliged repeatedly to attest, not just their competence in their special subjects (which, though unpleasant and uncommon, is certainly fair enough) but the innocuousness of their states of mind and intentions in doing their specialized work (which, the chapter states, is not). More generally, the cultural phenomenon of academic “antiscience” would be accorded a legitimacy it does not deserve, namely, positing its substantial and coherent existence. The dismal fate of the academic “antiscientists” would then be available as an object-lesson of what risks attend certain forms of inquiry. The university will be seen as a haven for free inquiry just on the condition that one naturalistically studies subjects. The lesson will be: study down; do not study up. (pages 233 - 237)
This chapter is available at:
    https://academic.oup.com/chica...

- Steven Weinberg
DOI: 10.7208/chicago/9780226467245.003.0024
[social setting, scientific research, science studies, Harry Collins, physical theories, scientific progress]
Although scientists recognize that their theories often bear the stamp of the social environment in which they are formulated, some like to think of this as an impurity, some slag left amid the metal, which they would eventually eliminate. There is the powerful attraction that true theories exert on scientists' thinking, an attraction that seems to have little to do with the social setting of their research. Scientists wonder why some historians of science such as Harry Collins are not interested in describing this process, the often slow and uncertain progress of physical theories toward an ultimate culture-free form that is the way it is because this is the way the world is. (pages 238 - 240)
This chapter is available at:
    https://academic.oup.com/chica...

Part Three: Rebuttals

- Jean Bricmont, Alan Sokal
DOI: 10.7208/chicago/9780226467245.003.0025
[Steven Shapin, radical relativism, beliefs, skepticism, witches, nonacademic culture]
This chapter addresses substantive criticisms. The chapter here attempts to set straight a few commentators' misreadings of the earlier arguments. For example, according to Steven Shapin, “Bricmont and Sokal attribute the cultural credibility of flat-Earth or witchcraft beliefs to ‘the existence of a radically relativist academic Zeitgeist.’” But no such thing happened. Rather, the text noted “the existence of a radically relativist academic Zeitgeist” in which some “otherwise reasonable researchers or university professors...will claim that witches are as real as atoms”—their obvious intent being to cast doubt on the existence of atoms, not to assert a sincere belief in the existence of witches—”or pretend to have no idea whether the Earth is flat, blood circulates, or the Crusades really took place” [emphasis added]. The text was thus discussing extreme relativism or skepticism in academia; it made no reference whatsoever to extreme credulity in the general nonacademic culture, much less did it claim that the latter is a causal consequence of the former. (pages 243 - 254)
This chapter is available at:
    https://academic.oup.com/chica...

- Harry Collins
DOI: 10.7208/chicago/9780226467245.003.0026
[science's authority, historical dimension, human action, climate change, political decision-making, science studies]
The different ways of justifying science are not mutually exclusive. Indeed, it may be the existence of crown jewels that makes one believe that rough diamonds are worth having. But to value murky and unfinished science only because of its potential to turn into something else is risky. There are many kinds of science that cannot be polished up to a fine sparkle. Think of any science that involves a long-term historical dimension, especially one involving human action such as the science of long-term climate change. However, to justify the use of science in political decision-making by reference only to polished jewels is to invite exclusion from political decision-making. (pages 255 - 260)
This chapter is available at:
    https://academic.oup.com/chica...

- Peter Dear
DOI: 10.7208/chicago/9780226467245.003.0027
[epistemography, David Mermin, modern beliefs, scientific beliefs, scientific ideas, historic events]
An observation of David Mermin's provides us with the occasion to clarify a little more the idea of “epistemography.” Mermin writes: “Peter Dear's remark that ‘modern scientific beliefs are in fact irrelevant to an epistemographical account’ can be too limiting when those beliefs can give clues about the objective circumstances of historic events.” There is, among other things, an important point about theory and practice here. As regards the practice of the historian of science, there is no doubt that knowledge of present-day scientific ideas often comes in very handy in wrestling with the (sometimes very alien) beliefs of the past—the more so the more recent the historical episode. That usefulness is, however, fundamentally heuristic. In orienting oneself with respect to a historical episode, it can happen that a more modern idea can cast an otherwise perhaps unexpected light on the material under investigation. (pages 261 - 262)
This chapter is available at:
    https://academic.oup.com/chica...

- Jay A. Labinger
DOI: 10.7208/chicago/9780226467245.003.0028
[Benveniste, conversation, Nature, peer review, ontology, science studies]
Both Pinch and Weinberg comment on the level of agreement the “conversation” seems to be approaching, which Pinch notes is the hallmark of a “real” conversation. This chapter seeks to offer some thoughts on the Benveniste case, which has been mentioned by a number of the commentators and also to respond to Bricmont and Sokal's criticism of the original contribution. Gregory notes that Benveniste's original work was replicated elsewhere, received positive peer review, and was published in Nature, which is “usually good enough” to start establishing the results as truth; and that anyone who understood these processes would be baffled by the subsequent developments. (pages 263 - 267)
This chapter is available at:
    https://academic.oup.com/chica...

