June 25, 2010

-page 65-

The development of such ‘externalist’ conceptions of intentionality informs the reception of Russell's legacy in contemporary philosophy of mind as well. Russell also helped to put in play a conception of the intentionality of mental states, according to which each such state is seen as involving the individual's ‘acquaintance with a proposition’ (counterpart to Frégean ‘grasping’) - which proposition is at once both what is understood in understanding expressions by which the state of mind is reported, and the content of the individual's state of mind. Thus, intentional states are ‘propositional attitudes.’ Also importantly, Russell's famous analysis of definite descriptions into phrases employing existential quantifiers and general predicates underlay many subsequent philosophers' rejection of any conception of intentionality (like Meinong's) that sees in it a relation to non-existent objects. And, Russell's treatment drew attention to cases of what he called ‘logically proper names’ that apparently defies such analysis in descriptive terms (paradigmatically, the terms ‘this’ and ‘that’), and which (he thought) thus must refer ‘directly’ to objects. Reflection on such ‘demonstratives’ and ‘indexical’ (e.g., ‘I,’ ‘here,’ ‘now’) reference has led some to maintain that the content of our states of mind cannot always be constituted by Frégean senses but must be seen as consisting partly of the very objects in the world outside our heads to which we refer, demonstratively, indexically - another source of support for an ‘externalist’ view of mental content, hence, of intentionality.


Yet another important source of externalist proclivities in twentieth century philosophy lies in the thought that the meaningfulness of a speaker's utterances depends on its potential intelligibility to hearers: language must be public - an idea that has found varying and influential expression in the work of Ludwig Wittgenstein, W.V.O. Quine, and Donald Davidson. This, coupled with the assumption that intentionality (or ‘thought’ in the broad (Cartesian) sense) must be expressible in language, has led some to conclude that what determines the content of one's mind must lie in the external conditions that enable others to attribute content.

However, the movement from Frége and Russell toward externalist views of intentionality should not simply be accepted as yielding a fund of established results: it has been subject to powerful and detailed challenges, but without plunging into the details of the internalism/externalism debate about mental content, we can recognize, in the issues just raised, certain themes bearing particularly on the connection between consciousness and intentionality.

For example: it is sometimes assumed that, whatever may be true of content or intentionality, the phenomenal character of one's experience, at least, is ‘fixed internally’ -, i.e., it involves no necessary relations to the nature of particular substances in one's external environment or to one's linguistic community. But then the purported externalist finding that meaning nor contents are ‘in the head’ and, of course, be read as showing the insufficiency of phenomenal consciousness to determine any intentionality or content. Something like this consequence is drawn by Putnam (1981), who takes the stream of consciousness to comprise nothing more than sensations and images, which (as Frége saw) should be sharply distinguished from thought and meaning. This interpretation of the import of externalist arguments may be reinforced by a tendency to tie (phenomenal) consciousness to non-intentional sensations, sensory qualities, or ‘raw feels,’ and hence to dissociate consciousness from intentionality (and allied notions of meaning and reference), a tendency that has been prominent in the analytic tradition.

But it is not at all evident that externalist theories of content require us to estrange consciousness from intentionality. One might argue (as do Martin Davies (1997) and Fred Dretske (1997)) that in certain relevant respects the phenomenal character of experience is also essentially determined by causal environmental connections. By contrast, one may argue (as do Ludwig (1996b) and Horgan and Tienson (2002)) that since it is conceivable that a subject has experience is much like our own in phenomenal character, but radically different in external causes from what we take our own to be (in the extreme case, a mind bewitched by a Cartesian demon into massive hallucination), there must indeed be a realm of mental content that is not externally determined.

One other aspect of the Frége-Russell tradition of theorizing about content that impinges on the consciousness/intentionality connection is this. If ‘content’ is identified with the sense or the truth-condition determiners of the expressions used in the object-clause reporting intentional states of mind, it will seem natural to suppose that possession of mental content requires the possession of conceptual capacities of the sort involved in linguistic understanding - ‘grasping senses.’ But then, to the extent the phenomenal character of experience is inadequate to endow a creature with such capacities, it may seem that phenomenal consciousness has little to do with intentionality.

However, this raises large issues. One is this: it should not be granted without question that the phenomenal character of our experience could be as it is in the absence to the sorts of conceptual capacities sufficient for (at least some types of) intentionality. And this is tied to the issue of whether or not the phenomenal character of experience is (as some suppose) a purely sensory affair. Some would maintain, on the contrary, that thought (not just imagistic, but conceptual thought) has phenomenal character too. If so, then it is very far from clear that phenomenal character can be divorced from whatever conceptual capacities are necessary for intentionality.

Moreover, we may ask: Are concepts, properly speaking, always necessary for intentionality anyway? Here another issue rears its head: is there not perhaps a form of sensory intentionality, which does not require anything as distinctively intellectual or conceptual as is needed for the grasping of linguistic senses or propositions? (This presumably would be a kind of intentionality had by the pre-linguistic (e.g., babies) or by non-linguistic creatures (e.g., dogs).) Suppose that there is, and that this type of intentionality is inseparable from the phenomenal character of perceptual experience. Then, even if one assumes that such phenomenal consciousness is insufficient to guarantee the possession of concepts, it would be wrong to say that it has little to do with intentionality. (Advocates of varying versions of the idea that there is a distinctively ‘non-conceptual’ kind of content include Bermudas 1998, Crane 1992, Evans 1982, Peacocke 1992, and Tye 1995 - for a notable voice of opposition to this trend, see McDowell 1994.) A deep difficulty in assessing these debates lies in getting an acceptable conception of concepts with which to work. We need to understand clearly what ‘having a concept of F’ does and does not require, before we can be clear about the content of and justification for the thesis of non-conceptual content.

These proposals about non-conceptual content bear some affinity with aspects of the Phenomenological tradition eluded too earlier: Husserl's notion of ‘pre-predicative’ experience as to Heidegger's procedures of ‘ready-to-hand;’ and Merleau-Ponty's idea that in normal active perception we are conscious of place, not via a determinate ‘representation’ of it, but rather, relative to our capacities for goal-directed bodily behaviour. Though to see the extent to which any of these are ‘non-conceptual’ in character would require not only more clarity about the conceptual/non-conceptual contrast, save that a considerable novel exegesis of these philosophers' works.

Also, one may plausibly try to find an affinity between externalist views in analytic philosophy, and the later phenomenologists' rejection of Husserl's reduction, based on their doubt that we can prise consciousness off from the world at which it is directed, and study its ‘intentional essence’ in solipsistic isolation. But if externalism can be defined broadly enough to encompass Heidegger, Merleau-Ponty, Kripke, and Burge, still the comparison is strained when we take account of the different sources of ‘externalism’ in the phenomenologists. These have to do it seems (very roughly) with the idea that the way we are conscious of things (or at least, for Heidegger, the way they ‘show themselves’ to us) in our everyday activity cannot be quite generally separated from our actual engagement with entities of which we are thus conscious (which show themselves in this way). Also relevant is the idea that one's use of language (hence one's capacity for thought) requires gearing one's activity to a social world or cultural tradition, in which antecedently employed linguistic meaning is taken up and made one's own through one's relation with others. All this is supposed to make it infeasible to study the nature of intentionality by globally uprooting, in thought, the connection of experience with one's spatial surroundings (and - crucially for Merleau-Ponty - one's own body), and one's social environment. Whatever the merits of this line of thought, we should note: Neither a causal connection with ‘natural kinds’ unmediated by reference-determining ‘modes of presentation,’ nor deference to the linguistic usage of specialists, nor belief in the need to reconstruct speakers’ meaning from observed behaviour, plays a role in the phenomenologists' doubts about the reduction.

The arduous exegesis required for a clearer and more detailed comparison of these views is not possible here. Nevertheless, following some of the main lines of thought in treatments of intentionality, descending on the one hand, primarily from Brentano and Husserl, and on the other, from Frége and Russell, certain fundamental issues concerning its relationship to consciousness have emerged. These include, first, the connection between consciousness and self-directed and self-reflexive intentionality. (It has already been seen that this topic preoccupied Brentano, Husserl and Sartre; its emergence as an important issue in analytic philosophy of mind will become more evident below, Second, there is concern with the way in which (and the extent to which) mind is world-involving. (In the Phenomenological tradition this can be seen in controversy over Husserl's Phenomenological reduction; That within Frégean cognitive traditions are exhibited through some formal critique as drawn upon sensationalism, in which only internalism/externalism are argued traditionally atypically in the passage through which are formally debated. Third, there is the putative distinction between conceptual and theoretical, and sensory or practical forms of intentionality. (In phenomenology this shows up in Husserl's contrast between judgment and pre-predicative experience, and related notions of his successors; In analytic philosophy this shows up in the (more recent) attention to the notion of ‘non-conceptual’ content.)

For more clarity regarding the consciousness-intentionality relationship and how these three topics figure prominently in views about it, it is necessary now to turn attention back to philosophical disagreements regarding consciousness that abruptly have of an abounding easy to each separate relation, til their distinctions have of occurring.

Consider the proposal that sense experience manifests a kind of intentionality distinct from and more basic than that involved in propositional thought and conceptual understanding. This might help form the basis for an account of consciousness. Perhaps conscious states of mind are distinguished partly by their possession of a type of content proper to the sensory subdivision of mind.

One source of the idea that a difference in type of content helps constitute a distinction between what is and is not phenomenally conscious, lies in the apparent distinction between sense experience and judgment. To have conscious visual experience of a stimulus - for it to look some way to you - is one thing. To make judgments about it is something else. (This seems evident in the persistence of a visual illusion, even once one has become convinced of the error.) However, on some accounts of consciousness, this distinction itself is doubtful, since conscious sense experience is taken to be nothing more than a form of judging. However, such to this view is expressed by Daniel Dummett (1991), who takes the relevant form of judging to consist in one's possession of information or mental content available to the appropriate sort of ‘probes’ - the availability of content he calls ‘cerebral celebrity.’ For Dummett what distinguishes conscious states of mind is not their possession of a distinctive type of intentional content, but rather the richness of that content, and its availability to the appropriate sort of cognitive operations. (Since the relevant class of operations is not sharply defined, neither, for Dummett, is the difference between which states of mind are conscious and which are not.)

Recent accounts of consciousness that, by contrast, give central place to a distinction between (conceptual) judgment and (non-conceptual - but still intentional) sense-experience includes Michael Tye's (1995) theory, holding that it is (by metaphysical necessity) sufficient to have a conscious sense-perception that some representation of sensory stimuli is formed in one's head, ‘map-like’ in character, whose (‘non-conceptual’) content is ‘poised’ to affect one's (conceptual) beliefs. This form of mental representation Tye would contrast with the ‘sentential’ form proper to belief and judgment - and in that way, he might preserve the judgment/experience contrast as Dummett does not. Consider also Fred Dretske's (1995) view, that phenomenally conscious sensory intentionality consists in a kind of mental representation whose content is bestowed through a naturally selected ‘function to indicate.’ Such natural (evolution-implanted) sensory representation can arise independently of learning (unlike the more conceptual, language dependent sort), and is found widely distributed among evolved lives.

Both Tye and Dretske's views of consciousness (unlike Dummett's) make crucial use of a contrast between the types of intentionality proper to sense-experience, and that proper to linguistically expressed judgment. On the other hand, there is also some similarity among the theories, which can be brought out by noting a criticism of Dummett's view, analogues of which arise for Tye and Dretske's views as well.

Some might think Dummett's account concerns only some variety of what Form would call ‘ascensive consciousness’. For on Dummett's account, it seems, to speak of visual consciousness is to speak of nothing over and above the sort of availability of informational content that is evinced in unprompted verbal discriminations of visual stimuli. And this view has been criticized for neglecting phenomenal consciousness. It seems we may conceive of a capacity for spontaneous judgment triggered by and responsive to visual stimuli, which would occur in the absence of the judger's phenomenally conscious visual experience of the stimuli: The stimuli do not look in any way impulsively subjective, and yet they trigger accurate judgments about their presence. The notion of such a (hypothetical) form of ‘blind-sight’ may be elaborated in such a way that we conceive of the judgment it affords for being at least as finely discriminatory (and as fine in informational content) as that enjoyed by those with extremely poor, blurry and un-acute conscious visual experience (as in the ‘legally blind’). But a view like Dummett's seems to make this scenario inconceivable.

However, this kind of criticism does not concern only those theories that would elide any experience/judgment distinction. For Tye and Dretske's theories, though they depend on forms of that contrast (and are offered as theories of phenomenal consciousness), can raise similar concerns. For one might think that the hypothetical blind-sighted would be as rightly regarded as having Tye ‘support’ some map-like representations in her visual system as would be someone with a comparable form of conscious vision. And one might find it unclear why we should think the visual system of such a blind-sighted must be performing naturally endowed indicating functions more poorly than the visual system of a consciously sighted subject would.

