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 31 毫秒
1.
A cognitive autopoietic system is a dynamic, self-generating, organized and self-organizing thing which self-regulates (by internal rearrangement) with respect to an external medium. The present model of the effect of stress on a cognitive autopoietic (ESCA) system captures the notion of how a priori cognitive structures (categories), combined with external sensations, constitute the basis for the development of cognitive structures (CS) and their architecture. The ESCA model integrates the fact that the mind–environment relation has a twofold effect: on one hand, it enables self-regulation of mind (the matching of external sensations with CS), but on the other hand, it poses a potential perturbation on the same, which may result in the breakdown of the self-regulation of mind. The architecture of the CS developed on the basis of the ESCA model is consistent with the manifestations of the effect of stress on mind behavior at different levels. The ESCA model predicts that the faculty to concatenate synthetic propositions, which enables enhanced categorical conscious cognition (ECCC) on the basis of CS, is inhibited by stress, thus reducing cognition to a mechanized heuristic categorical conscious cognition (HCCC) and/or an unconscious cognition (UC) level. The ESCA model explains the casual relation between cognition of persistent social stress and self-esteem, sensory deprivation and self-cognition, a mechanized mind state and accumulated stress, and the effect of stress activated short-term emotions on cognition. Finally, it is speculated how persistently perceived uncertainty may induce stress.  相似文献   

2.
This paper argues that nearly every proposition could in principle be known a priori, with exceptions for necessary falsehoods and a few other odd cases. The argument for this conclusion has two main premises: (i) Odd exceptions aside, if it is possible that p, then it is possible that someone knows innately that p. (ii) Necessarily, whatever is known innately is known a priori. After defending (i) and (ii), I conclude by suggesting that the best way to recover a reasonably limited and interesting conception of the a priori is to adopt an anthropocentric conception of a priori knowability, one that is relativized to our own innate cognitive capacities. However, this proposal has consequences that contradict prevailing views about the a priori. More importantly, this proposal has the result that many philosophical applications of the notion of apriority are misguided.  相似文献   

3.
Sandra Lapointe 《Synthese》2010,174(2):263-281
This paper is aimed at understanding one central aspect of Bolzano’s views on deductive knowledge: what it means for a proposition and for a term to be known a priori. I argue that, for Bolzano, a priori knowledge is knowledge by virtue of meaning and that Bolzano has substantial views about meaning and what it is to know the latter. In particular, Bolzano believes that meaning is determined by implicit definition, i.e. the fundamental propositions in a deductive system. I go into some detail in presenting and discussing Bolzano’s views on grounding, a priori knowledge and implicit definition. I explain why other aspects of Bolzano’s theory and, in particular, his peculiar understanding of analyticity and the related notion of Ableitbarkeit might, as it has invariably in the past, mislead one to believe that Bolzano lacks a significant account of a priori knowledge. Throughout the paper, I point out to the ways in which, in this respect, Bolzano’s antagonistic relationship to Kant directly shaped his own views.  相似文献   

4.
Jessica F. Leech 《Ratio》2010,23(2):168-183
Williamson (1986) presents a troublesome example of the contingent a priori; troublesome, because it does not involve indexicals, and hence cannot be defused via the usual two‐dimensional strategies. Here I explore how the example works, via an examination of crucial belief‐forming method M, partly in response to Hawthorne (2002) and the questions there raised for ‘hyperreliable’ belief‐forming methods. I suggest that, when used to form a belief, M does its special work through creating a verifying state of affairs which guarantees the truth of the belief thus formed. This creative link can be said to account for the knowledge‐conferring status of M. But it also provides us with a way to defuse the purported example of the contingent a priori. The knowledge at issue is only a priori in virtue of this creative link, an importantly different epistemic achievement from standard cases of a priori knowledge. One important moral to be drawn is that the a priori/a posteriori distinction does not appear to be slicing the epistemological beast at its joints.  相似文献   

