首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 215 毫秒
1.
Alexandre Koyré was one of the most prominent historians of science of the twentieth century. The standard interpretation of Koyré is that he falls squarely within the internalist camp of historians of science—that he focuses on the history of the ideas themselves, eschewing cultural and sociological interpretations regarding the influence of ideologies and institutions on the development of science. When we read what Koyré has to say about his historical studies (and most of what others have said about them), we find him embracing and championing this Platonic view of his work. Ultimately I think this interpretation of Koyré's history of science is lopsided and in need of correction. I claim, rather, that a careful reading of Koyré's work suggests that a tension exists between internal and external methodological considerations. The external considerations stem from Koyré's commitment to the unity of human thought and the influence he admits that the ‘transscientifiques’ (philosophy, metaphysics, religion) have on the development of science. I suggest in conclusion then, that if we are to put a philosophical label on his work, rather than ‘Platonist’, as has been the custom, ‘Hegelian’ makes a better fit.  相似文献   

2.
For the history of science the 1940s were a transformative decade, when salient scholars like Herbert Butterfield or Alexandre Koyré set out to shape postwar culture by promoting new standards for understanding science. Some years ago I placed these developments in a tradition of enduring arts-science tensions and the contemporary notion that previous, “scientistic”, historical practices needed to be confronted with disinterested codes of historical craft (Mayer, 2000). Here, I want to further explore the ideological dimensions of the processes through which the academic study of science became institutionalized. Butterfield’s generation of science historians moulded perception of science in highly specific ways. Whereas the scientist-historians of the 1930s put scientific innovation into its socio-economic contexts, postwar accounts portrayed the birth of modern science as an intellectual revolution. Anti-Marxism formed a defining feature of the process by which the image of scientific work as a disinterested journey of the mind came to be institutionalized. Rather than spelling the end of ideology, appointments processes in the early Cold War years reveal disagreement about what science was to be invariably coextensive with dissent about social and political order. Rather than testifying to irreconcilable conflicts between interestedness and historical craft, the work of both the 1930s and 40s speaks of surprisingly productive relations between the two.  相似文献   

3.
Since the late 1980s, presentism has seen a resurgence among some historians of science. Most of them draw a line between a good form of presentism and typical anachronism, but where the line should be drawn remains an open question. The present article aims at resolving this problem. In the first part I define the four main distinct forms of presentism at work in the history of science and the different purposes they serve. Based on this typology, the second part reconsiders what counts as anachronism, Whiggism and positivist history. This clarification is used as a basis to rethink the research program of historical epistemology in the third section. Throughout this article, I examine the conceptual core of historical epistemology more than its actual history, from Bachelard to Foucault or others. Its project should be defined – as Canguilhem suggested – as an attempt to account for both the contingency and the rationality of science. As such, historical epistemology is based on a complex fifth form of presentism, which I call critical presentism. The critical relation at stake not only works from the present to the past, because of the acknowledged rationality of science, but also from the past to the present because of the contingency and historicity of scientific knowledge.  相似文献   

4.
At first glance twentieth-century philosophy of science seems virtually to ignore chemistry. However this paper argues that a focus on chemistry helped shape the French philosophical reflections about the aims and foundations of scientific methods. Despite patent philosophical disagreements between Duhem, Meyerson, Metzger and Bachelard it is possible to identify the continuity of a tradition that is rooted in their common interest for chemistry. Two distinctive features of the French tradition originated in the attention to what was going on in chemistry.French philosophers of science, in stark contrast with analytic philosophers, considered history of science as the necessary basis for understanding how the human intellect or the scientific spirit tries to grasp the world. This constant reference to historical data was prompted by a fierce controversy about the chemical revolution, which brought the issue of the nature of scientific changes centre stage.A second striking—albeit largely unnoticed—feature of the French tradition is that matter theories are a favourite subject with which to characterize the ways of science. Duhem, Meyerson, Metzger and Bachelard developed most of their views about the methods and aims of science through a discussion of matter theories. Just as the concern with history was prompted by a controversy between chemists, the focus on matter was triggered by a scientific controversy about atomism in the late nineteenth-century.  相似文献   

5.
This article examines how Hans G. Gadamer’s philosophical hermeneutics can contribute to contemporary debates on the concept of ‘presentism’. In the field of the history of science, this term is usually employed in two ways. First, ‘presentism’ refers to the kind of historiography which judges the past to legitimate the present. Second, this concept designates the inevitable influence of the present in the interpretation of the past. In this paper, I argue that both dimensions of the relationship between the present and the past are explored by Hans G. Gadamer in Truth and Method and other texts. In the first place, Gadamer’s critique of historicism calls into question the anti-presentist ideal of studying the past for ‘its own sake’. In the second place, Gadamer’s thesis that all understanding inevitably involves some prejudice poses the question of the inherent “present-centredness” of historical interpretations. By examining Gadamer’s hermeneutics, I seek to provide historians with new arguments and perspectives on the question of ‘presentism’.  相似文献   