Whatever the cogency of these concerns, one should note their distinctness from the issues about ‘kinds of intentionality’ that appear to separate both Tye and Dretske from Dummett. The notion that there is a fundamental distinction to be drawn in kinds of intentional content (separating the more intellectual from the more sensory departments of mind) sometimes forms the basis of an account of consciousness (as with Dretske and Tye's, though not with Dummett's). But it is also important to recognize what unites Dummett, Tye, and Dretske. Despite their differences, all propose to account for consciousness by starting with a general understanding of intentionality (or mental content or representation) to which consciousness is inessential. Dummett is known for an uncompromising re-evaluation of the Western tradition, viewing writings before the rise of anaclitic philosophy as fatty and flawed by having take epistemology to be fundamental, whereas the correct approach, giving a foundational place to a concern with language, only took to a point-start with the work of Frége. Equally, the supposedly pure investigation of language in the 20th century has often kept some dubious epistemological and metaphysical company.

They then offer to explain consciousness as a special case of intentionality thus understood - so, in terms of the operations the content is available for, or the form in which it is represented, or the nature of its external source. The blind-sight-based objection to Dennett, and its possible extension to Dretske and Tye, helps bring this commonality to light. The last of these issues showed how some theories purport to account for consciousness on the basis of intentionality, in a way that focuses attention on attempts to discern a distinctively sensory type of intentionality. A different strategy for explaining consciousness via intentionality highlights the importance of clarity regarding the connection between consciousness and reflexivity. On such a view (roughly): Experiences or states of mind are conscious just insofar as the mind represents itself as having them.

In David Rosenthal's variant of this approach, a state is conscious just when it is a kind of (potentially non-conscious) mental state one has, which one (seemingly without inference) thinks that one is in. A theory of this sort starts with some way of classifying mental states that is supposed to apply to conscious and non-conscious states of mind alike. The proposal then is that such a state is conscious just when it belongs to one of those mental kinds, and the (‘higher order’) thought occurs to the person in that state that he or she is in a state of that kind. So, for example it is maintained that certain non-conscious states of mind can possess ‘sensory qualities’ of various sorts - one may, in a sense, be in pain without feeling pain, one may have a red sensory quality, even when nothing looks red to one. The idea is that one has a conscious visual experience of red, or a conscious pain sensation, just when one has such a red sensory quality, or pain-quality, and the thought (itself also potentially non-conscious) occurs to one that one has a red sensory quality, or pain-quality.

This way of accounting for consciousness in terms of intentionality may, like theories mentioned, provoke the concern that the distinctively phenomenal sense of consciousness has been slighted - though this time, not in favour of some ‘access’ consciousness, but in favour of reflexive consciousness. One focus of such criticism lies in the idea that such higher-order thought requires the possession of concepts - concepts of types of mental states - that may be lacking in creatures with first order mentality. And it is unclear (in fact it seems false to say) these beings would therefore have no conscious sensory experience in the phenomenal sense. Might that they enduringly exist in a way the world looks to rabbits, dogs, monkeys, and human babies, and might they agreeably feel pain, though they lack the conceptual wherewithal to think about their own experience?

One line of response to such concerns is simply to bite the bullet: dogs, babies and the like might altogether lack higher order thought, but that is no problem for the theory because, indeed, they also altogether lack feelings. Rosenthal, for his part, takes a different line: lack of cognitive sophistication need not instantly disqualify one for consciousness, since the possession of primitive mentalistic concepts requires so little that practically any organism we would consider a serious candidate for sensory consciousness (certainly babies, dogs and bunnies) would obviously pass conscription.

A number of additional worries have been raised about both the necessity and the sufficiency of ‘higher order thought’ for conscious sense experience. In the face of such doubts, one may preserve the idea that consciousness consists in some kind of higher order representation - the mind's ‘scanning’ itself - by abandoning ‘higher order thought’ for another form of representation: one that is not thought-like or conceptual, but somehow sensory in character. Maybe somewhat as we can distinguish between primitive sensory perception of things in our environment, and the more intellectual, conceptual operations based on them, so we can distinguish the thoughts we have about our own (‘inner’) mental goings-on from the (‘inner’) sensing of them. And, if we propose that consciousness consist in this latter sort of higher order representation, it seems we will escape the worries occasioned by the Rosenthalian variant of the ‘reflexivist’ doctrine. In considering such theories, two of the consciousness-themes that earlier discern had in coming together, namely the reflexivity of thought, or higher order representations, and, by contrast, between the conceptual and non-conceptual presentations, as sensory data,

Criticism of ‘inner sense’ theories is likely to focus not so much on the thought that such inner sensing can occur without phenomenal consciousness, or that the latter can occur without the former, as on the difficulty in understanding just what inner sensing (as distinct from higher order thought) is supposed to be, and why we should think we have it. It seems the inner sense theorist’s share with those who distinguish between conceptual and non-conceptual (or sensory) flavours of intentionality the challenge of clarifying and justifying some version of this distinction. But they bear the additional burden of showing how such a distinction can be applied not just to intentionality directed at tables and chairs, but at the ‘furniture of the mind’ as well. One may grant that there are non-conceptual sensory experiences of objects in one's external environment while doubting one has anything analogous regarding the ‘inner’ landscape of mind.

It should be noted that, in spite of the difficulties faced by higher order representation theories, they draw on certain perennially influential sources of philosophical appeal. We do have some willingness to speak of conscious states of mind as states we are conscious or aware of being in. It is tempting to interpret this as indicating some kind of reflexivity. And the history of philosophy reveals many thinkers attracted to the idea that consciousness is inseparable from some kind of self-reflexivity of mind. As noted, varying versions of this idea can be found in Brentano, Husserl, and Sartre, as well as we can go further back in which case of Kant (1787) who spoke explicitly of ‘inner sense,’ and Locke (1690) defined consciousness as the ‘perception of what passes in a man's mind.’ Brentano (controversially) interpreted Aristotle's enigmatic and terse discussion of ‘seeing that one sees’ in De Anima, as an anticipation of his own ‘inner perception’ view. However, there is this critical difference between the thinkers just cited and contemporary purveyors of higher order representation theories. The former does not maintain, as do the latter, that consciousness consists in one's forming the right sort of higher order representation of a possible non-conscious type of mental state. Even if they think that consciousness is inseparable from some sort of mental reflexivity, they do not suggest that consciousness can, so to speak, be analysed into mental parts, none of which they essentially require consciousness. (Some could not maintain this, since they explicitly deny mentality without consciousness.) There is a difference between saying that reflexivity is essential to consciousness and saying that consciousness just consists in or is reducible to a species of mental reflexivity. Advocacy of the former without advocacy of the latter is certainly possible.

Suppose one holds that phenomenal consciousness is distinguishable both from ‘access’ and ‘reflexivity,’ and that it cannot be explained as a special case of intentionality. One might conclude from this that phenomenal consciousness and intentionality are two composite structures exhibiting of themselves of distinct realms as founded in the psychic domain as called the mental, and embrace the idea that the phenomenal are a matter of non-intentional qualia or raw feels. One important current in the analytic tradition has evinced this attitude - it is found, for example, in Wilfrid Sellars' (1956) distinction between ‘sentience’ (sensation) and ‘sapience.’ Whereas the qualities of feelings involved in the former - mere sensations - require no cognitive sophistication and are readily attributable to brutes, the latter - involving awareness of, awareness that - requires that one have the appropriate concepts, which cannot be guaranteed by just having sensations; one needs learning and inferential capacities of a sort Sellars believed possibly only with language. ‘Awareness,’ Sellars says, ‘is a linguistic affair.’

Thus we may arrive at a picture of mind that places sensation on one side, and thought, concepts, and ‘propositional attitudes’ on the other. If one recognizes the distinctively phenomenal consciousness not captured in ‘representationalist’ theories of the kinds just scouted, one may then want to say: that is because the phenomenal belong to mere sentience, and the intentional to sapience. Other influential philosophers of mind have operated with a similar picture. Consider Gilbert Ryle's (1949) contention that the stream of consciousness contains nothing but sensations that provide ‘no possibility of deciding whether the creature that had these was an animal or a human being; An ignoramus, simpletons, or a sane man, only from which nothing is appropriately asked of whether it is correct or incorrect, veridical or nonveridical. And Wittgenstein's (1953) influential criticism of the notion of understanding as an ‘inner process,’ and of the idea of a language for private sensation divorced from public criteria, could be interpreted in ways that sever (phenomenal) consciousness from intentionality. (Such an interpretation would assume that if consciousness could secure understanding, understanding would be an ‘inner process,’ and if phenomenal character bore intentionality with it, private sensations could impart meaning to words.) Also recall Putnam's conviction that the (internal) stream of consciousness cannot furnish the (externally fixed) content of meaning and belief. A similar attitude is evident in Donald Davidson's distinction between sensation and thought (the former is nothing more than a causal condition of knowledge, while the latter can furnish reasons and justifications, but cannot occur without language). Richard Rorty (1979) makes a Sellarsian distinction between the phenomenal and the intentional key to his polemic against epistemological philosophy overall, and ‘foundationalism’ in particular (and takes a generally deflationary view of the phenomenal or ‘qualitative’ side of this divide).

But it is possible to reject attempts to subsume the phenomenal under the intentional as in the ‘representationalist’ accounts of consciousness variously exemplified in Dennett, Dretske, Lycan, Rosenthal, and Tye, without adopting this ‘two separate realms’ conception. We can believe that there is no conception of the intentional from which the phenomenal can be explanatorily derived that does not already include the phenomenal, but still believe also that the phenomenal character of experience cannot be separated from its intentionality, and that having experience of the right sort of phenomenal character is sufficient for having certain forms of intentionality.

Here one might leave open the question whether there is also some kind of phenomenal character (perhaps that involved in some kinds of bodily sensation or after-images) whose possession is not sufficient for intentionality. (Though if we say there is such non-intentional phenomenal character, this would give us a special reason for rejecting the representationalist explanations of phenomenal consciousness) on the other hand, we say phenomenal character always brings intentionality with it, that might be ‘representational’’ of a sort. But its endorsement is consistent with a rejection of attempts to derive phenomenality from intentionality, or reduce the former to a species of the latter, which commonly attract the ‘representationalist’ label. We should distinguish the question of whether the phenomenal can be explained by the intentional from the question of whether the phenomenal are separable from the intentional.

Closer consideration of two of the three themes earlier identified as common to Phenomenological and analytic traditions is needed to come to grips with the latter question. It is necessary to inquire: (1) whether an externalist conception of intentionality can justify separating phenomenal character from intentionality. And one needs to ask: (2) whether one's verdict on the ‘separability’ question stands or falls with acceptance of some version of a distinction between conceptual and non-conceptual (or distinctively sensory) form of intentionality.

The dialectical situation regarding (1) is complex. One way it may seem plausible to answer question (1) in the affirmative, and restrict phenomenal character and intentionality to different sides of some internal/external divide, is to conduct a Cartesian thought experiment, in which one conceives of consciousness with all its subjective riches surviving the utter annihilation of the spatial realm of nature. (Similarly, but less radical, one may conceive of a ‘brain in a vat’ generating an extended history of sense experience indistinguishable in phenomenal character from that of an embodied subject.) If one is committed to an externalist view of intentionality - but rejects the internationalizing strategies for dealing with consciousness - one may conclude that phenomenal character is altogether separable from (and insufficient for) intentionality. However, one may draw rather different conclusions from the Cartesian thought experiment - turning it against externalism. It may seem to one that, since the intentionality of experience would apparently survive along with its phenomenal character, one may then infer that the causal tie between the mind's content and the world of objects beyond it that (according to some versions of externalism) fixes content, is in reality and in at least some cases (or for some contents), no more than contingent. Alternatively, whatever one relies on to argue that this or that relation of experience and world is essential to having any intentionality at all, one might take this to show that phenomenal character is also externally determined in a way that renders the Cartesian scenario of consciousness totally unmoored from the world an illusion. And, if Merleau-Ponty or Heidegger thinks that Husserl's Phenomenological reduction to a sphere of ‘pure’ consciousness cannot be completed, and their reasons make them externalists of some sort, it hardly seems to establish that they are committed to a realm of raw sensory phenomenal consciousness, devoid of intentionality. In fact their rejection of Husserl's notion of ‘uninterpreted’ sensory or ‘hyletic’ data in experience would seem to indicate them, at least, would strongly deny they held such views.

In this arena it is far from clear what we are entitled to regard as secure ground and what as ‘up for grabs.’ However, there do seem to be ways in which all would probably admit that the phenomenal character of experience and externally individuated content come apart, ways in which such content goes beyond anything phenomenal consciousness can supply. For the way it seems to me to experience this computer screen may be no different from the way it seems to my twin to experience some entirely distinct one. Thus where intentional contents are distinguished in such a way as to include the particular objects experienced or thought of, phenomenal character cannot determine the possession of content. Still, that does not show that no content of any sort is fixed by phenomenal character. Perhaps, as some would say, phenomenal character determines ‘narrow’ or ‘notional’ content, but not ‘wide’ (externally ‘fixed’) content. Nor is it even clear that we must judge the sufficiency of phenomenal character for intentionality by adopting some general account of content and its individuation (as ‘narrow’ or ‘wide’ for instance), and then ask whether one's possession of content so considered is entailed by the phenomenal character of one's experience. One may argue that the phenomenal character of one's experience suffices for intentionality as long as having it makes one assessable for truth, accuracy (or other sorts of ‘satisfaction’) without the addition of any interpretation, properly so-called, such as is involved in assessment of the truth or accuracy of sentences or pictures.