5.
Conclusions At the outset of this discussion, I undertook to present an argument from design which would follow Swinburne's example in making use of a priori judgments, while avoiding some of the objections which have been posed in response to his treatment of these issues. So we need to ask: how does this approach to the question of design compare with Swinburne's?Swinburne argues that a chaotic world is a priori more likely than an ordered world: this consideration provides one central reason, on his account, for giving an explanation of some sort for the world's regularity. The other central argument he advances for this claim is the argument from analogy (in terms of the coins) which we noted earlier. The approach I have taken offers an alternative route to this same conclusion. In particular, it substitutes the simpler a priori judgments recorded in (i) and (ii) for the rather difficult and contentious claim that chaos is a priori more likely than order. In place of this claim, I have offered the judgment that order, or recurrence, is more likely given the activity of a common source or common kind of source than otherwise: this proposal does not commit us to a view either way on the question of whether order is a priori likely per se. Moreover, in place of Swinburne's analogical argument, I have offered an a priori approach, with the advantages I have noted.Given that recurrence is to be explained, we might ask: why offer an explanation in terms of design? On this point, Swinburne argues, for instance, that no other explanation of temporal regularity is even possible a priori. Again, the a priori principles which I have used, in (iv) and (v), may be less ambitious, but at the same time more persuasive. In support of this same idea, Swinburne also cites various ideas to do with the predictive power of the idea of design. I have tried to bring out the role of this sort of consideration in terms of my principle (v). Principle (iv) has no place in Swinburne's account, in view of his reliance on the principle of simplicity as a measure of prior probability.Lastly, we may ask: if we are to cite a designer, are there reasons for attributing to this agent more powers than are needed for the production of the effect to be explained? On this point, Swinburne cites the principle of simplicity. Again, my approach avoids what has proved to be a relatively controversial judgment about the nature of a priori probabilities, offering in place of the principle of simplicity the less ambitious principle recorded in (iii). At this point, I have moreover inverted the logical sequence of Swinburne's argument: it seems to me that, in the ways I have indicated, it is helpful to consider the extent of the powers of the source of recurrence before addressing the question of design.In these various ways, I hope I have made good my undertaking to present an argument which avoids some of the controversy surrounding the particular measures of a priori probability which figure in Swinburne's argument. Moreover, I hope that this approach provides an indication of how a priori judgments may function in a relatively unproblematic way within an argument from design, in so far as (i)–(v) are all rather modest proposals. In sum, the argument I have presented is distinguished by its explicit use of the a priori judgments recorded in (i)–(v), by its attempt to buttress in this fashion analogical forms of argument, and by the logical role it gives to the idea that the source of regularity possesses more powers than are required for the production of this effect.Lastly, we might ask: how persuasive is this argument? Of course, the cogency of the idea of design depends upon the balance of debate in other areas of the philosophy of religion, especially upon our ability to provide some account of the existence of evil. In this paper, I have been concerned to argue simply that recurrence by kind provides evidence for design: I have not addressed the question of whether other features of the world provide good evidence against the idea of (benevolent) design. However, if we confine our attention to this one phenomenon, there is it seems to me good evidence for the idea of design, (i) and (ii) suggest that recurrence surely calls for some explanation; (iii), together with the existence in nature of statistical irregularities, suggests that whatever provides this explanation could have brought about other effects besides; and design seems the only clear explanation of why this effect should have been brought about, if (as I have argued) analogies drawn from vegetable and animal reproduction fail, and if we cannot explain the effect satisfactorily by reference to the conditions of observation. Moreover, I have argued that there are reasons for supposing that the probability a priori of design is relatively high in relation to the probability a priori of any rival hypothesis of equivalent predictive power. In brief, this is because the design hypothesis (unlike the hypothesis of theism) can cite an agent of relatively indeterminate power in order to account for the phenomenon to be explained. In this regard, it is less precisely defined than any rival hypothesis of equivalent predictive power. If all of this is so, then as philosophers from early times have supposed, temporal regularity provides the basis for a powerful argument in favour of design. It remains true, of course, that its import can be judged in full only when we have taken into account the relevance of other phenomena, many of which are apparently less favourable to the idea of design.  相似文献   