6.
In the eighteenth century, the historiography of astronomy was part of a wider discussion concerning the history of the human spirit. The concept of the human spirit was very popular among Enlightenment authors because it gave the history of human knowledge continuity, unity and meaning. Using this concept, scientists and historians of science such as Montucla, Lalande, Bailly and Laplace could present the history of astronomy in terms of a progress towards contemporary science that was slow and could be interrupted at times, but was still constant, regular, and necessary. In my paper I intend to explain how the originally philosophical concept of the human spirit was transferred to the history of astronomy. I also introduce the basic principles to which the development of the spirit is subject in astronomy, according to historians of astronomy. The third part of the paper describes how historians of astronomy took into account the effect of social and natural factors on the history of astronomy.  相似文献   

7.
Recently, many historians of science have chosen to present their historical narratives from the ‘actors’-eye view’. Scientific knowledge not available within the actors’ culture is not permitted to do explanatory work. Proponents of the Sociology of Scientific Knowledge (SSK) purport to ground this historiography on epistemological relativism. I argue that they are making an unnecessary mistake: unnecessary because the historiographical genre in question can be defended on aesthetic and didactic grounds; and a mistake because the argument from relativism is in any case incoherent.The argument of the present article is self-contained, but steers clear of metaphysical debates in the philosophy of science. To allay fears of hidden assumptions, the sequel, to be published in the following issue, will consider SSK’s prospects of succour from scientific realism, instrumentalism, and a metaphysical system of Bruno Latour’s own devising.  相似文献   

8.
In this paper I endeavour to bridge the gap between the history of material culture and the history of ideas. I do this by focussing on the intersection between metaphysics and technology—what I call ‘applied metaphysics’—in the oeuvre of the Jesuit scholar Athanasius Kircher. By scrutinising the interplay between texts, objects and images in Kircher’s work, it becomes possible to describe the multiplicity of meanings related to his artefacts. I unearth as yet overlooked metaphysical and religious meanings of the camera obscura, for instance, as well as of various other optical and magnetic devices. Today, instruments and artefacts are almost exclusively seen in the light of a narrow economic and technical concept. Historically, the ‘use’ of artefacts is much more diverse, however, and I argue that it is time to historicize the concept of ‘utility’.  相似文献   

9.
In my paper I argue for mobilising recent material heritage at universities in teaching history of contemporary science. Getting your hands dirty in the messy worlds of the laboratory and the storage room, and getting entangled with the commemorative practices of scientists and technicians does not belong to the common experiences of students in history and philosophy of science. Despite the recent material turn in cultural studies, students’ engagement with the material world often remains a linguistic exercise, extending at most to an excursion to the sanitised and academically encultured world of the museum exhibit.I contrast this approach by drawing on experiences of taking students to the Atomei, Germany’s oldest research reactor at the Garching campus of the Munich University of Technology. Decommissioned since 2000, the installation and its history are still controlled by scientists. Studying contemporary laboratories and their materiality has so far been the domain of sociologists and ethnographers. I argue for opening these spaces to historians of science and engaging with the ‘unfinished’ material world of contemporary science. Taking the material seriously beyond the linguistic turn and asking students to explore laboratories and other sites of knowledge production challenges existing histories and historiographies. By exploring local university departments and their recent histories through their material heritage, we can observe everyday science and confront scientists and technicians’ cultures with those of historians’. By engaging with recent material heritage as historians and archivists, students can make an important contribution to enhancing the awareness about this heritage, its implications for history writing, as well as its documentation and preservation.  相似文献   

10.
In his account of probable reasoning, Poincaré used the concept, or at least the language, of conventions. In particular, he claimed that the prior probabilities essential for inverse probable reasoning are determined conventionally. This paper investigates, in the light of Poincaré's well known claim about the conventionality of metric geometry, what this could mean, and how it is related to other views about the determination of prior probabilities. Particular attention is paid to the similarities and differences between Poincaré's conventionalism as it applies to probabilities and de Finetti's subjectivism. The aim of the paper is to suggest that in accounts of the development of ideas about probable reasoning, particularly those customarily described as Bayesian, Poincaré's discussion deserves more attention than it has so far received.  相似文献   

