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 15 毫秒
1.
This paper aims to analyse Karl Popper's conception of ‘three worlds’, and especially the problem of world 3—the world of objective knowledge. Firstly, I try to explain Popper's turn to ontological questions which I link to his antipsychologism and to issues raised by the development of logic after World War II. I then consider Popper's concept of the autonomy of world 3 and his attempt to introduce world 3 as a world of knowledge without a knowing subject. I conclude that Popper did not succeed in unifying his central idea of autonomy of knowledge with the requirement of the creative role of the critical subject carrying out the evolution of knowledge. I see the core of this contradiction especially in his co‐existing ideas of the timeless existence of world 3 and the elimination of the subject from it. The attempt to desubjectivize the realm of objective knowledge leads to a philosophically unbalanced standpoint which presupposes a creative subject and at the same time neglects it. Finally, I question Popper's account of the growth of world 3. Popper considers only cognitive motivations, and excludes a broad range of motivating factors which originate in the problems which we face in our lives, and affect our cognitive interests in world 3.  相似文献   

2.
Eric Scerri has proposed an account of how reduction might be understood in chemistry. He claims to build on a general aspect of Popper's views which survives his otherwise heavy criticism, namely adherence to actual scientific practice. This is contrasted with Nagel's conception, which Scerri takes to be the philosopher's standard notion. I argue that his proposal, interesting though it is, is not so foreign to ideas in the tradition within which Nagel wrote as Scerri would have us believe. Moreover, actual scientific practice can be commandeered in support of a holistic conception which Popper contrasted with what he saw as the admirable strivings towards reduction in science.  相似文献   

3.
Most philosophers of science maintain Confirmationism's central tenet, namely, that scientific theories are probabilistically confirmed by experimental successes. Against this dominant (and old) conception of experimental science, Popper's well-known, anti-inductivistic Falsificationism (’Deductivism’) has stood, virtually alone, since 1934. Indeed, it is Popper who tells us that it was he who killed Logical Positivism. It is also pretty well-known that Popper blames Wittgenstein for much that is wrong with Logical Positivism, just as he despises Wittgenstein and Wittgensteinian philosophers for abdicating philosophy's true mission. What is not well-known, however, especially because Popper neglected to tell us in 1934, is that Wittgenstein is very much an ally. It was Wittgenstein who rejected induction in the strongest possible terms as early as 1922, and it was Wittgenstein who similarly rejected Confirmationism approximately four years prior to Popper. The aims of this paper are to illuminate the substantial agreements between Popper and Wittgenstein and, by doing so, to clarify their important disagreement regarding the status of “strictly universal,” scientific theories (or hypotheses). This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

4.
On the basis of the Suppes–Sneed structuralview of scientific theories, we take a freshlook at the concept of refutability,which was famously proposed by K.R. Popper in 1934 as a criterion for the demarcation of scientific theories from non-scientific ones, e.g., pseudo-scientificand metaphysical theories. By way of an introduction we argue that a clash between Popper and his critics on whether scientific theories are, in fact, refutablecan be partly explained by the fact Popper and his criticsascribed different meanings to the term theoryThen we narrow our attention to one particular theory,namely quantum mechanics, in order to elucidate general matters discussed. We prove that quantum mechanics is irrefutable in a rather straightforward sense, but argue that it is refutable in a more sophisticated sense, which incorporates someobservations obtained by looking closely at the practiceof physics. We shall locate exactly where non-rigourous elements enter the evaluation of a scientific theory – thismakes us see clearly how fruitful mathematics isfor the philosophy of science.  相似文献   

5.
Popper has provided a model for the scientific explanation of human actions and a metaphysical theory of man which can guide scientific research. In this paper I discuss the problems of the empirical content and nomicity of the Rationality Principle and extend the method of situational analysis to the problem of explaining beliefs. The domain of applicability of the Rationality Principle is bounded on one side by cases in which behavior is determined by processes which can not be influenced by criticism and on the other side by the phenomenon of substantive creativity. However, a large part of human activity lies within its scope.  相似文献   