6.
A recurrent idea in the history of psychology is that one is conscious of outputs but not of the complex processes underlying the generation of outputs, which is evident in the out-of-the-blue, “eureka-like” experiences associated with intuition. We examine how this idea may suffer from a logical fallacy and may thus have inadvertently hindered progress on the study of the intimate liaisons among high-level central processes, intuition, and overt action. It is proposed that, for various reasons, the only undisputable output in the nervous system is overt action. Once this is accepted, the overlooked relationship between conscious central processes and overt action can be examined. A review of the evidence reveals that conscious processing is in the business of, not low-level perceptual processing, motor control, or action production per se, but of constraining a peculiar form of knowledge-based, integrated action-goal selection, which can lead to integrated actions such as holding one's breath. Unconscious processing can influence behavior indirectly, by producing these conscious constraining dimensions that modulate action-goal selection, or directly, through unintegrated actions such as reflexively inhaling or responding to a subliminal stimulus. From this standpoint, eureka-like intuitions reflect not an atypical brain process but the general nature by which unconscious machinations influence action either directly or indirectly, through the limited purview of consciousness.  相似文献   

7.
Conclusion Some of Tichý's conclusions rest on an assumption about substitutivity which Kripke would not accept. If we grant the assumption, then Tichý successfully shows that we can discover true identity statements involving names a priori, but not that we can discover a priori what properties things have essentially. Many of Tichý's arguments require an implausible rejection of the possibility of indirect belief as described in Section III. 25Are there necessary a posteriori propositions? I have argued that we certainly can discover necessary propositions a posteriori, but have left it an open question whether there are necessary propositions which we can only discover a posteriori.What effect do the considerations here presented have on the positivist doctrine that the a priori and the necessary coincide? My explanation of how we discover necessary propositions a posteriori involves our believing them indirectly, in virtue of believing contingent propositions. I would argue that Kripke's examples of the contingent a priori involve, similarly, our believing the contingent propositions in directly, in virtue of believing necessary propositions.This suggests that a reformulation of the positivist thesis along something like the following lines may well be correct. Let us say that someone directly believes a proposition just in case he could not fail to believe it without being in a different cognitive state. Then perhaps one can directly believe a proposition on the basis of a priori evidence only if it is necessary, and can directly believe a proposition on the basis of a posteriori evidence only if it is contingent.  相似文献   

8.
The paper presents the two main assumptions of Attentional Semantics—(A) and (B), and its main aim (C). (A) Conscious experience is determined by attention: there cannot be consciousness without attention. Consciousness is explained as the product of attentional activity. Attentional activity can be performed thanks to a special kind of energy: nervous energy. This energy is supplied by the organ of attention. When we perform attentional activity, we use our nervous energy. This activity directly affects the organ of attention, causing a variation in the state of the nervous energy. This variation constitutes the phenomenal aspect of consciousness. (B) Words are tools to pilot attention. The meanings of words isolate, de-contextualize, “freeze” and classify in an articulated system the ever-changing and multiform stream of our conscious experiences. Each meaning is composed of the sequence of invariable elements that, independently of any individual occurrence of a given conscious experience, are responsible for the production of any instance of that conscious experience. The elements composing the meanings of words are attentional operations: each word conveys the condensed instructions on the attentional operations one has to perform if one wants to consciously experience what is expressed through and by it. (C) Attentional Semantics aims at finding the attentional instruction conveyed by the meanings of words. To achieve this goal, it tries: (1) to identify the sequence of the elementary conscious experiences that invariably accompany, and are prompted by, the use of the word being analyzed; (2) to describe these conscious experiences in terms of the attentional operations that are responsible for their production; and (3) to identify the unconscious and non-conscious operations that, directly or indirectly, serve either as the support that makes it possible for the attentional operations to take place, be completed, and occur in a certain way, or as the necessary complement that makes it possible to execute and implement the activities determined and triggered by the conscious experiences. The origins of Attentional Semantics are also presented, and the methodological problems researchers encounter when analyzing meanings in attentional terms are discussed. Finally, a brief comparison with the other kinds of semantics is made.  相似文献   