Even if one does not globally divide phenomenal character from intentionality along some inner/outer boundary line, to address questions of the sufficiency of phenomenal character for intentionality (and thus of the separability of the latter from the former), one still needs to look at question (2) as above, and the potential relevance of distinctions that have been proposed between conceptual and non-conceptual forms of content or intentionality. Again the situation is complex. Suppose one regards the notion of non-conceptual intentionality or content as unacceptable on the grounds that all content is conceptual. But suppose one also thinks it is clear that phenomenal character is confined to sensory experience and imagery, and that this cannot bring with it the rational and inferential capacities required for genuine concept possession. Then one will have accepted the separability of phenomenal consciousness from intentionality. However, one may, by contrast, take the apparent susceptiblity of phenomenally conscious sense experience to assessment for accuracy, without need for additional, potentially absent interpretation, to show that the phenomenal character of experience is inherently intentional. Then one will say that the burden lies on anyone who claims conceptual powers are crucial to such assessability and can be detached from the possession of such experience: They must identify those powers and show that they are both crucial and detachable in this way. Additionally, one may reasonably challenge the assumption that phenomenal consciousness is indeed confined to the sensory realm; One may say that conceptual thought also has phenomenal character. Even if one does not, one may still base one's confidence in the sufficiency of phenomenal character for intentionality on one's confidence that there is a kind of non-conceptual intentionality that clearly belongs essentially to sense experience.

These considerations, we can see that it is critical to answer the following questions in order to decide whether or not phenomenal character is wholly or significantly separable from intentionality. Does every sort of intentionality that belongs to thought and experiences require an external connection, for which phenomenal characters are insufficient?

Does every sort of intentionality that belongs to sense-experience and sensory imageries require conceptual abilities for which phenomenal character is insufficient? And does every sort of intentionality that belongs to thought require conceptual capacities for which phenomenal character is insufficient?

Suppose one finds phenomenal character quite generally inadequate for the intentionality of thought and sense-experience by answering ‘yes’ either to (i), or to both (ii) and (iii). And suppose one makes the plausible (if non-trivial) assumption that what guarantees’ intentionality for neither sensory experience, nor imagery, nor conceptual thought, guarantees no intentionality that belongs to our minds (including that of emotion, desire and intention - for these later presuppose the former). Then one will find phenomenal character altogether separable from intentionality. Phenomenal character could be as it is, even if intentionality were completely taken away. There is no form of phenomenal consciousness, and no sort of intentionality, such that the first suffices for the second.

A more moderate view might merely answer only one of either (ii) or (iii) in the affirmative (and probably (iii) would be the choice). But still, in that case one recognizes some broad mental domain whose intentionality is in no respect guaranteed by phenomenal character. And that too would mark a considerable limitation on the extent to which phenomenal consciousness brings intentionality with it.

On the other hand, suppose that one answer ‘no’ to (i), and to either (ii) or (iii). Now, external connections and conceptual capacities seem to be what we might most plausibly regard as conditions necessary for the intentionality of thought and experience that could be stripped away while phenomenal character remains constant. So if one thinks that actually neither are generally essential to intentionality and removable while phenomenal character persists unchanged, and one can think of nothing else that is essential for thought and experience to have any intentionality, but for which phenomenal character is insufficient, it seems reasonable to conclude that phenomenal character is indeed sufficient for intentionality of some sort. If one has gone this far, it seems unlikely that one will then think that actual differences in phenomenal character still leave massively underdetermined the different forms of intentionality we enjoy in perceiving and thinking. So, one will probably judge that some kind of phenomenal character suffices for, and is inseparable from, many significant forms of intentionality in at least one of these domains (sensory or cognitive): There are many differences in phenomenal character, and many in intentionality, such that you cannot have the former without the latter. If one also rejects both (ii) and (iii), then one will accept that appropriate forms of phenomenal consciousness are sufficient for a very broad and important range of human intentionality.

Suppose one rejects both the views that consciousness is explanatorily derived from a more fundamental intentionality, as well as the view that phenomenal character is insufficient for intentionality because it is a matter of a purely inward feeling. It seems one might then press farther, and argue for what Flanagan calls ‘consciousness essentialism’ - the view that the phenomenal character of experience is not only sufficient for various forms of intentionality, but necessary also.

This type of thesis needs careful formulation. It does not necessarily commit one to a Cartesian (or Brentanian or Sartrean) claim that all states of mind are conscious - a total denial of the reality of the unconscious. A more qualified thesis does seem desirable. Freud's waning prestige has weakened tendencies to assume that he had somehow demonstrated the reality of unconscious intentionality, the rise of cognitive science has created a new climate of educated opinion that also takes elaborate non-conscious mental machinations for granted. Even if we do not acquiesce in this view, we do (and long have) appealed to explanations of human behaviour that recognize some sort of intentional state other than phenomenally conscious experiences and thoughts.

The way of maintaining the necessity of consciousness to mind that can preserve some space for mind that is not conscious is Searle's agreement, roughly, that we should first distinguish between what he calls ‘intrinsic’ intentionality on the one hand, and merely ‘as if’ intentionality, and ‘interpreter relative’ intentionality, on the other. We may sometimes speak as if artifacts (like thermostats) had beliefs or desires - but this is not to be taken literally. And we may impose ‘conditions of satisfaction’ on our acts and creations (words, pictures, diagrams, etc.) by our interpretation of them - but they have no intentionality independent of our interpretive practices. Intrinsic intentionality, on the other hand - the kind that pertains to our beliefs, perception, and intentions - is neither a mere ‘manner of speech,’ nor our possession of it derived from others' interpretive stance toward us. But then, Searle asks, what accounts for the fact that some state of affairs in the world for which in having an intrinsic intentionality - that they are directed at objects under aspects - and why they are directed under the aspects they are (why they have the content they do)? With conscious states of mind, Searle says, their phenomenal or subjective character determines their ‘aspectual shape.’ Where non-conscious states of mind are concerned, there is nothing to do the job, but their relationship to consciousness. The right relationship, he holds, is this non-conscious state of mind and must be ‘potentially conscious.’ If some psychological theories (of language, of vision) postulated an unconscious so deeply buried that its mental representations cannot even potentially become conscious, so much the worse for those theories.

Searle's views have aroused a number of criticisms. Among the problems areas are these. First, how are we to explain the requirement that intrinsically intentional states be ‘potentially conscious,’ without making it either too easy or too difficult to satisfy? Second, just why is it that the intrinsic intentionality of non-conscious states need’s accounting for, while conscious states are somehow unproblematic. Third, it appears Searle's argument does not offer some general reason to rule out all efforts to give ‘naturalistic’ accounts of conditions sufficient to impose - without the help of consciousness - genuine and not merely interpreter relative intentionality.

Another approach is taken by Kirk Ludwig, who argues that there is nothing to determine whose state of mind a given non-conscious state of mind is, unless that state consists in a disposition to produce a conscious mental state of the right sort. Alleged mental processes that did not tend to produce someone's conscious states of mind appropriately would be no one's, which is to say that they would not be mental states at all. Roughly: consciousness is needed to provide that unity of mind without which there would be no mind. And Ludwig argues that it is therefore a mistake to attribute many of the unconscious inferences with which psychological theorists have long been wont to populate our minds.

The persuasiveness of Searle and Ludwig's arguments depends heavily on demonstrating the failure of alternative accounts of the job that they enlist consciousness to do (such as secure ‘aspectual shape,’ or ownership). One may grant (as does Colin McGinn 1991) that phenomenal character is inseparable from intentionality, but cannot be explained by it, while still maintaining that genuine intentionality (mental content) is quite adequately imposed on animal brains by their acquisition of natural functions of content-bearing - in which consciousness evidently plays no essential role. Or one may (like Jerry Fodor 1987) maintain a robust realist ‘representational theory of mind,’ proposing that the content of mental symbols is stamped on them by their being in the ‘right causal relation’ to the world - while despairing of the prospects for a credible naturalistic theory of consciousness.

The preceding discussion has conveyed some of the complexities and potential ambiguities in talk of ‘consciousness’ and ‘intentionality’ that must be appreciated if one is to resolve questions about the relationship between consciousness and intentionality with any clarity. Brief surveys of relevant aspects of Phenomenological and analytic traditions have brought out some shared areas of interest, namely: The relationship of consciousness to reflexivity and ‘self-directed’ intentionality manages to distinguish events between conceptual and non-conceptual (or sensory) forms of intentionality, and the concerns with which the extent is characterized by either conscious experience or intentional states of mind is essentially ‘world-involving.’ These concerns were seen to bear on attempts to account for consciousness in terms of intentionality, and on questions that arise even if those attempts are rejected - questions regarding the separability of phenomenal consciousness and intentionality. Some attention is given to views that, in some sense, reverse the order of explanation proposed by internationalizing views of consciousness, and take the facts of consciousness to explain the facts of intentionality. Now it is possible to step back and distinguish four general views of the consciousness-intentionality relationship discernable in the philosophical positions canvassed above, as follows.

(1) Consciousness is explanatorily derived from intention

(2) Consciousness is explanatorily derived from intentionality.

(3) Consciousness is underived and separable from intentionality.

(4) Consciousness is underived but also inseparable from intentionality.

(5) Consciousness is underived from, inseparable from, and essential to intentionality.

To adopted view (1) is to accept some internationalizing strategy with respect to consciousness, such as is variously represented by Dennett, Dretske, Lycan, Rosenthal, and Tye. These views differ importantly among themselves. Their differences have much to do with how they treat consciousness-reflexivity issues and the conceptual/non-conceptual (or conceptual/sensory) contrast, and how they view the intersection between the two. But if we accept (1), then our answer to the question of what consciousness has to do with intentionality will ultimately be given in some prior general account of content or intentionality. And there will be no special issue regarding the internal or external fixation of the phenomenal character of experience, over and above what arises for mental content generally.

On the other hand, suppose one reject (1), and holds that experiences are conscious in a phenomenal sense that does not yield to an approach in which one conceives of intentionality (or content, or information bearing) independently of consciousness, and then, by adverting to special operations, or sources, or contents, tells us what consciousness is. At this point, one would face a choice between (2) and (3).

By embracing (2) we yield the ‘raw feel’ in its conception of phenomenality seemingly implicit in Sellars and Ryle. If, on the other hand, we accept (3), we endorse a much more intimate relationship between consciousness and intentionality. Without proposing to account for the former on the basis of the latter, we would hold that phenomenal character is sufficient for intentionality.

But adoption of (3) leaves open a further basic question. Consciousness (of the appropriate sort) may be sufficient, but underived from intentionality. Yet, intentionality does not require consciousness. Thus we come to ask whether having conscious experience of an appropriate sort is necessary to having either sensory or more-than-sensory (conceptual) intentionality. Adopting theses (4), we say ‘yes’ - that such intentionality can come only with consciousness - we will probably have gone as far in making consciousness fundamental to mind as one reasonably can. Again, this is not necessarily to deny the reality of non-conscious mental phenomena. But it could, in a broad way, be interpreted as siding with Husserl, Ludwig and Searle in thinking of consciousness as the irreplaceable source of intentionality and meaning.

This abstract list of four options might leave one without a sense of what is at stake in adopting this or that view. Perhaps the positions themselves will become a little clearer if we make explicit four broad areas of philosophical concern to which the choice among them is relevant.

First, they are relevant to the issue of how to conceive of the mind or the domain of psychology as a whole. Is there some unity to the concept of mind or psychologically phenomenal? Is there something that deserves to be considered the essence of the mental? If consciousness can be thoroughly internationalized (as (1) would have it), maybe (with suitable qualifications), we could uphold the thesis that intentionality is the ‘mark of the mental.’ If we disapprove of (1) and embrace (3), seeing intentionality as inseparable from the phenomenal character of experience, then we still might maintain that both consciousness and intentionality are necessary for real minds - at least, if we adopt (4) as well. But a unified view of the mind seems difficult (if possible) to maintain if one segregates phenomenal character to non-intentional sensation - as in (2). Even if one does not, one may lack a unifying conception of the mental domain, if one is not satisfied with arguments that show that phenomenal consciousness is essential to genuine (not merely ‘as if’ or ‘interpreter derived’) intentionality. In any case, both consciousness and intentionality signify a broad enough psychological categories, in that one's view of their extension and relationship will do much to draw one's map of psychology's terrain.