11.
Traditionally the domain of scientists, the history of science became an independent field of inquiry only in the twentieth century and mostly after the Second World War. This process of emancipation was accompanied by a historiographical departure from previous, ‘scientistic’ practices, a transformation often attributed to influences from sociology, philosophy and history. Similarly, the liberal humanists who controlled the Cambridge History of Science Committee after 1945 emphasized that their contribution lay in the special expertise they, as trained historians, brought to the venture. However, the scientists who had founded the Committee in the 1930s had already advocated a sophisticated contextual approach: innovation in the history of science thus clearly came also from within the ranks of scientists who practised in the field. Moreover, unlike their scientist predecessors on the Cambridge Committee, the liberal humanists supported a positivistic protocol that has since been criticized for its failure to properly contextualize early modern science. Lastly, while celebrating the rise of modern science as an international achievement, the liberal humanists also emphasized the peculiar Englishness of the phenomenon. In this respect, too, their outlook had much in common with the practices from which they attempted to distance their project.  相似文献   

12.
The characteristics of inductivist historiography of science, as practised by earlier scientist/historians, and Whig historiography, as practised by earlier political historians, are described, according to the accounts of Agassi and Butterfield. It is suggested that the writings of Geikie on the history of geology allow us to characterize him as a Whig/inductivist historian of science who formulated anachronistic judgements. It is further suggested that his writings have had a considerable long-term effect on interpretations of the history of geology. The character of Geikie's historiography is related to his social, political and religious views, his historicism, and his romantic enthusiasm for Nature. His methodological pronouncements are examined: he believed that the present is the key to the past, and also that the past is the key to the present. He was an empiricist and inductivist. These epistemological and methodological views impinged on his historiography of science. If one attempts to criticize Geikie's historiography, though one may try to judge his work according to the norms of his own day and age, historical anachronism cannot be avoided entirely and one may oneself be charged with acting ‘Whiggishly’. Historians of science, in the process of professionalization, have accepted the historiographical norms of general historians (perhaps for socio-economic reasons), but in so doing have inherited a problem that arises whenever they contemplate earlier historical writings. It is suggested that there may be more room for Whiggish historiography of science than is presently deemed acceptable. Alternatively, one may wish to draw a careful distinction between science and its meta-discipline, the historiography of science. Whig historiography conflates the two, and the work of Geikie (a scientist/historian par excellence) provides a good illustration of this.  相似文献   

13.
In this article I respond to the defences of the Strong Programme put forward by David Bloor and Márta Fehér in this issue. I dispute the claim that it is attention to only limited parts of the Strong Programme framework that allows me to argue that this approach: (i) leads to weak idealism, (ii) undermines the idea that theories have varying levels of instrumental success, and (iii) challenges the theoretical claims of scientific actors. Rather, I argue that these problematic positions are entailed by the constructionist tenets at the core of the Strong Programme.  相似文献   

14.
This paper analyzes the claim that the Left Vienna Circle (LVC) offers a theoretical and historical precedent for a politically engaged philosophy of science today. I describe the model for a political philosophy of science advanced by LVC historians. They offer this model as a moderate, properly philosophical approach to political philosophy of science that is rooted in the analytic tradition. This disciplinary-historical framing leads to weaknesses in LVC scholars’ conception of the history of the LVC and its contemporary relevance. In this light, I examine the claim that there are productive enrichments to be gained from the engagement of feminist philosophy of science with the LVC, finding this claim ill-formulated. The case of LVC historiography and feminist philosophy of science presents a revealing study in the uses and ethics of disciplinary history, showing how feminist and other perspectives are misconceived and marginalized by forms of disciplinary self-narrativizing.  相似文献   

15.
This paper employs the revised conception of Leibniz emerging from recent research to reassess critically the ‘radical spiritual revolution’ which, according to Alexandre Koyré’s landmark book, From the closed world to the infinite universe (1957) was precipitated in the seventeenth century by the revolutions in physics, astronomy, and cosmology. While conceding that the cosmological revolution necessitated a reassessment of the place of value-concepts within cosmology, it argues that this reassessment did not entail a spiritual revolution of the kind assumed by Koyré, in which ‘value-concepts, such as perfection, harmony, meaning and aim’ were shed from the conception of the structure of the universe altogether. On the contrary, thanks to his pioneering intuition of the distinction between physical and metaphysical levels of explanation, Leibniz saw with great clarity that a scientific explanation of the universe which rejected the ‘closed world’ typical of Aristotelian cosmology did not necessarily require the abandonment of key metaphysical doctrines underlying the Aristotelian conception of the universe. Indeed the canon of value-concepts mentioned by Koyré—meaning, aim, perfection and harmony—reads like a list of the most important concepts underlying the Leibnizian conception of the metaphysical structure of the universe. Moreover, Leibniz’s universe, far from being a universe without God—because, as Clarke insinuated, it does not need intervention from God—is a universe which in its deepest ontological fabric is interwoven with the presence of God.  相似文献   