9.
Tyler Burge notably offers a truth‐first account of perceptual entitlement in terms of a priori necessary representational functions and norms: on his account, epistemic normativity turns on natural norms, which turn on representational functions. This paper has two aims: first, it criticises Tyler Burge's truth‐first a priori derivation on functionalist and value‐theoretic grounds. Second, it develops a novel, knowledge‐first a priori derivation of perceptual entitlement. According to the view developed here, it is a priori that we are entitled to believe the deliverances of our perceptual belief formation system, in virtue of the latter's constitutive function of generating knowledge.  相似文献   

10.
Spinoza scholars have claimed that we are faced with a dilemma: either Spinoza's definitions in his Ethics are real, in spite of indications to the contrary, or the definitions are nominal and the propositions derived from them are false. I argue that Spinoza did not recognize the distinction between real and nominal definitions. Rather, Spinoza classified definitions according to whether they require a priori or a posteriori justification, which is a classification distinct from either the real/nominal or the intensional/extensional classification. I argue that Spinoza uses both a priori and a posteriori definitions in the Ethics and that recognizing both types of definitions allows us to understand Spinoza's geometric method in a new way. We can now understand the geometric method as two methods, one resulting in propositions that Spinoza considers to be absolutely certain and another resulting in propositions that Spinoza does not consider certain. The latter method makes use of a posteriori definitions and postulates, whereas the former method uses only a priori definitions and axioms.  相似文献   

11.
In reply to Michael Bertrand, I clarify my view that the problem of physical evil is not an a priori problem but an a posteriori one.  相似文献   

12.
We provide an account of necessary a posteriori identity statements that relies upon Perry's multipropositionalism. On our account an utterance of, e.g., ‘Hesperus is Phosphorus', semantically makes available several propositions, one of which is necessary (and a priori) and another of which is a posteriori (and contingent). Since our view resembles two‐dimensionalism, one might assume that it is undermined by the sorts of nesting arguments that Soames and others have raised against two‐dimensionalism. We demonstrate, however, that our account is immune to such nesting arguments.  相似文献   

13.
In the Bayesian framework, a language learner should seek a grammar that explains observed data well and is also a priori probable. This paper proposes such a measure of prior probability. Indeed it develops a full statistical framework for lexicalized syntax. The learner's job is to discover the system of probabilistic transformations (often called lexical redundancy rules) that underlies the patterns of regular and irregular syntactic constructions listed in the lexicon. Specifically, the learner discovers what transformations apply in the language, how often they apply, and in what contexts. It considers simpler systems of transformations to be more probable a priori. Experiments show that the learned transformations are more effective than previous statistical models at predicting the probabilities of lexical entries, especially those for which the learner had no direct evidence.  相似文献   

14.
When asked in a questionnaire to describe a spiritual person, William James named one instead: Phillips Brooks. This article focuses on Brooks—his life, his sermons, and his poem “O Little Town of Bethlehem”—to make the case that he exemplified James’ view of spirituality as “a susceptibility to ideals, but with a certain freedom to indulge in imagination about them.” It also supports Belzen’s (Mental Health, Religion & Culture, 12:205–222, 2009) view that there is no spirituality in general but only individual manifestations of it, a point that James’ nomination of Brooks implicitly supports.  相似文献   

15.
According to Epistemic Two-Dimensional Semantics (E2D), expressions have a counterfactual intension and an epistemic intension. Epistemic intensions reflect cognitive significance such that sentences with necessary epistemic intensions are a priori. We defend E2D against an influential line of criticism: arguments from epistemic misclassification. We focus in particular on the arguments of Speaks [2010] and Schroeter [2005]. Such arguments conclude that E2D is mistaken from (i) the claim that E2D is committed to classifying certain sentences as a priori, and (ii) the claim that such sentences are a posteriori. We aim to show that these arguments are unsuccessful as (i) and (ii) undercut each other. One must distinguish the general framework of E2D from a specific implementation of it. The framework is flexible enough to avoid commitment to the apriority of any particular sentence; only specific implementations are so committed. Arguments from epistemic misclassification are therefore better understood as arguments for favouring one implementation of E2D over another, rather than as refutations of E2D.  相似文献   