Second (and relatedly), views about the consciousness-intentionality relationship bear significantly on general questions about the explanation of mental phenomena. One may ask what kinds of things we might try to explain in the mental domain, what sorts of explanations we should seek, and what prospects of success we have in finding them. If we accept (1) and some internationalizing account of consciousness, we will not suppose as do some (Chalmers 1996, Levine 2001, McGinn 1991, and Nagel 1974) those phenomenal consciousness poses some specially recalcitrant (maybe hopelessly unsolvable) problem for reductive physicalist or materialist explanations. Rather, we will see the basic challenge as that of giving a natural scientific account of intentionality or mental representation. And this indeed is a reason some are attracted to (1). One may believe that it offers us the only hope for a natural scientific understanding of consciousness. The underlying thought is that a science of consciousness must adopt this strategy: First conceive of intentionality (or content or mental representation) in a way that separates it from consciousness, and see intentionality as the outcome of familiar (and non-intentional) natural causal processes. Then, by further specifying the kind of intentionality involved (in terms of its use, its sources, its content), we can account for consciousness. In other words: ‘naturalize’ intentionality, then internationalized consciousness, and mind has found its place in nature.

However, we should recognize a distinction between those whose envisioned naturalistic explanation would require underlying forms of necessity and impossibility stronger that pertaining to laws of nature generally - such as either conceptual or ‘metaphysical’ necessity - and those who see the link between explanans and explanandum as simply one of natural scientific law. David Chalmers' (1996) proposals for ‘naturalistic dualism’ (unlike those of the aforementioned naturalizers) put him in the second group. He argues that phenomenal consciousness in its various forms supervenes (not conceptually or metaphysically but only as a matter of nature's laws) on functional organization, and that this permits us to envisage (‘non-reductive’) ways of explaining consciousness by appeal to such organization.

Those who reject attempts to explain the phenomenal consciousness via a theory of intentionality still may reasonably proclaim allegiance to ‘naturalism.’ One may take phenomenal consciousness to be, in a sense, psychologically basic (if all that is mental is phenomenally either conscious or intentional, and no internationalizing account of phenomenal character is feasible). But one might still hold that some non-intentional neuropsychological, or other recognizable physicalist, there to some explanation of the phenomenal character of experience is to be had, because the explanatory link is otherwise to exhibit an appropriately strong conceptual or metaphysical necessity. Only for measures for which are regarded to that of nothing stronger than psychophysical laws of nature are needed to give us the prospect of a natural scientific account of consciousness.

However, if we not only reject internationalizing accounts of phenomenal character, but also see it as inseparable from intentionality (if we reject both (1) and (2) and agree to whatever problems are attached to physicalist explanations of consciousness will also infect prospects for explaining intentionality - to some extent at least. And this will hold, even if we remain aloof from (d), and do not claim that phenomenal consciousness is essential to intentionality. For if we think that much of the intentionality we have in perceiving, imagining, and thinking is integral to the phenomenal character of such experience, then without a reductive explanation of that phenomenal character, our possession of the intentionality it brings with it will not be reductively explained either.

Finally, it should be noted that if one holds (4), this may have important consequences for what forms of psychological explanation that once acceptably found the remaining agreement stems for that which one's mental processes must have the right relationship to one's conscious experiences to count as one's mental processes at all. If they are right, postulated processes that do not bear this relation to our experiential lives cannot be going on in our minds.

Regardlessly, another enlarging area of concern is of choice, in that between (1)-(4) tells of a direction among proven rights is to continue in having epistemological connections, in that if one is to embrace of (2), and something like a Sellarsian or Davidsonian distinction between sensation and thought, putting phenomenal character exclusively on the ‘sensation’ side, and intentionality exclusively on the ‘thought’ side of this divide, the place of consciousness in a philosophical account of knowledge will likely be meager - at most phenomenal character will be a causal condition, without a role to play in the warrant or justification of claims to knowledge. However, if one takes routes (1) or (3) the situation will appear rather different. If one is to consider the consequent for, either of which is to internationalize consciousness, or else views intentionality as inseparable from phenomenal character, there will then be more room to view consciousness as central to accounts of the warrant involved in first-person (‘introspective’) knowledge of mind, and empirical or perceptual knowledge. Though just how one goes about this, and with what success, will depend on how (if one chooses (1)) one internationalizes consciousness, and (if one chooses (1) or (3), that will depend on what sort of intentionality or content one thinks phenomenal consciousness brings with it. The place of consciousness in one's understanding of introspective or empirical knowledge will be rather different, depending on how one resolves the issues regarding: Reflexivity, the conceptual/non-conceptual distinction, and externalisms.

A fourth area of philosophical concern we may indicate broadly, closely bound to our conception of the relation of consciousness and intentionality, has to do with value. How intimately is consciously bound up with those features of our own and others' lives that give them intrinsic or non-instrumental value for us? We may think that the pleasure and suffering that demand our ethical concern are necessarily phenomenally conscious - and that this evaluative significance remains even if phenomenal character is non-intentional. However, the more intentionality is seen as inherent to the phenomenal character of experience, the more the latter will be bound to manifestations of intelligence, emotion, and understanding that appear to give human (and perhaps at least another animal life) its special importance for us. It may seem that those opting for (3) share at least this much ground with their internationalizing opponents who go for (1): They both (unlike those who adopt (2)) are in a position to claim consciousness is crucial to whatever special moral regard we think appropriate only toward those whose psychologies involve a kind of intentionality for which possession of painful or pleasant experience is not sufficient. However, this needs qualification on two counts. First, if one's embrace of (1) includes an internationalizing strategy that limits phenomenal character to the sensory realm, one will limit the moral significance of phenomenal consciousness accordingly. Second, to those who hold, it may seem their opponents' internationalizing theories remove from view those very qualities of experience that make life worth living, and so they will hardly seem like allies on the issue of value. Further, if the proponent of hesitant anticipations were in going insofar as to be taken on (4) - conscious essentialism - those who make that additional commitment might wonder how those who do not could ultimately accord the possession of consciousness much greater non-instrumental value than the possession of a sophisticated but totally non-conscious mind.

From this survey it seems fair to conclude that working out a detailed view of the relation between consciousness and intentionality is hardly a peripheral matter philosophically. Potentially it has extensive consequences for one's views concerning these four important, broad topics: (I) The unity of mental phenomena (Do consciousness or intentionality (or both together) somehow unifies the domain of the psychological?) (II) The explanation of mental phenomena (Can consciousness and intentionality are explained separately? (III) Is explaining the one key to explaining the other? Introspective and empirical knowledge (What relation to intentionality would give consciousness a central epistemological role in either?) (IV) The value of human and other animal life. (What relation of consciousness and intentionality (if any) underlies the non-instrumental value we accord ourselves and others?)

We collectively glorify our ability to think as the distinguishing characteristic of humanity; We personally and mistakenly glorify our thoughts as the distinguishing pattern of whom we are. From the inner voice of thought-as-words to the wordless images within our minds, thoughts create and limit our personal world. Through thinking we abstract and define reality, reason about it, react to it, recall past events and plan for the future. Yet thinking remains both woefully underdeveloped in most of us, as well as grossly overvalued. We can best gain some perspective on thinking in terms of energies.

Automatic thinking draws us away from the present, and wistfully allows our thoughts to meander where they would, carrying our passive attention along with them. Like water running down a mountain stream, thoughts running on auto-pilot careens through the spaces of perception, randomly triggering associative links within our vast storehouse of memory. By way of itself, such associative thought is harmless. However, our tendency to believe in, act upon, and drift away with such undirected thought keeps us operating in an automatic mode. Lulled into an inner passivity by our daydreams and thought streams, we lose contact with the world of actual perceptions, of real life. In the automatic mode of thinking, I am completely identified with my thoughts, believing my thoughts are I, and believing that I am the conceptualization forwarded by me to think of thoughts that are sometimes thought as unthinkable.

Another mode of automatic thinking consists of repetitious and habitual patterns of thought. These thought tapes and our running commentary on life, unexamined by the light of awareness, keep us enthralled, defining who we are and perpetuating all our limiting assumptions about what is possible for us. Driving and driven by our emotions, these ruts of thought create our false persona, the mask that keeps us disconnected from others and from our own authentic self. More than any other single factor, automatic thinking hinders our contact with presence, limits our being, and Forms our path. The autopilot of thought constantly calls us away from the most recent or the current of immediacy. Thus, keeping us fixed on the most superficial levels of our being.

Sometimes we even notice strange, unwanted thoughts that we consider horrible or shameful. We might be upset or shaken that we would think such thoughts, but those reactions only serves to sustain the problematic thoughts by feeding them energy. Furthermore, that self-disgust is based on the false assumption that we are our thoughts, that even unintentional thoughts, arising from our conditioned minds, are we. They are not we and we need not act upon or react to them. They are just thoughts with no inherent power and no real message about whom we are. We can just relax and let them go - or not. Troubling thoughts that recur over a long period and hinder our inner work may require us to examine and heal their roots in our conditioning, perhaps with the help of a psychotherapist.

Sensitive thinking puts us in touch with the meaning of our thoughts and enables us to think logically, solve problems, make plans, and carry on a substantive conversation. A good education develops our ability to think clearly and intentionally with the sensitive energy. With that energy level in our thinking brain, no longer totally submerged in the thought stream, we can move about in it, choosing among and directing our thoughts based on their meaning.

Conscious thinking means stepping out of the thought stream altogether, and surveying it from the shore. The thoughts themselves may even evaporate, leaving behind a temporary empty streambed. Consciousness reveals the banality and emptiness of ordinary thinking. Consciousness also permits us to think more powerfully, holding several ideas, their meanings and ramifications in our minds at once.

When the creative energy reaches thought, truly new ideas spring up. Creative thinking can happen after a struggle, after exhausting all known avenues of relevant ideas and giving up, shaping and emptying the stage so the creative spark may enter. The quiet, relaxed mind also leaves room for the creative thought, a clear channel for creativity. Creative and insightful thoughts come to all of us in regard to the situations we face in life. The trick is to be aware enough to catch them, to notice their significance, and if they withstand the light of sober and unbiased evaluation, to act on them.

In the spiritual path, we work to recognize the limitations of thought, to recognize its power over us, and especially to move beyond it. Along with Descartes, we subsist in the realm of ‘thoughts‘, but thoughts are just thoughts. They are not we. They are not who we are. No thought can enter the spiritual realms. Rather, the material world defines the boundaries of thought, despite its power to conceive lofty abstractions. We cannot think our way into the spiritual reality. On the contrary, identification with thinking prevents us from entering the depths. As long as we believe that refined thinking represents our highest capacity, we shackle ourselves exclusively to this world. All our thoughts, all our books, all our ideas wither before the immensity of the higher realms.

A richly developed body of spiritual practices engages of thought, from repetitive prayer and mantras, to contemplation of an idea, to visualizations of deities. In a most instructive and invaluable exercise, we learn to see beyond thought by embracing the gaps, the spaces between thoughts. After sitting quietly and relaxing for some time, we turn our attention toward the thought stream within us. We notice thoughts come and go of their own accord, without prodding or pushing from us. If we can abide in this relaxed watching of thought, without falling into the stream and flowing away with it, the thought stream begins to slow, the thoughts fragment. Less enthralled by our thoughts, we begin to see that we are not our thoughts. Less controlled by, and at the mercy of, our thoughts, we begin to be aware of the gaps between thought particles. These gaps open to consciousness, underlying all thought. Settling into these gaps, we enter and become the silent consciousness beneath thought. Instead of being in our thoughts, our thoughts are in us.

There is potentially a rich and productive interface between neuroscience/cognitive science. The two traditions, however, have evolved largely independent, based on differing sets of observations and objectives, and tend to use different conceptual frameworks and vocabulary representations. The distributive contributions to each their dynamic functions of finding a useful common reference to further exploration of the relations between neuroscience/cognitive science and psychoanalysis/psychotherapy.

Recent historical gaps between neuroscience/cognitive science and psychotherapy are being productively closed by, among other things, the suggestion that recent understandings of the nervous system as a modeler and predictor bear a close and useful similarity to the concepts of projection and transference. The gap could perhaps be valuably narrowed still further by a comparison in the two traditions of the concepts of the ‘unconscious’ and the ‘conscious’ and the relations between the two. It is suggested that these be understood as two independent ‘story generators’ - each with different styles of function and both operating optimally as reciprocal contributors to each others' ongoing story evolution. A parallel and comparably optimal relation might be imagined for neuroscience/cognitive science and psychotherapy.

For the sake of argument, imagine that human behaviour and all that it entails (including the experience of being a human and interacting with a world that includes other humans) is a function of the nervous system. If this were so, then there would be lots of different people who are making observations of (perhaps different) aspects of the same thing, and telling (perhaps different) stories to make sense of their observations. The list would include neuroscientists and cognitive scientists and psychologists. It would include as well psychoanalysts, psychotherapists, psychiatrists, and social workers. If we were not too fussy about credentials, it should probably include as well educators, and parents and . . . babies? Arguably, all humans, from the time they are born, spend a considerable reckoning of time making observations of how people (others and themselves) behave and why, and telling stories to make sense of those observations.

The stories, of course, all differ from one another to greater or lesser degrees. In fact, the notion that ‘human behaviour and all that it entails . . . are a function of the nervous system’ is itself a story used to make sense of observations by some people and not by any other? It is not my intent here to try to defend this particular story, or any other story for that matter. Very much to the contrary, what I want to do is to explore the implications and significance of the fact that there are different stories and that they might be about the same (some)thing.