- Michael Lynch
DOI: 10.7208/chicago/9780226467245.003.0029
[methodological relativism, philosophical relativism, working philosophies, causality, contemporary science]
Most sociologists and historians agree that their studies of historical and contemporary cases must take into account what scientists say about their own theories and practices. It also seems that many of the scientists recognize that studies of what historical and contemporary scientists do will not always vindicate what scientists and philosophers say about science in general. It is also clear that what seems to be emerging is not a truce between two discrete sides. Instead, it involves a much more varied collection of agreements and disagreements that crisscross the two-cultures divide. Many substantive disagreements remain, for example, about the relationship between methodological and philosophical relativism, but the overall tenor of many of the arguments is inquisitive rather than inquisitorial. Despite his support for the overall dialogue, this chapter expresses a couple of reservations about some of the terms of the peace agreement that may be emerging from several discourses. (pages 268 - 274)
This chapter is available at:
    https://academic.oup.com/chica...

- N. David Mermin
DOI: 10.7208/chicago/9780226467245.003.0030
[science, pseudoscience, scientific knowledge, astrology, David Bloor, Scientific Knowledge]
Years of public attacks on astrology by eminent scientists seem to have created the presumption that no scientist can address the subject in any context without deploring and condemning. This chapter criticizes that astrology provided a bad textbook illustration of an area that might shift across the boundary scientists draw between science and pseudoscience, because far too much scientific knowledge would have to be thrown out to accommodate astrology with any plausibility. This same disposition appeared in David Bloor's reply to a review of Scientific Knowledge. In criticizing their choice of astrology as an example the review emphasized that there was a difference between drawing sharp, precisely defined boundaries between science and nonscience and distinguishing between extreme cases. (pages 275 - 279)
This chapter is available at:
    https://academic.oup.com/chica...

- Trevor Pinch
DOI: 10.7208/chicago/9780226467245.003.0031
[natural science, social science, determinism, social resources, material resources, theoretical arguments]
Most controversies in the natural sciences are brought to a close in a comparatively short period of time with a clear victor. Most debates in the humanities and social sciences are seldom resolved so quickly and easily. “Free Will proclaimed as victor over Determinism at philosophy conference!” is not a headline one will find. In scientific controversies scientists seem too good at closing them down. Why this is so is, of course, itself a matter of dispute. For those in science studies such debates are settled by the presence of an effective community—a core set—within which a combination of rhetorical, cognitive, material, and social resources are brought to bear. For scientists it is more the case that the “truth will out,” sometimes by means of a few crucial experiments in tandem with irresistible theoretical arguments. (pages 280 - 282)
This chapter is available at:
    https://academic.oup.com/chica...

- Peter R. Saulson
DOI: 10.7208/chicago/9780226467245.003.0032
[natural world, state of knowledge, Newton's laws, quantum mechanics, wave functions, space-time]
Weinberg believes that among the present state of knowledge there exists a large measure of “hard” physics, on which one might agree even with creatures from other planets. Within restricted ranges of applicability, we have laws that describe outstandingly well the behavior of the natural world. Newton's laws will always be “true, mostly” in this sense, as will Maxwell's equations of the electromagnetic field, Einstein's field equations for gravity, and so on. Nonphysicists need to be reminded that what most interests most of them is probably among the “soft” material destined to be swept away. The “vision of reality that we use to explain to ourselves why the equations work” includes nearly all of the metaphysical concepts that accompany many theories: Newton's absolute space or Einstein's elastic space-time, the luminiferous ether or fields in the vacuum, the deterministic trajectories of classical physics or the probabilistic wave functions of quantum mechanics. (pages 283 - 288)
This chapter is available at:
    https://academic.oup.com/chica...

- Steven Weinberg
DOI: 10.7208/chicago/9780226467245.003.0033
[historiography, scientific knowledge, social influences, historical record, data, calculations]
The work of scientists in any era is motivated by the data and the calculations available to them, as well as a whole host of social, cultural philosophical, and psychological influences. But the data are the way they are (at least some of the time) because that is the way the world is, and scientists know more now about the way the world is than they did in the past. So why not use this knowledge? Admittedly, there would be no need to do so had there been perfect knowledge of the data and calculations available to past scientists. Knowing something more about the world than what the scientists of the past knew can help us to fill in the gaps that will always be in the historical record. (pages 289 - 290)
This chapter is available at:
    https://academic.oup.com/chica...

- Kenneth G. Wilson, Constance K. Barsky
DOI: 10.7208/chicago/9780226467245.003.0034
[social construction, scientific progress, social sciences, natural sciences, measurements, diverse cultures]
The process of scientific progress, while not error-free, can be used for many purposes by many diverse cultures as long as the measurements become accurate enough for their errors not to matter. Meanwhile, the larger body of scientific results that have yet to benefit from many rounds of continuous improvement can be problematic for any culture that tries to use them. There are many results from the natural sciences that are in this latter category; there are many results from the social sciences that are in this latter category, too. Research on the nature of continuous improvement in science could be beneficial for future users of both the natural and the social sciences. (pages 291 - 295)
This chapter is available at:
    https://academic.oup.com/chica...

35. Conclusion

References

Contributors

Index