16.
In this journal Andy Egan argued that, contrary to what I have claimed, quasi-realism is committed to a damaging asymmetry between the way a subject regards himself and the way he regards others. In particular, a subject must believe it to be a priori that if something is one of his stable or fundamental beliefs, then it is true. Whereas he will not hold that this is a priori true of other people. In this paper I rebut Egan's argument, and give further consideration to the correct way to think about our own fallibility.  相似文献   

17.
The sharing of bodily states elicits in mimicker and mimickee corresponding conceptualisations, which facilitates liking. There are many studies showing the relatedness of mimicry and liking. However, the mimicry‐liking link has not been investigated under conditions in which the mimickee is liked or disliked a priori. In two studies, we examined moderating effects of a priori liking on the mimicry‐liking link. Liking was measured via self‐report measures (Studies 1 and 2) and behavioural measures using a virtual environment technology (Study 2). Results showed that when participants intentionally mimicked a disliked person, liking for that person was not improved, whereas when participants mimicked a liked person, liking for that person increased. These effects were shown to be mediated by affiliation. These studies not only provided further evidence of a link between mimicry and liking, but also demonstrated that this relationship is moderated by a priori liking. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
During the last decades it has become widely accepted that scientific observations are ‘theory‐laden’. Scientists ‘see’ the world with their theories or theoretical presuppositions. In the present paper it is argued that they ‘see’ with their scientific instruments as well, as the uses of scientific instruments is an important characteristic of modern natural science. It is further argued that Euclidean geometry is intimately linked to technology, and hence that it plays a fundamental part in the construction and operation of scientific instruments. Finally, Euclidean geometry is compared to fractal geometry, and the question of its a priori status is raised. Although the position that Euclidean geometry is a priori in the original Kantian sense is untenable, the paper concludes that in some restricted sense Euclidean geometry may be said to be a priori.  相似文献   

19.
SOCIAL JUSTICE     
Social justice (which includes retributive and distributive justice) is most clearly satisfied by a system of Divine rewards and punishments: an omnipotent, omniscient, perfectly just Being could determine in each case how much effort was made and effect the appropriate distribution of rewards and punishments. A correct understanding of social justice naturally leads us to suppose that there is an afterlife, a God, a free choice — though it is logically possible at least that social justice could be satisfied in some future (very advanced) human society. There will still be those who have their doubts about the correctness of any view according to which justice cannot be attained by fallible creatures who have an incomplete knowledge of one another's behaviour. But, surely, these doubts are not sufficient to discredit my view. There is no a priori reason for rejecting such a view. There is nothing about our use of the term ‘justice’ and its cognates which implies that such a view is mistaken. (Otherwise the statement “There is no justice in this world’ would be meaningless.) To the contrary, there are widely held religious views, Christian as well as non-Christian, which take this view quite seriously. If there is no a priori reason for rejecting this view, then there must be some independent reason for rejecting it. In other words, we need some independent reasons for believing that social justice can be attained by fallible creatures with limited knowledge. The mere fact that we might feel uncomfortable with my theory is not reason enough to reject it. Finally, those who do experience this discomfort might ask themselves whether such discomfort stems from their moral experience or whether they are simply intent on finding justice in imperfect human institutions.  相似文献   

20.
I reconstruct and critique two arguments Laurence BonJour has recently offered against skepticism about the a priori. While the arguments may provide anti-skeptical, internalist foundationalists with reason to accept the a priori, I show that neither argument provides sufficient reason for believing the more general conclusion that there is no rational alternative to accepting the a priori.  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