In so doing, I want to try to create a new story that helps to facilitate an enhanced dialogue between neuroscience/cognitive science, on the one hand, and psychotherapy, on the other. That new stories of itself are stories of conflicting historical narratives . . . what is within being called the ‘nervous system’ but others are free to call the ‘self,’ ‘mind,’ ‘soul,’ or whatever best fits their own stories. What is important is the idea that multiple things, evident by their conflicts, may not in fact be disconnected and adversarial entities but could rather be fundamentally, understandably, and valuably interconnected parts of the same thing.

‘Non-conscious Prediction and a Role for Consciousness in Correcting Prediction Errors’ by Regina Pally (Pally, 2004) is the take-off point for my enterprise. Pally is a practising psychiatrist, psychoanalyst, and psychotherapist who have actively engaged with neuroscientists to help make sense of her own observations. I am a neuroscientists who recently spent two years as an Academic Fellow of the Psychoanalytic Centre of Philadelphia, an engagement intended to expand my own set of observations and forms of story-telling. The significance of this complementarity, and of our similarities and differences, is that something will emerge in this commentary.

Many psychoanalysts (and psychotherapists too, I suspect) feel that the observations/stories of neuroscience/cognitive science are for their own activities at best, find to some irrelevance, and at worst destructive or are they not the same probability that holds for many neuroscientists/cognitive scientists. Pally clearly feels otherwise, and it is worth exploring a bit why this is so in her case. A general key, I think, is in her line ‘In current paradigms, the brain has intrinsic activity, is highly integrated, is interactive with the environment, and is goal-oriented, with predictions operating at every level, from lower systems to . . . the highest functions of abstract thought.’ Contemporary neuroscience/cognitive science has indeed uncovered an enormous complexity and richness in the nervous system, ‘making it not so different from how psychoanalysts (or most other people) would characterize the self, at least not in terms of complexity, potential, and vagary.’ Given this complexity and richness, there is substantially less reason than there once was to believe psychotherapists and neuroscientists/cognitive scientists are dealing with two fundamentally different things.

Pally suspect, more aware of this than many psychotherapists because she has been working closely with contemporary neuroscientists who are excited about the complexity to be found in the nervous system. And that has an important lesson, but there is an additional one at least as important in the immediate context. In 1950, two neuroscientists wrote that, ‘the sooner we recognize the certainty of the complexity that is highly functional, just as those who recognize the Gestalts under which they leave the reflex physiologist confounded, in fact they support the simplest functions in the sooner that we will see that the previous terminological peculiarities that seem insurmountably carried between the lower levels of neurophysiology and higher behavioural theory simply dissolve away.’

And in 1951 another said: ‘ I am coming more to the conviction that the rudiments of every behavioural mechanism will be found far down in the evolutionary scale and represented in primitive activities of the nervous system.’

Neuroscience (and what came to be cognitive science) was engaged from very early on in an enterprise committed to the same kind of understanding sought by psychotherapists, but passed through a phase (roughly from the 1950 through to the 1980's) when its own observations and stories were less rich in those terms. It was a period that gave rise to the notion that the nervous system was ‘simple’ and ‘mechanistic,’ which in turn made neuroscience/cognitive science seem less relevant to those with broader concerns, perhaps even threatening and apparently adversarial if one equated the nervous system with ‘mind,’ or ‘self,’ or ‘soul,’ since mechanics seemed degrading to those ideas. Arguably, though, the period was an essential part of the evolution of the contemporary neuroscience/cognitive science story, one that laid needed groundwork for rediscovery and productive exploration of the richness of the nervous system. Psychoanalysis/psychotherapy, and, of course, move through their own story of evolution over its presented time. That the two stories seemed remote from one another during this period was never adequate evidence that they were not about the same thing but only an expression of their needed independent evolutions.

An additional reason that Pally is comfortable with the likelihood that psychotherapists and neuroscientists/cognitive scientists are talking about the same thing is her recognition of isomorphism (or congruities, Pulver 2003) between the two sets of stories, places where different vocabularies in fact seem to be representing the same (or quite similar) things. I am not sure I am comfortable calling these ‘shared assumptions’ (as Pally does) since they are actually more interesting and probably more significant if they are instead instances of coming to the same ideas from different directions (as I think they are). In this case, the isomorphism tend to imply that rephrasing Gertrude Stein, that ‘there proves to be the actualization in the exception of there.’ Regardless, Pally has entirely appropriately and, I think, usefully called attention to an important similarity between the psychotherapeutic concept of ‘transference’ and an emerging recognition within neuroscience/cognitive science that the nervous system does not so much collect information about the world as generate a model of it, act in relation to that model, and then check incoming information against the predictions of that model. Pally's suggestion that this model reflects in part early interpersonal experiences, can be largely ‘unconscious,’ and so may cause inappropriate and troubling behaviour in current time seems to be entirely reasonable. So too, are those that constitute her thought, in that of the interactions with which an analyst can help by bringing the model to ‘consciousness’ through the intermediary of recognizing the transference onto the analyst.

The increasing recognition of substantial complexity in the nervous system together with the presence of identifiable isomorphism that provide a solid foundation for suspecting that psychotherapists and neuroscientists/cognitive scientists are indeed talking about the same thing. But the significance of different stories for better understanding a single thing lies as much in the differences between the stories as it does in their similarities/isomorphism, in the potential for differing and not obviously isomorphic stories to modify another productively, and yielding a new story in the process. With this thought in mind, I want to call attention to some places where the psychotherapeutic and the neuroscientific/cognitive scientific stories have edges that rub against one another than smoothly fitting together. And perhaps to ways each could be usefully further evolved in response to those non-isomorphism.

Unconscious stories and ‘reality.’ Though her primary concern is with interpersonal relations, Pally clearly recognizes that transference and related psychotherapeutic phenomena are one (actually relatively small) facet of a much more general phenomenon, the creation, largely unconsciously, of stories that are understood to be but are not that any necessary thoughtful pronunciations inclined for the ‘real world.’ Ambiguous figures illustrate the same general phenomenon in a much simpler case, that of visual perception. Such figures may be seen in either of two ways; They represent two ‘stories’ with the choice between them being, at any given time, largely unconscious. More generally, a serious consideration of a wide array of neurobiological/cognitive phenomena clearly implies that, as Pally says, we do not see ‘reality,’ but only have stories to describe it that result from processes of which we are not consciously aware.

All of this raises some quite serious philosophical questions about the meaning and usefulness of the concept of ‘reality.’ In the present context, what is important is that it is a set of questions that sometimes seem to provide an insurmountable barrier between the stories of neuroscientists/cognitive scientists, who largely think they are dealing with reality, and psychotherapists, who feel more comfortable in more idiosyncratic and fluid spaces. In fact, neuroscience and cognitive science can proceed perfectly well in the absence of a well-defined concept of ‘reality’ and, without being fully conscious of it, committing to fact as they do so. And psychotherapists actually make more use of the idea of ‘reality’ than is entirely appropriate. There is, for example, a tendency within the psychotherapeutic community to presume that unconscious stories reflect ‘traumas’ and other historically verifiable events, while the neurobiological/cognitive science story says quite clearly that they may equally reflect predispositions whose origins reflect genetic information and hence bear little or no relation to ‘reality’ in the sense usually meant. They may, in addition, reflect random ‘play,’ putting them even further out of reach of easy historical interpretation. In short, with regard to the relation between ‘story’ and ‘reality,’ each set of stories could usefully be modified by greater attention to the other. Differing concepts of ‘reality’ (perhaps the very concept itself) gets in the way of usefully sharing stories. The mental/cognitive scientists' preoccupation with ‘reality’ as an essential touchstone could valuably be lessened, and the therapist's sense of the validation of stories in terms of personal and historical idiosyncrasies could be helpfully adjusted to include a sense of actual material underpinnings.

The Unconscious and the Conscious. Pally appropriately makes a distinction between the unconscious and the conscious, one that has always been fundamental to psychotherapy. Neuroscience/cognitive science has been slower to make a comparable distinction but is now rapidly beginning to catch up. Clearly some neural processes generate behaviour in the absence of awareness and intent and others yield awareness and intent with or without accompanying behaviour. An interesting question however, raised at a recent open discussion of the relations between neuroscience and psychoanalysis, is whether the ‘neurobiological unconscious’ is the same thing as the ‘psychotherapeutic unconscious,’ and whether the perceived relations between the ‘unconscious’ and the’conscious’ are the same in the two sets of stories. Is this a case of an isomorphism or, perhaps more usefully, a masked difference?

An oddity of Pally's article is that she herself acknowledges that the unconscious has mechanisms for monitoring prediction errors and yet implies, both in the title of the paper, and in much of its argument, that there is something special or distinctive about consciousness (or conscious processing) in its ability to correct prediction errors. And here, I think, there is evidence of a potentially useful ‘rubbing of edges’ between the neuroscientific/cognitive scientific tradition and the psychotherapeutic one. The issue is whether one regards consciousness (or conscious processing) as somehow ‘superior’ to the unconscious (or unconscious processing). There is a sense in Pally of an old psychotherapeutic perspective of the conscious as a mechanism for overcoming the deficiencies of the unconscious, of the conscious as the wise father/mother and the unconscious as the willful child. Actually, Pally does not quite go this far, as I will point out in the following, but there is enough of a trend to illustrate the point and, without more elaboration, I do not think of many neuroscientists/cognitive scientists will catch Pally's more insightful lesson. I think Pally is almost certainly correct that the interplay of the conscious and the unconscious can achieve results unachievable by the unconscious alone, but think also that neither psychotherapy nor neuroscience/cognitive science are yet in a position to say exactly why this is so. So let me take a crack here at a new, perhaps bi-dimensional story that could help with that common problem and perhaps both traditions as well.

A major and surprising lesson of comparative neuroscience, supported more recently by neuropsychology (Weiskrantz, 1986) and, more recently still, by artificial intelligence, is that an extraordinarily rich repertoire of adaptive behaviour can occur unconsciously, in the absence of awareness of intent (be supported by unconscious neural processes). It is not only modelling the world and prediction. Error correction that can occur this way but virtually (and perhaps literally) the entire spectrum of behaviour externally observed, including fleeing from a threat, and of approaching good things, generating novel outputs, learning from doing so, and so on.

This extraordinary terrain, discovered by neuroanatomists, electro-physiologists, neurologists, behavioural biologists, and recently extended by others using more modern techniques, is the unconscious of which the neuroscientists/cognitive scientist speaks. It is the area that is so surprisingly rich that it creates, for some people, the puzzle about whether there is anything else at all. Moreover, it seems, at first glance, to be a totally different terrain from that of the psychotherapist, whose clinical experience reveals a territory occupied by drives, unfulfilled needs, and the detritus with which the conscious would prefer not to deal.

As indicated earlier, it is one of the great strengths of Pally's article to suggest that the two areas may in fact, turns out to be the same as in many ways that if they are of the same, then its question only compliments in what way are the ‘unconscious’ and the ‘conscious’ of showing to any difference? Where now are the ‘two stories?’ Pally touches briefly on this point, suggesting that the two systems differ not so much (or at all?) In what they do, but rather in how they do it. This notion of two systems with different styles seems to me worth emphasizing and expanding. Unconscious processing is faster and handles many more variables simultaneously. Conscious processing is slower and handles numerously fewer variables at one time. It is likely that their equalling a host of other differences in style as well, in the handling of number for example, and of time.

In the present context, however, perhaps the most important difference in style is one that Lacan called attention to from a clinical/philosophical perspective - the conscious (conscious processing) have in themselves forwarded by some objective ‘coherence,’ that it attempts to create a story that makes sense simultaneously of all its parts. The unconscious, on the other hand, is much more comfortable with bits and pieces lying around with no global order. To a neurobiologist/cognitive scientist, this makes perfectly good sense. The circuitry embodies that of the unconscious (sub-cortical circuitry?) Is an assembly of different parts organized for a large number of different specific purposes, and only secondarily linked together to try to assure some coordination? The circuitry, has, once, again, to involve in conscious processing (neo-cortical circuitry?) On the other hand, seems to both be more uniform and integrated and to have an objective for which coherence is central.

That central coherence is well-illustrated by the phenomena of ‘positive illusions,’ exemplified by patients who receive a hypnotic suggestion that there is an object in a room and subsequently walk in ways that avoid the object while providing a variety of unrelated explanations for their behaviour. Similar ‘rationalization’ is, of course, seen in schizophrenic patients and in a variety of fewer dramatic forms in psychotherapeutic settings. The ‘coherent’ objective is to make a globally organized story out of the disorganized jumble, a story of (and constituting) the ‘self.’

What all this introduces that which is the mind or brain for which it is actually organized to be constantly generating at least two different stories in two different styles. One, written by conscious processes in simpler terms, is a story of/about the ‘self’ and experienced as such, for developing insights into how such a story can be constructed using neural circuitry. The other is an unconscious ‘story’ about interactions with the world, perhaps better thought of as a series of different ‘models’ about how various actions relate to various consequences. In many ways, the latter are the grist for the former.