6.
The tu quoque argument is the argument that since in the end rationalism rests on an irrational choice of and commitment to rationality, rationalism is as irrational as any other commitment. Popper's and Polanyi's philosophies of science both accept the argument, and have on that account many similarities; yet Popper manages to remain a rationalist whereas Polanyi decided for an irrationalist version of rationalism. This is more marked in works of their respective followers, W. W. Bartley III and Thomas S. Kuhn. Bartley declares the rationalist's very openness to criticism open to criticism, in the hope of rendering Popper's critical rationalism quite comprehensive. Kuhn makes rationality depend on the existence of an accepted model for scientific research (paradigm), thus rendering Polanyi's view of the authority of scientific leadership a sine qua non for scientific progress. The question raised here is, in what sense is a rationalist committed to his rationality, or an irrationalist to his specific axiom ? The tradition views only the life‐long commitment as real. Viewing rationality as experimental open‐mindedness, we may consider a rationalist unable to retreat into any life‐long commitment — even commitment to science. In this way the logic of the tu quoque argument is made irrelevant: anyone able to face the choice between rationality and commitment is already beyond such a choice; it is one thing to be still naïve and another — and paradoxical — thing to return to one's naïveté.  相似文献   

7.
The present work takes the decease of Horst Wessel as an opportunity to present and honour his work (and that of his group), which has not received the attention it deserves. The focus will be on works which might not be sufficiently well-known. Wessel was, as we aim to show, familiar with the international debate concerning logical and philosophical issues and strived to solve them by considering theories of logical consequence, a non-traditional theory of predication and the theory of logical terms, all of which he developed in joint work with Alexander Sinowjew. Wessel had to significantly refine these theories in order to formulate his criticism towards alternative logics and to apply the theories to a treatment of intensional contexts. He was vehemently opposed to approaches which revised fundamental laws of classical logic. Questions concerning the history of Philosophy are addressed in Wessel’s criticism of the Kantian antinomies of pure reason, in Wessel’s contributions to the problem of universals and his analyses of fatalism, tychism and antifatalism.  相似文献   

8.
In a paper entitled “Revolution in Permanence”, published in the collection “Karl Popper: Philosophy and Problems”, John Worrall (1995) severely criticised several aspects of Karl Popper’s work before commenting that “I have no doubt that, given suffi-cient motivation, a case could be constructed on the basis of such remarks that Popper had a more sophisticated version of theory production......” (p. 102). Part of Worrall’s criticism is directed at a “strawpopper”: in his “Darwinian Model” emphasising the similarities and differences between genetic mutation, variation in animal behaviour and the gestation of scientific theories, Popper (1975, 1981, 1994) never stated that tentative scientific conjec-tures “while more or less random, are not completely blind.” He was referring to variation in animal species behaviour, and about tentative scientific conjectures he said nothing, although common sense would indicate that presumably he regarded them as being less blind and less random. In Popper (1977, 1983), giving a summary of his “Darwinian Model”, he repaired this omission about tentative scientific conjectures by inserting the sentence “On a level of World 3 theory formation they are of the character of planned gropings into the unknown.” Recent developments in the field of genetics (see for example Raff (1996), Lewis (1999), Korn (2002)) indicate that Popper’s intuitions were along the modern lines while Worrall’s intuitions are old fashioned. Therefore Popper’s “Darwinian Model” remains both viable and fruitful.  相似文献   

9.
John Wettersten 《Ratio》2007,20(2):219-235
All fallibilist theories may appear to be defective, because they allegedly underestimate the security of at least some scientific knowledge and thereby leave science less defensible than it otherwise might be. When they call all scientific knowledge conjectural they may seem at first blush to underestimate the superiority of science vis a vis pseudo‐science. Fallibilists apparently fail to account for the fact that science turns theory into facts, because even “facts” are held only provisionally. This impression is false: the relatively secure establishment of facts can be accounted for with a fallibilist view. After theories have been honed through sharp criticism, there is often no reason to doubt some aspects of them. These aspects are what we regard to be factual knowledge, even though these facts are also provisionally accepted as such. We then explain the newly won factual knowledge with deeper theories, which often correct our factual knowledge in spite of its apparent security. Theories of justification add nothing useful to the fallibilists' observation that science finds the best theories because it has the highest standards of criticism. Fallibilist theories today give the best account and defence of science. We may abandon the quest for some kind of assurance that goes beyond the determination that some theory can answer all known objections to it and take up more interesting problems, such as how we can find new objections and how criticism may be improved and made institutionally secure. 1 1 I am grateful to Joseph Agassi and an anonymous referee of this journal for comments on an earlier draft of this essay.
  相似文献   

10.
11.
In several places Popper describes a little experiment in which an audience is given the non-specific command Observe! He draws a number of conclusions from this experiment, in particular that observation takes place in the presence of theoretical problems, questions, hypotheses or points of view. The paper argues that while Popper's experiment is instructive, it hardly supports the strong conclusions he draws about the theory-dominance of observation in science. In particular, it is argued that talk of principles of selection which guide us to relevant observations, rather than the host of irrelevant observations of the naive inductivist, is misleading. Rather, it is the goals, aims, motives or interests of an observer that guide observation and these need not always involve a theoretical component.  相似文献   