16.
Nineteenth-century spiritism was a blend of religious elements, the philosophy of mind, science and popular science and contacts with extraterrestrials were a commonplace phenomenon during spiritistic séances. Using the example of Carl du Prel (1839–1899) I show how his comprehensive mystic philosophy originated in a theory of extraterrestrial life. Carl du Prel used a Darwinian and monistic framework, theories of the unconscious and a Neo-Kantian epistemology to formulate a philosophy of astronomy and extraterrestrial life. He claimed that the mechanism of Darwinian selection is responsible for the distribution of stars and the orbits of the planets. In his speculations on the nature of extraterrestrial life he used the concept of organ projection to argue that technical solutions on earth will be realized organically on other planets and claimed that superior extraterrestrials have quantitatively and qualitatively different senses and thus different forms of intuition. A comparison with Camille Flammarion, spiritist and populariser of astronomy, demonstrates the contextual complexities of spiritism. In contrast to du Prel’s sober Neo-Kantian philosophical speculations, Flammarion was a late proponent of a French esoteric tradition that was rooted in romantic socialism, painted grand cosmological vistas and emphasized reincarnation. I put forward the hypothesis that current discourses on extraterrestrial life are affected by the spiritist tradition mainly through the ‘Golden Age’ science fiction literature of the 1940s and 50s and its successors. However, neither Carl du Prel nor Camille Flammarion contributed significantly to this tradition, which is mainly shaped by the psychical research of J. B. Rhine.  相似文献   

17.
Historical research on John Dalton has been dominated by an attempt to reconstruct the origins of his so-called “chemical atomic theory”. I show that Dalton’s theory is difficult to define in any concise manner, and that there has been no consensus as to its unique content among his contemporaries, later chemists, and modern historians. I propose an approach which, instead of attempting to work backward from Dalton’s theory, works forward, by identifying the research questions that Dalton posed to himself and attempting to understand how his hypotheses served as answers to these questions. I describe Dalton’s scientific work as an evolving set of puzzles about natural phenomena. I show how an early interest in meteorology led Dalton to see the constitution of the atmosphere as a puzzle. In working on this great puzzle, he gradually turned his interest to specifically chemical questions. In the end, the web of puzzles that he worked on required him to create his own novel philosophy of chemistry for which he is known today.  相似文献   

18.
In his 1785-review of the Ideen zur Philosophie der Geschichte der Menschheit, Kant objects to Herder's conception of nature as being imbued with active forces. This attack is usually evaluated against the background of Kant's critical project and his epistemological concern to caution against the “metaphysical excess” of attributing immanent properties to matter. In this paper I explore a slightly different reading by investigating Kant's pre-critical account of creation and generation. The aim of this is to show that Kant's struggle with the forces of matter has a long history and revolves around one central problem: that of how to distinguish between the non-purposive forces of nature and the intentional powers of the mind. Given this history, the epistemic stricture that Kant's critical project imposes on him no longer appears to be the primary reason for his attack on Herder. It merely aggravates a problem that Kant has been battling with since his earliest writings.  相似文献   

19.
This paper discusses the historiography of the ‘two cultures’ controversy. C. P. Snow’s lament about the ‘two cultures’, literary and scientific, has inspired a wide range of comment—much of which begins by citing Snow and his thesis, before going on to discuss very different things. This paper focuses upon one strand of this commentary, the historical analysis of the controversy itself. A ‘historical’ analysis is defined here as one that resists the impulse to enter the argument on behalf of Snow or Leavis, to conceive of their argument in the terms that Snow defined, or to invoke their argument as a precursor to some contemporary issue. Instead, a historical interpretation registers distance between that day and this, takes the controversy itself as its object of study, and explores the tensions and associations that came to be packed into those now familiar terms. As the fiftieth anniversary of Snow’s Rede Lecture nears, this approach—rather than the repetition of clichés about the bridging of cultures—offers both analytical perspective on the controversy and interpretive possibilities for its examination.  相似文献   

20.
Historians of science have frequently sought to exclude modern scientific knowledge from their narratives. Part I of this paper, published in the previous issue, cautioned against seeing more than a literary preference at work here. In particular, it was argued—contra advocates of the Sociology of Scientific Knowledge (SSK)—that a commitment to epistemological relativism should not be seen as having straightforward historiographical consequences. Part II considers further SSK-inspired attempts to entangle the currently fashionable historiography with particular positions in the philosophy of science. None, I argue, is promising. David Bloor’s proposed alliance with scientific realism relies upon a mistaken view of contrastive explanation; Andrew Pickering’s appeal to instrumentalism is persuasive for particle physics but much less so for science as a whole; and Bruno Latour’s home-grown metaphysics is so bizarre that its compatibility with SSK is, if anything, a further blow to the latter’s plausibility.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号