In this sense, we are safely back to the two stories that are ideologically central in their manifestations as pronounced in psychotherapy, but perhaps with some added sophistication deriving from neuroscience/cognitive science. In particular, there is no reason to believe that one story is ‘better’ than the other in any definitive sense. They are different stories based on different styles of story telling, with one having advantages in certain sorts of situations (quick responses, large numbers of variables, more direct relation to immediate experiences of pain and pleasure) and the other in other sorts of situations (time for more deliberate responses, challenges amenable to handling using smaller numbers of variables, more coherent, more able to defer immediate gratification/judgment.

In the clinical/psychotherapeutic context, an important implication of the more neutral view of two story-tellers outlined above is that one ought not to over-value the conscious, nor to expect miracles of the process of making conscious what is unconscious. In the immediate context, the issue is if the unconscious is capable of ‘correcting prediction errors,’ then why appeal to the conscious to achieve this function? More generally, what is the function of that persistent aspect of psychotherapy that aspires to make the unconscious conscious? And why is it therapeutically effective when it is? Here, it is worth calling special attention to an aspect of Pally's argument that might otherwise get a bit lost in the details of her article: . . . the therapist encourages the wife consciously to stop and consider her assumption that her husband does not properly care about her, and effortfully to consider an alternative view and inhibit her impulse to reject him back. This, in turn, creates a new type of experience, one in which he is indeed more loving, such that she can develop new predictions.’

It is not, as Pally describes it, the simple act of making something conscious that is therapeutically effective. What is necessary is to decompose the story consciously (something that is made possible by its being a story with a small number of variables) and, even what is more important, to see if the story generates a new ‘type of experience’ that in turn causes the development of ‘new predictions.’ The latter, is an effect of the conscious on the unconscious, an alteration of the unconscious brought about by hearing, entertaining, and hence acting on a new story developed by the conscious. It is not ‘making things conscious’ that is therapeutically effective; it is the exchange of stories that encourages the creation of a new story in the unconscious.

For quite different reasons, Grey (1995) earlier made a suggestion not dissimilar to Pally's, proposing that consciousness was activated when an internal model detected a prediction failure, but acknowledged he could see no reason ‘why the brain should generate conscious experience of any kind at all.’ Seemingly, in spite of her title, there seems of nothing really to any detection of prediction errors, especially of what is important that Pally's story is the detection of mismatches between two stories. One unconscious and the other conscious, and the resulting opportunity for both to shape a less trouble-making new story. That, briefly may be why the brain ‘should generate conscious experience,’ to reap the benefits of having a second story teller with a different style. Paraphrasing Descartes, one might say ‘I am, and I can think, therefore I can change who I am.’ It is not only the neurobiological ‘conscious’ that can undergo change; it is the neurobiological ‘unconscious’ as well.

More generally, I want to suggest that the most effective psychotherapy requires the recognitions, rapidly emanating from the neuro- sciences and their cognitive counterpart for which are exposed of each within the paradigms of science, that the brain/mind has evolved with two (or more) independent story tellers and has done so precisely because there are advantages to having independent story tellers that generate and exchange different stories. The advantage is that each can learn from the other, and the mechanisms to convey the stories and forth and for each story teller to learn from the stories of the others occurring as a part of our evolutionary endowment as well. The problems that bring patients into a therapist's office are problems in the breakdown of story exchange, for any of a variety of reasons, and the challenge for the therapist is to reinstate the confidence of each story teller in the value of the stories created by the other. Neither the conscious nor the unconscious is primary; they function best as an interdependent loop with each developing its own story facilitated by the semi-independent story of the other. In such an organization, there are not only no ‘real,’ and no primacy for consciousness, there is only the ongoing development and, ideally, effective sharing of different stories.

There are, in the story I am outlining, implications for neuroscience/cognitive science as well. The obvious key questions are what does one mean (in terms of neurons and neuronal assemblies) by ‘stories,’ and in what ways are their construction and representation different in unconscious and conscious neural processing. But even more important, if the story I have outlined makes sense, what are the neural mechanisms by which unconscious and conscious stories are exchanged and by which each kind of story impacts on the other? And why (again in neural terms) does the exchange sometimes break down and fail in a way that requires a psychotherapist - an additional story teller - to be repaired?

Just as the unconscious and the conscious are engaged in a process of evolving stories for separate reasons and using separate styles, so too have been and will continue to be neuroscience/cognitive science and psychotherapy. And it is valuable that both communities continue to do so. But there is every reason to believe that the different stories are indeed about the same thing, not only because of isomorphism between the differing stories but equally because the stories of each can, if listened to, are demonstrably of value to the stories of the other. When breakdowns in story sharing occur, they require people in each community who are daring enough to listen and be affected by the stories of the other community. Pally has done us all a service as such a person. I hope to further the constructs that bridge her to lay, and that others will feel inclined to join in an act of collectivity such that has enormous intellectual potential and relates directly too more seriously psychological need in the mental health arena. Indeed, there are reasons to believe that an enhanced skill at hearing, respecting, and learning from differing stories about similar things would be useful in a wide array of contexts.

The physical basis of consciousness appears to be the major and most

singular challenge to the scientific, reductionist world view. In the closing years of the second millennium, advances in the ability to record the activity of individual neurons in the brains of monkeys or other animals while they carry out particular tasks, combined with the explosive development of functional brain imaging in normal humans, has lead to a renewed empirical program to discover the scientific explanation of consciousness. This article reviews some of the relevant experimental work and argues that the most advantageous strategy for now is to focus on discovering the neuronal correlates of consciousness.

Consciousness is a puzzling state-dependent property of certain types of complex, adaptive systems. The best example of one type of such systems is a healthy and attentive human brain. If the brain is anaesthetized, consciousness ceases. Small lesions in the midbrain and thalamus of patients can lead to a complete loss of consciousness, while destruction of circumscribed parts of the cerebral cortex of patients can eliminate very specific aspects of consciousness, such as the ability to be aware of motion or to recognize objects as faces, without a concomitant loss of vision usually. Given the similarity in brain structure and behaviour, biologists commonly assume that at least some animals, in particular non-human primates, share certain aspects of consciousness with humans. Brain scientists, in conjunction with cognitive neuroscientists, are exploiting a number of empirical approaches that shed light on the neural basis of consciousness. Since it is not known to what extent, artificial systems, such as computers and robots, can become conscious, this article will exclude these from consideration.

Largely, neuroscientists have made a number of working assumptions that, in the fullness of time, need to be justified more fully.

(1) There is something to be explained; that is, the subjective content associated with a conscious sensation - what philosophers point to the qualia - does exist and has its physical basis in the brain. To what extent qualia and all other subjective aspects of consciousness can or cannot be explained within some reductionist framework remains highly controversially.

(2) Consciousness is a vague term with many usages and will, in the fullness of time, be replaced by a vocabulary that more accurately reflect the contribution of different brain processes (for a similar evolution, consider the usage of memory, that has been replaced by an entire hierarchy of more specific concepts). Common to all forms of consciousness is that it feels like something (e.g., to ‘see blue,’ to ‘experience a headache,’ or to ‘reflect upon a memory’). Self-consciousness is but one form of consciousness.

It is possible that all the different aspects of consciousness (smelling, pain, visual awareness, effect, self-consciousness, and so on) employ a basic common mechanism or perhaps a few such mechanisms. If one could understand the mechanism for one aspect, then one will have gone most of the way toward understanding them all.

(3) Consciousness is a property of the human brain, a highly evolved system. It therefore must have a useful function to perform. Crick and Koch (1998) assumes that the function of the neuronal correlate of consciousness is to produce the best current interpretation of the environment-in the light of past experiences-and to make it available, for a sufficient time, to the parts of the brain that contemplate, plan and execute voluntary motor outputs (including language). This needs to be contrasted with the on-line systems that bypass consciousness but that can generate stereotyped behaviours.

Note that in normally developed individuals motor output is not necessary for consciousness to occur. This is demonstrated by lock-in syndrome in which patients have lost (nearly) all ability to move yet are clearly conscious.

(4) At least some animal species posses some aspects of consciousness. In particular, this is assumed to be true for non-human primates, such as the macaque monkey. Consciousness associated with sensory events in humans is likely to be related to sensory consciousness in monkeys for several reasons. Firstly, trained monkeys show similar behaviour to that of humans for many low-level perceptual tasks (e.g., detection and discrimination of visual motion or depth. Secondly, the gross neuroanatomy of humans and non-human primates are rather similar once the difference in size has been accounted for. Finally, functional magnetic resonance imaging of human cerebral cortex is confirming the existence of a functional organization in sensory cortical areas similar to that discovered by the use of single cell electrophysiology in the monkey. As a corollary, it follows that language is not necessary for consciousness to occur (although it greatly enriches human consciousness).

It is important to distinguish the general, enabling factors in the brain that are needed for any form of consciousness to occur from modulating ones that can up-or-down regulate the level of arousal, attention and awareness and from the specific factors responsible for a particular content of consciousness.

An easy example of an enabling factor would be a proper blood supply. Inactivate the heart and consciousness ceases within a fraction of a minute. This does not imply that the neural correlate of consciousness is in the heart (as Aristotle thought). A neuronal enabling factor for consciousness is the intralaminar nuclei of the thalamus. Acute bilateral loss of function in these small structures that are widely and reciprocally connected to the basal ganglia and cerebral cortex leads to an immediate coma or profound disruption in arousal and consciousness.

Among the neuronal modulating factors are the various activities in nuclei in the brain stem and the midbrain, often collectively referred to as the reticular activating system, that control in a widespread and quite specific manner the level of noradrenaline, serotonin and acetylcholine in the thalamus and forebrain. Appropriate levels of these neurotransmitters are needed for sleep, arousal, attention, memory and other functions critical to behaviour and consciousness.

Yet any particular content of consciousness is unlikely to arise from these structures, since they probably lack the specificity necessary to mediate a sharp pain in the right molar, the percept of the deep, blue California sky, the bouquet associated with a rich Bordeaux, a haunting musical melody and so on. These must be caused by specific neural activity in cortex, thalamus, basal ganglia and associated neuronal structures. The question motivating much of the current research into the neuronal basis of consciousness is the notion of the minimal neural activity that is sufficient to cause a specific conscious percept or memory.

For instance, when a subject consciously perceives a face, the retinal ganglion cells whose axons make up the optic nerve that carries the visual information to the brain proper are firing in response to the visual stimulus. Yet it is unlikely that this retinal activity directly correlates with visual perception. While such activity is evidently necessary for seeing a physical stimulus in the world, retinal neurons by themselves do not give rise to consciousness.

Given the comparative ease with which the brains of animals can be probed and manipulated, it seems opportune at this point in time to concentrate on the neural basis of sensory consciousness. Because primates are highly visual animals and much is known about the neuroanatomy, psychology and computational principles underling visual perception, visions has proven to be the most popular model systems in the brain sciences.

Cognitive and clinical research demonstrates that much complex information processing can occur without involving consciousness. This includes visual, auditory and linguistic priming, implicit memory, the implicit recognition of complex sequences, automatic behaviours such as driving a car or riding a bicycle and so on (Velmans 1991). The dissociations found in patients with lesions in the cerebral cortex (e.g., such as residual visual functions in the professed absence of any visual awareness known as clinical blind-sight in patients with lesions in preliminary visual cortex.

It can be said, that if one is without idea, then one is without concept, and as well, if one is without concept one is without an idea. An idea (Gk., eidos, visible form) be it a notion stretching all the way from one pole, where it denotes a subjective, internal presence in the mind, somehow thought of as representing something about the world, to the other pole, where it represents an eternal, timeless unchanging form or concept: The concept o the number series or of justice, for example, thought of as independent objected of enquiry and perhaps of knowledge. These two poles are not distinct meanings of the therm, although they give rise to many problems of interpretation, but between tem they define a space of philosophical problems. On the other hand, ideas are that with which er think, or in Locke’s terms, whatever the mind may be employed about in thinking. Looked at that way they seem to be inherently transient, fleeting, and unstable private presences. On the other hand, ideas provide the way in which objective knowledge can be expressed. They are the essential components of understanding, and any intelligible proposition that is true must be capable of being understood. Plato’s theory of Forms is a celebration of objective and timeless existence of ideas as concepts, and in his hands ideas are reified to the point where they make up the only rea world, of separate and perfect models of which the empirical world is only a poor cousin. This doctrine, notable in the Timaeus, opened the way for the Neoplatonic notion of ideas as the thoughts of God. The concept gradually lost this other-worldly aspect until after Descartes ideas become assimilated to whatever it is that lies in the mind of any thinking being.

The philosophical doctrine that reality is somehow mind-correlatives or mind co-ordinated - that the real objects comprising the ‘external world’ are mot independent of cognizing minds, but only exist as in some way correlative to the mental operations. The doctrine centres on the conception that reality as we understand it reflects the working of mind. And it construes this as meaning that the inquiring mind itself to make a formative contribution not merely to our understanding character we attribute to it.