12.
This paper intends to re-emphasize the relationship among Psychology, Popper and his Philosophy. Popper, who is often considered as one of the most important philosophers of science, had associations with the discipline of psychology in his early years. Popper was associated with Würzburg school of psychology, especially the psychologies of Külpe, Selz and Bühler. However, there was a change in Popper’s interest from the psychology of discovery to an objectivist epistemology—that is, to the logic of discovery, which he himself acknowledged (1976, p. 55). Popper, not only turned away from psychology, as early as 1930 or thereabouts, he later became one of the most outspoken opponents of a psychological approach to science. This antipathy has worked both ways! Very few psychologists study and discuss Popper. His place in the history of psychology, certainly remains inadequate. But why should psychologists ignore Popper? Instead of turning away from Popper, Psychologists’ efforts should be directed towards bringing into focus Popper, his works and association with psychology. This paper argues and tries to discuss the relationships among Popper, his training in psychology and his philosophy.  相似文献   

13.
The article discusses two puzzles about Plato's account of the democratic person: (1) unlike his account of the democratic city, his characterization of a democratic person is markedly incorrect. (2) His criticism of a person so characterized is criticism of a straw man. The article argues that the first puzzle is resolved if we see it as a result of Plato's assumption that a democratic person is a person whose soul is isomorphic to a democratic constitution. Such a person has a desire satisfaction theory of good and adopts liberty and equality of desires as a basis for action. The article then argues that Plato's criticism brings up two problems endemic to desire satisfaction theories of good, the problem of bad desires and the problem of conflicts of desires. The criticism is that the democratic person's way of dealing with these problems, by applying the social principles of liberty and equality to his desires, is irrational.  相似文献   

14.
Summary  Popper uses the “Humean challenge” as a justification for his falsificationism. It is claimed that in his basic argument he confuses two different doubts: (a) the Humean doubt (Popper’s problem of induction), and (b) the “Popperean” doubt whether – presupposing that there are laws of nature – the laws we accept are in fact valid. Popper’s alleged solution of the problem of induction does not solve the problem in a straightforward way (as Levison and Salmon have remarked before). But if Popper’s solution of the Humean challenge is re-interpreted as being close to Kant’s it makes sense. Even though Popper explicitly rejects Kant’s synthetic judgements a priori, it is claimed here that this is so because he misinterprets Kant’s argument. If he had understood Kant correctly he should have been a modern “Kantianer”!  相似文献   

15.
The deep formal and conceptual link existing between artificial life and artificial intelligence can be highlighted using conceptual tools derived by Karl Popper's evolutionary epistemology. Starting from the observation that the structure itself of an organism embodies knowledge about the environment which it is adapted to, it is possible to regard evolution as a learning process. This process is subject to the same rules indicated by Popper for the growth of scientific knowledge: causal conjectures (mutations) and successive refutations (extinction). In the field of machine learning such a paradigm is represented by genetic algorithms that, simulating biological processes, emulate cognitive processes. From a practical viewpoint, that perspective allows to identify the two different kinds of learning considered by artificial intelligence, knowledge acquisition and skill improvement, and to get a different view of the problem of heuristic knowledge in learning systems. From a theoretical point of view, these considerations can shade a new light on an old epistemological problem: why do we live in a learnable world?  相似文献   

16.
Stuart Mathieson 《Zygon》2021,56(1):254-274
The Victoria Institute was established in London in 1865. Although billed as an anti‐evolutionary organization, and stridently anti‐Darwinian in its rhetoric, it spent relatively little time debating the theory of natural selection. Instead, it served as a haven for a specific set of intellectual commitments. Most important among these was the Baconian scientific methodology, which prized empiricism and induction, and was suspicious of speculation. Darwin's use of hypotheses meant that the Victoria Institute members were unconvinced that his work was truly scientific, but even more concerning for them was the specter of biblical criticism. This approach to biblical studies incorporated techniques from literary criticism, treating it as any other document. Since it also relied on hypotheses, the Victoria Institute members were similarly skeptical that biblical criticism was scientific, and spent much of their time attempting to refute it. In this way, they functioned as an incubator for the concerns that would animate the fundamentalist–modernist controversies of the early twentieth century.  相似文献   