The cognitive scientist Jackendoff (1987) argues at length against the notion that consciousness and thoughts are inseparable and that introspection can reveal the contents of the mind. What is conscious about thoughts, are sensory aspects, such as visual images, sounds or silent speech? Both the process of thought and its content are not directly accessible to consciousness. Indeed, one tradition in psychology and psychoanalysis - going back to Sigmund Freud-hypothesizes that higher-level decision making and creativity are not accessible at a conscious level, although they influence behaviour.

Within the visual modality, Milner and Goodale (1995) have made a masterful case for the existence of so-called on-line systems that by-pass consciousness. Their function is to mediate relative stereotype visuo-motor behaviours, such as eye and arm movements, reaching, grasping, and postural adjustment and so on. In a very rapid, reflex-like manner. On-line systems work in egocentric coordinate systems, and lack certain types of perceptual illusions (e.g., size illusion) as well as direct access to working memory. These contrasts are well within the function of consciousness as alluded to from above, namely to synthesize information from many different sources and use it to plan behavioural patterns over time. Milner and Goodale argue that on-line systems are associated with the dorsal stream of visual information in the cerebral cortex, originating in the primary visual cortex and terminating in the posterior parietal cortex. The problem of consciousness can be broken down into several separate questions. Most, if not all of these, can then be subjected to scientific inquiry.

The major question that neuroscience must ultimately answer can be bluntly stated as follows: It is probable that at any moment some active neuronal processes in our head correlates with consciousness, while others do not; what is the difference between them? The specific processes that correlate with the current content of consciousness are referred to as the neuronal correlate of consciousness, or as the NCC. Whenever some information is represented in the NCC, it is represented in consciousness. The NCC is the minimal (minimal, since it is known that the entire brain is sufficient to give rise to consciousness) set of neurons, most likely distributed throughout certain cortical and subcortical areas, whose firing directly correlates with the perception of the subject at the time. Conversely, stimulating these neurons in the right manner with some yet unheard of technology should give rise to the same perception as before.

Discovering the NCC and its properties will mark a major milestone in any scientific theory of consciousness.

What is the character of the NCC? Most popular has been the belief that consciousness arises as an emergent property of a very large collection of interacting neurons (for instance, Libet 1993). In this view, it would be foolish to locate consciousness at the level of individual neurons. An alternative hypothesis is that there are special sets of ‘consciousness’ neurons distributed throughout cortex and associated systems. Such neurons represent the ultimate neuronal correlate of consciousness, in the sense that the relevant activity of an appropriate subset of them is both necessary and sufficient to give rise to an appropriate conscious experience or percept (Crick and Koch 1998). Generating the appropriate activity in these neurons, for instance by suitable electrical stimulation during open skull surgery, would give rise to the specific percept.

Any-one subtype of NCC neurons would, most likely, be characterized by a unique combination of molecular, biophysical, pharmacological and anatomical traits. It is possible, of course, that all cortical neurons may be capable of participating in the representation of one percept or another, though not necessarily doing so for all percepts. The secret of consciousness would then be the type of activity of a temporary subset of them, consisting of all those cortical neurons that represent that particular percept at that moment. How activity of neurons across a multitude of brain areas that encode all of the different aspects associated with an object (e.g., the colour of the face, its facial expression, its gender and identity, the sound issuing from its mouth) is combined into some single percept remains puzzling and is known as the binding problem.

What, if anything, can we infer about the location of neurons whose activity correlates with consciousness? In the case of visual consciousness, it was surmised that these neurons must have access to visual information and project to the planning stages of the brain; That is to premotor and frontal areas. Since no neurons in the primary visual cortex of the macaque monkey project to any area forward of the central sulcus, Crick and Koch (1998) propose that neurons in V1 do not give rise to consciousness (although it is necessary for most forms of vision, just as the retina is). Ongoing electro physiological, psycho physical and imaging research in monkeys and humans is evaluating this prediction.

While the set of neurons that can express anyone particular conscious percept might constitute a relative small fraction of all neurons in anyone area, many more neurons might be necessary to support the firing activity leading up to the NCC. This might resolve the apparent paradox between clinical lessoning data suggesting that small and discrete lesions in the cortex can lead to very specific deficits (such as the inability to see colours or to recognize faces in the absence of other visual losses) and the functional imaging data that anyone visual stimulus can activate large swaths of cortex.

Conceptually, several other questions need to be answered about the NCC. What type of activity corresponds to the NCC (it has been proposed as long ago as the early part of the twentieth century that spiking activity synchronized across a population of neurons is a necessary condition for consciousness to occur)? What causes the NCC to occur? And, finally, what effect does the NCC have on postsynaptic structures, including motor output.

A promising experimental approach to locate the NCC is the use of bistable percepts in which a constant retinal stimulus gives rise to two percepts alternating in time, as in a Necker cube (Logothetis 1998). One version of this is binocular rivalry in which small images, say of a horizontal grating, are presented to the left eye and another image, say the vertical grating is shown to the corresponding location in the right eye. In spite of the constant visual stimulus, observers ‘see’ the horizontal grating alternately every few seconds with the vertical one (Blake 1989). The brain does not allow for the simultaneous perception of both images.

It is possible, though difficult, to train a macaque monkey to report whether it is currently seeing the left or the right image. The distribution of the switching times and the way in which changing the contrast in one eye affects these leaves little to doubt, in that monkeys and humans experience the same basic phenomenon. In a series of elegant experiments, Logothetis and colleagues (Logothetis 1998) recorded from a variety of visual cortical areas in the awake macaque monkey while the animal performed a binocular rivalry task. In undeveloped visual cortices, only a small fraction of cells modulates their response as a function of the percept of the monkey, while 20 to 30% of neurons in higher visual areas in the cortex do so. The majority of cells increased their firing rate in response to one or the other retinal stimulus with little regard to what the animal perceives at the time. In contrast, in a high-level cortical area such as the inferior temporal cortex, almost all neurons responded only to the perceptual dominant stimulus (in other words, a ‘face’ cell only fired when the animal indicated by its performance that it saw the face and not the pattern presented to the other eye). This makes it likely that the NCC involves activity in neurons in the inferior temporal lobe. Lesions in the homologous area in the human brain are known to cause very specific deficits in the conscious face or object recognition. However, it is possible that specific interactions between IT cells and neurons in parts of the prefrontal cortex are necessary in order for the NCC to be generated

Functional brain imaging in humans undergoing binocular rivalry has revealed that areas in the right prefrontal cortex are in activating during the perceptual switch from one percept to the other.

A number of alternate experimental paradigms are being investigated using electro physiological recordings of individual neurons in behaving animals and human patients, combined with functional brain imaging. Common to these is the manipulation of the complex and changing relationship between physical stimulus and the conscious percept. For instance, when subjects are forced rapidly to respond to a low saliency target, both monkeys and human’s sometimes claim to perceive such a target in the absence of any physical target consciously (false alarm) or fail to respond to a target (miss). The NCC in the appropriate sensory area should mirror the perceptual report under these dissociated conditions. Visual illusions constitute another rich source of experiments that can provide information concerning the neurons underlying these illusory percepts. A classical example is the motion affected in which a subject stares at a constantly moving stimulus (such as a waterfall) for a fraction of a minute or longer. Immediately after this conditioning period, a stationary stimulus will appear to move in the opposite direction. Because of the conscious experience of motion, one would expect, the subject’s cortical motion areas to be activated in the absence of any moving stimulus.

Future techniques, most likely based on the molecular identification and manipulation of discrete and identifiable subpopulations of cortical cells in appropriate animals, will greatly help in this endeavour

Identifying the type of activity and the type of neurons that gives rise to specific conscious percept in animals and humans would only be the first, even if critical, step in understanding consciousness. One also needs to know where these cells project to, their postsynaptic action, how they develop in early childhood, what happens to them in mental diseases known to affect consciousness in patients, such as schizophrenia or autism, and so on. And, of course, a final theory of consciousness would have to explain the central mystery, why a physical system with particular architectures gives rise to feelings and qualia.

The central structure of an experience is its intentionality, its being directed toward something, as it is an experience of or about some object. An experience is directed toward an object by virtue of its content or meaning (which represents the object) together with appropriate enabling conditions.

Phenomenology as a discipline is distinct from but related to other key disciplines in philosophy, such as ontology, epistemology, logic, and ethics. Phenomenology has been practised in various guises for centuries, however, its maturing qualities have begun in the early parts of the 20th century. The works that have dramatically empathized the growths of phenomenology are accredited through the works of Husserl, Heidegger, Sartre, Merleau-Ponty and others. Phenomenological issues of intentionality, consciousness, qualia, and first-person perspective have been prominent in recent philosophy of mind.

Phenomenology is commonly understood in either of two ways: as a disciplinary field in philosophy, or as a movement in the history of philosophy.

The discipline of phenomenology may be defined initially as the study of structures of experience, or consciousness. Literally, phenomenology is the study of ‘phenomena’: Appearances of things, or things as they appear in our experience, or the ways we experience things, thus the meaning’s things have in our experience. Phenomenology studies conscious experience as experienced from the subjective or first person point of view. This field of philosophy is then to be distinguished from, and related to, the other main fields of philosophy: Ontology (the study of being or what is), epistemology (the study of knowledge), logic (the study of valid reasoning), ethics (the study of right and wrong action), etc.

The historical movement of phenomenology is the philosophical tradition launched in the first half of the 20th century by Edmund Husserl, Martin Heidegger, Maurice Merleau-Ponty, Jean-Paul Sartre. In that movement, the discipline of phenomenology was prized as the proper foundation of all philosophy - as opposed, say, to ethics or metaphysics or epistemology. The methods and characterization of the discipline were widely debated by Husserl and his successors, and these debates continue to the present day. (The definitions of Phenomenological offered above will thus be debatable, for example, by Heideggerians, but it remains the starting point in characterizing the discipline.)

In recent philosophy of mind, the term ‘phenomenology’ is often restricted to the characterization of sensory qualities of seeing, hearing, etc.: What it is like to have sensations of various kinds. However, our experience is normally much richer in content than mere sensation. Accordingly, in the Phenomenological tradition, phenomenology is given a much wider range, addressing the meaning things have in our experience, notably, the significance of objects, events, tools, the flow of time, the self, and others, as these things arise and are experienced in our ‘life-world.’

Phenomenology as a discipline has been central to the tradition of continental European philosophy throughout the 20th century, while philosophy of mind has evolved in the Austro-Anglo-American tradition of analytic philosophy that developed throughout the 20th century. Yet the fundamental character of our mental activity is pursued in overlapping ways within these two traditions. Accordingly, the perspective on phenomenology drawn in this article will accommodate both traditions. The main concern here will be to characterize the discipline of phenomenology, in contemporary views, while also highlighting the historical tradition that brought the discipline into its own.

Basically, phenomenology studies the structure of various types of experience ranging from perception, thought, memory, imagination, emotion, desire, and volition to bodily awareness, embodied action, and social activity, including linguistic activity. The structure of these forms of experience typically involves what Husserl called ‘intentionality,’ that is, the directedness of experience toward things in the world, the property of consciousness that it is a consciousness of or about something. According to classical Husserlian phenomenology, our experiences abide toward the direction that represents or ‘intends’ of things only through particular concepts, thoughts, ideas, images, etc. These make up the meaning or content of a given experience, and are distinct from the things they present or mean.

The basic intentional structure of consciousness, we come to find in reflection or analysis, in that of which involves further forms of experience. Thus, phenomenology develops a complex account of temporal awareness (within the stream of consciousness), spatial awareness (notably in perception), attention (distinguishing focal and marginal or ‘horizonal’ awareness), awareness of one's own experience (self-consciousness, in one sense), self-awareness (awareness-of-oneself), the self in different roles (as thinking, acting, etc.), embodied action (including kinesthetic awareness of one's movement), determination or intention represents its desire for action (more or less explicit), awareness of other persons (in empathy, intersubjectivity, collectivity), linguistic activity (involving meaning, communication, understanding others), social interaction (including collective action), and everyday activity in our surrounding life-world (in a particular culture).

Furthermore, in a different dimension, we find various grounds or enabling conditions - conditions of the possibility - of intentionality, including embodiment, bodily skills, cultural context, language and other social practices, social background, and contextual aspects of intentional activities. Thus, phenomenology leads from conscious experience into conditions that help to give experience its intentionality. Traditional phenomenology has focussed on subjective, practical, and social conditions of experience. Recent philosophy of mind, however, has focussed especially on the neural substrate of experience, on how conscious experience and mental representation or intentionality is grounded in brain activity. It remains a difficult question how much of these grounds of experience fall within the province of phenomenology as a discipline. Cultural conditions thus seem closer to our experience and to our familiar self-understanding than do the electrochemical workings of our brain, much less our dependence on quantum-mechanical states of physical systems to which we may belong. The cautious thing to say is that phenomenology leads in some ways into at least some background conditions of our experience.

The discipline of phenomenology is defined by its domain of study, its methods, and its main results. Phenomenology studies structures of conscious experience as experienced from the first-person point of view, along with relevant conditions of experience. The central structure of an experience is its intentionality, the way it is directed through its content or meaning toward a certain object in the world.