17.
In order to develop an account of scientific rationality, two problems need to be addressed: (i) how to make sense of episodes of theory change in science where the lack of a cumulative development is found, and (ii) how to accommodate cases of scientific change where lack of consistency is involved. In this paper, we sketch a model of scientific rationality that accommodates both problems. We first provide a framework within which it is possible to make sense of scientific revolutions, but which still preserves some (partial) relations between old and new theories. The existence of these relations help to explain why the break between different theories is never too radical as to make it impossible for one to interpret the process in perfectly rational terms. We then defend the view that if scientific theories are taken to be quasi-true, and if the underlying logic is paraconsistent, it’s perfectly rational for scientists and mathematicians to entertain inconsistent theories without triviality. As a result, as opposed to what is demanded by traditional approaches to rationality, it’s not irrational to entertain inconsistent theories. Finally, we conclude the paper by arguing that the view advanced here provides a new way of thinking about the foundations of science. In particular, it extends in important respects both coherentist and foundationalist approaches to knowledge, without the troubles that plague traditional views of scientific rationality.  相似文献   

18.
Sir Karl Popper has claimed that behaviorism is misguided because it holds that conditioning occurs through repetition. According to Popper, there is no such thing as learning through repetition. To the limited extent that philosophers of science have concerned themselves with behaviorism, this attack is one of the most direct and unique in that the battleground is not over the value of mentalism/cognitivism but a bold claim that conditioning—the heart and soul of behaviorism—is fictitious. This paper examines the soundness of Popper's argument against behaviorism by examining whether operant and classical conditioning rely on learning through repetition and suffer from other problems Popper attributes to the notion of conditioning. Although Popper correctly attributes certain properties to classical conditioning, he fails to undermine the empirical evidence that such conditioning occurs. Second, we claim that although B. F. Skinner is never entirely clear, operant conditioning does not rely on repetition nor does it suffer from the other problems Popper attributes to conditioning. Thus, Popper's argument also fails because of his assumption that all conditioning is classical conditioning and therefore his misattributing properties of classical conditioning to operant conditioning. We conjecture that the earlier polemics of John Watson and Watson's sole reliance on classical conditioning probably contributed to Popper's confusion on this point.  相似文献   

19.
Conclusion From a philosophical standpoint, the work presented here is based on van Fraassen [26]. The bulk of that paper is organized around a series of arguments against the assumption, built into standard deontic logic, that moral dilemmas are impossible; and van Fraassen only briefly sketches his alternative approach. His paper ends with the conclusion that the problem of possibly irresolvable moral conflict reveals serious flaws in the philosophical and semantic foundations of orthodox deontic logic, but also suggests a rich set of new problems and methods for such logic. My goal has been to suggest that some of these methods might be found in current research on nonmonotonic reasoning, and that some of the problems may have been confronted there as well.I have shown that nonmonotonic logics provide a natural framework for reasoning about moral dilemmas, perhaps even more useful than the ordinary modal framework, and that the issues surrounding the treatment of exceptional information within these logics run parallel to some of the problems posed by conditional oughts. However, there is also another way in which deontic logic might benefit from a connection to nonmonotonic reasoning. A familiar criticism among ethicists of work in deontic logic is that it is too abstract, and too far removed from the kind of problems confronted by real agents in moral deliberation. It must be said that similar criticisms of abstraction and irrelevance are often lodged against work in nonmonotonic reasoning by more practically minded researchers in artificial intelligence; but here, at least, the criticisms are taken seriously. Nonmonotonic logic aims at a qualitative account of commonsense reasoning, which can be used to relate planning and action to defeasible goals and beliefs; and at least some of the theories developed in this area have been tested in realistic situations. By linking the subject of deontic logic to this research, it may be possible also to relate the idealized study of moral reasoning typical of the field to a more robust treatment of practical deliberation.  相似文献   

20.
Grünbaum's approach to psychoanalysis suffers from several difficulties. It imposes a standard of logical reductionism and methodological purity that not only violates the nature of psychoanalytic knowledge, but imposes an invalid standard of verification and scientific confirmation. It utilizes a brand of dichotomous reasoning that forces psychoanalytic propositions into artificial positions that do not reflect the actuality of analytic practice. It imposes a standard of verification that is impossible for psychoanalysis, along with all forms of psychological knowledge, to reach. It visualizes psychoanalysis as encompassing only one form of knowledge of human psychic life, forcing it into a model that eliminates other aspects of the psychoanalytic process, so that psychoanalysis is subjected to criticism only on one dimension among several--a kind of psychoanalytic straw man. The psychoanalysis that is so impaled often is difficult for the psychoanalytic practitioner to recognize. To the extent that Grünbaum's skillful and highly informed criticism of the philosophical bases of psychoanalysis encounters these difficulties, the value of his argument falls short of providing a useful basis for advancing psychoanalytic knowledge and particularly for promoting the quest for pertinent standards of validation within psychoanalysis.  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