We all experience various types of experience including perception, imagination, thought, emotion, desire, volition, and action. Thus, the domain of phenomenology is the range of experiences including these types (among others). Experience includes not only relatively passive experience as in vision or hearing, but also active experience as in walking or hammering a nail or kicking a ball. (The range will be specific to each species of being that enjoys consciousness; Our focus is on our own human experience. Not all conscious beings will, or will be able to, practice phenomenology, as we do.)

Conscious experiences have a unique feature: we experience them, we live through them or perform them. Other things in the world we may observe and engage. But we do not experience them, in the sense of living through or performing them. This experiential or first-person feature - that of being experienced - is an essential part of the nature or structure of conscious experience: as we say, ‘I see/think/desire/do . . .’ This feature is both a Phenomenological and an ontological feature of each experience: it is part of what it is for the experience to be experienced (Phenomenological) and part of what it is for the experience to be (ontological).

How shall we study conscious experience? We reflect on various types of experiences just as we experience them. That is to say, we proceed from the first-person point of view. However, we do not normally characterize an experience at the time we are performing it. In many cases we do not have that capability: a state of intense anger or fear, for example, consumes the entire focus at the time. Rather, we acquire a background of having lived through a given type of experience, and we look to our familiarity with that type of experience: hearing a song, seeing a sunset, thinking about love, intending to jump a hurdle. The practice of phenomenology assumes such familiarity with the type of experiences to be characterized. Importantly, also, it is types of experience that phenomenology pursues, rather than a particular fleeting experience - unless its type is what interests us.

Classical phenomenologists practised some three distinguishable methods. (1) We describe a type of experience just as we find it in our own (past) experience. Thus, Husserl and Merleau-Ponty spoke of pure description of lived experience. (2) We interpret a type of experience by relating it to relevant features of context. In this vein, Heidegger and his followers spoke of hermeneutics, the art of interpretation in context, especially social and linguistic context. (3) We analyse the form of a type of experience. In the end, all the classical phenomenologists practised analysis of experience, factoring out notable features for further elaboration.

These traditional methods have been ramified in recent decades, expanding the methods available to phenomenology. Thus: (4) In a logico-semantic model of phenomenology, we specify the truth conditions for a type of thinking (say, where I think that dogs chase cats) or the satisfaction conditions for a type of intention (say, where I intend or will to jump that hurdle). (5) In the experimental paradigm of cognitive neuroscience, we design empirical experiments that tend to confirm or refute aspects of experience (say, where a brain scan shows electrochemical activity in a specific region of the brain thought to subserve a type of vision or emotion or motor control). This style of ‘neurophenomenology’ assumes that conscious experience is grounded in neural activity in embodied action in appropriate surroundings - mixing pure phenomenology with biological and physical science in a way that was not wholly congenial to traditional phenomenologists.

What makes an experience conscious is a certain awareness one has of the experience while living through or performing it. This form of inner awareness has been a topic of considerable debate, centuries after the issue arose with Locke's notion of self-consciousness on the heels of Descartes' sense of consciousness (conscience, co-knowledge). Does this awareness-of-experience consist in a kind of inner observation of the experience, as if one were doing two things at once? (Brentano argued no.) Is it a higher-order perception of one's mind's operation, or is it a higher-order thought about one's mental activity? (Recent theorists have proposed both.) Or is it a different form of inherent structure? (Sartre took this line, drawing on Brentano and Husserl.) These issues are beyond the scope of this article, but notice that these results of Phenomenological analysis, that shape the characterlogical domain of study and the methodology appropriate to the domain. For awareness-of-experience is a defining trait of conscious experience, the trait that gives experience a first-person, lived character. It is that lived character of experience that allows a first-person perspective on the object of study, namely, experiences, and that perspective is characteristic of the methodology of phenomenology.

Conscious experience is the starting point of phenomenology, but experience shades off into fewer overtly conscious phenomena. As Husserl and others stressed, we are only vaguely aware of things in the margin or periphery of attention, and we are only implicitly aware of the wider horizon of things in the world around us. Moreover, as Heidegger stressed, in practical activities like walking along, or hammering a nail, or speaking our native tongue, we are not explicitly conscious of our habitual patterns of action. Furthermore, as psychoanalysts have stressed, much of our intentional mental activity is not conscious at all, but may become conscious in the process of therapy or interrogation, as we come to realize how we feel or think about something. We should allow, then, that the domain of phenomenology - our own experience - spreads out from conscious experience into semi-conscious and even unconscious mental activity, along with relevant background conditions implicitly invoked in our experience. (These issues are subject to debate; the point here is to open the door to the question of where to draw the boundary of the domain of phenomenology.)

To begin an elementary exercise in phenomenology, consider some typical experiences one might have in everyday life, characterized in the first person: (1) I see that fishing boat off the coast as dusk descends over the Pacific. (2) I hear that helicopter whirring overhead as it approaches the hospital. (3) I am thinking that phenomenology differs from psychology. (4) I wish that warm rain from Mexico were falling like last week. (5) I imagine a fearsome creature like that in my nightmare. (6) I intend to finish my writing by noon. (7) I walk carefully around the broken glass on the sidewalk. (8) I stroke a backhand cross-court with that certain underspin. (9) I am searching for the words to make my point in conversation.

Here are rudimentary characterizations of some familiar types of experience. Each sentence is a simple form of Phenomenological description, articulating in everyday English the structure of the type of experience so described. The subject term ‘I’ indicate the first-person structure of the experience: The intentionality proceeds from the subject. The verb indicates the type of intentional activity describing recognition, thought, imagination, etc. Of central importance is the way that objects of awareness are presented or intended in our experiences, especially, the way we see or conceive or think about objects. The direct-object expression (‘that fishing boat off the coast’) articulates the mode of presentation of the object in the experience: the content or meaning of the experience, the core of what Husserl called noema. In effect, the object-phrase expresses the noema of the act described, that is, to the extent that language has appropriate expressive power. The overall form of the given sentence articulates the basic form of intentionality in the experience: Subject-act-content-object.

Rich Phenomenological description or interpretation, as in Husserl, Merleau-Ponty et al., will far outrun such simple Phenomenological descriptions as above. But such simple descriptions bring out the basic form of intentionality. As we interpret the Phenomenological description further, we may assess the relevance of the context of experience. And we may turn to wider conditions of the possibility of that type of experience. In this way, in the practice of phenomenology, we classify, describe, interpret, and analyse structures of experiences in ways that answer to our own experience.

In such interpretive-descriptive analyses of experience, we immediately observe that we are analysing familiar forms of consciousness, conscious experience of or about this or that. Intentionality is thus the salient structure of our experience, and much of the phenomenology proceeds as the study of different aspects of intentionality. Thus, we explore structures of the stream of consciousness, the enduring self, the embodied self, and bodily action. Furthermore, as we reflect on how these phenomena work, we turn to the analysis of relevant conditions that enable our experiences to occur as they do, and to represent or intend as they do. Phenomenology then leads into analyses of conditions of the possibility of intentionality, conditions involving motor skills and habits, backgrounding social practices, and often language, with its special place in human affairs, presents the following definition: ‘Phoneme, . . .’

The Oxford English Dictionary indicated of its knowledge, where science as itself is a contained source of phenomena as distinct from being (ontology). That division of any science that describes and classifies its phenomena. From the Greek phainomenon, appearance. In philosophy, the term is used in the first sense, amid debates of theory and methodology. In physics and philosophy of science, the term is used in the second sense, but only occasionally.

So its root meaning, then, phenomenology is the study of phenomena: Literally, appearances as opposed to reality. This ancient distinction launched philosophy as we emerged from Plato's cave. Yet the discipline of phenomenology did not blossom until the 20th century and remains poorly understood in many circles of contemporary philosophy. What is that discipline? How did philosophy move from a root concept of phenomena to the discipline of phenomenology?

Originally, in the 18th century, ‘phenomenology’ meant the theory of appearances fundamental to empirical knowledge, especially sensory appearances. The term seems to have been introduced by Johann Heinrich Lambert, a follower of Christian Wolff. Subsequently, Immanuel Kant used the term occasionally in various writings, as did Johann Gottlieb Fichte and G. W. F. Hegel. By 1889 Franz Brentano used the term to characterize what he called ‘descriptive psychology. From there Edmund Husserl took up the term for his new science of consciousness, and the rest is history.

Suppose we say phenomenology study’s phenomena: Of what appears to us - and its appearing. How shall we understand phenomena? The term has a rich history in recent centuries, in which we can see traces of the emerging discipline of phenomenology.

In a strict empiricist vein, what appears before the mind accedes of sensory data or qualia: either patterns of one's own sensations (seeing red here now, feeling this ticklish feeling, hearing that resonant bass tone) or sensible patterns of worldly things, say, the looks and smells of flowers (what John Locke called secondary qualities of things). In a strict rationalist vein, by contrast, what appears before the mind of ideas, rationally formed ‘clear and distinct ideas’ (in René Descartes' ideal). In Immanuel Kant's theory of knowledge, fusing rationalist and empiricist aims, what appears to the mind are phenomena defined as things-as-they-appear or things-as-they-are-represented (in a synthesis of sensory and conceptual forms of objects-as-known). In Auguste Comte's theory of science, phenomena (phenomenes) are the facts (faits, what occurs) that a given science would explain.

In 18th and 19th century epistemology, then, phenomena are the starting points in building knowledge, especially science. Accordingly, in a familiar and still current sense, phenomena are whatever we observe (perceive) and seek to explain. Discipline of psychology emerged late in the 19th century, however, phenomena took on a somewhat different guise. In Franz Brentano's Psychology from an Empirical Standpoint (1874), phenomena are of what is to occur in the mind: Mental phenomena are acts of consciousness (or their contents), and physical phenomena are objects of external perception starting with colours and shapes. For Brentano, physical phenomena exist ‘intentionally’ in acts of consciousness. This view revives a Medieval notion Brentano called ‘intentional in-existence. Nevertheless, the ontology remains undeveloped (what is it to exist in the mind, and do physical objects exist only in the mind?). More generally, we might say that phenomena are whatever we are conscious of: objects and events around us, other people, ourselves, even (in reflection) our own conscious experiences, as we experience these. In a certain technical sense, phenomena are things as they are given to our consciousness, whether in perception or imagination or thought or volition. This conception of phenomena would soon inform the new discipline of phenomenology.

Brentano distinguished descriptive psychology from genetic psychology. Where genetic psychology seeks the causes of various types of mental phenomena, descriptive psychology defines and classifies the various types of mental phenomena, including perception, judgment, emotion, etc. According to Brentano, every mental phenomenon, or act of consciousness, is directed toward some object, and only mental phenomena are so directed. This thesis of intentional directedness was the hallmark of Brentano's descriptive psychology. In 1889 Brentano used the term ‘phenomenology’ for descriptive psychology, and the way was paved for Husserl's new science of phenomenology.

Phenomenology as we know it was launched by Edmund Husserl in his Logical Investigations (1900-01). Two importantly different lines of theory came together in that monumental work: Psychological theory, on the heels of Franz Brentano (and William James, whose Principles of Psychology appeared in 1891 and greatly impressed Husserl); Its logically semantic theory, are the heels of Bernard Bolzano and Husserl's contemporaries who founded modern logic, including Gottlob Frége. (Interestingly, both lines of research trace back to Aristotle, and both reached importantly new results in Husserl's day.)

Husserl's Logical Investigations was inspired by Bolzano's ideal of logic, while taking up Brentano's conception of descriptive psychology. In his Theory of Science (1835) Bolzano distinguished between subjective and objective ideas or representations (Vorstellungen). In effect Bolzano criticized Kant and before him the classical empiricists and rationalists for failing to make this sort of distinction, thereby rendering phenomena merely subjective. Logic studies objective ideas, including propositions, which in turn make up objective theories as in the sciences. Psychology would, by contrast, study subjective ideas, the concrete contents (occurrences) of mental activities in particular minds at a given time. Husserl was after both, within a single discipline. So phenomena must be reconceived as objective intentional contents (sometimes called intentional objects) of subjective acts of consciousness. Phenomenology would then study this complex of consciousness and correlated phenomena. In Ideas I (Book One, 1913) Husserl introduced two Greek words to capture his version of the Bolzanoan distinction: noesis and noema (from the Greek verb noéaw, meaning to perceive, thinks, intend, from where the noun nous or mind). The intentional process of consciousness is called noesis, while its ideal content is called noema. The noema of an act of consciousness Husserl characterized both as an ideal meaning and as ‘the object as intended.’ Thus the phenomenon, or object-as-it-appears, becomes the noema, or object-as-it-is-intended. The interpretations of Husserl's theory of noema have been several and amount to different developments of Husserl's basic theory of intentionality. (Is the noema an aspect of the object intended, or rather a medium of intention?)

For Husserl, then, phenomenology integrates a kind of psychology with a kind of logic. It develops a descriptive or analytic psychology in that it describes and Analysed types of subjective mental activity or experience, in short, act of consciousness. Yet it develops a kind of logic - a theory of meaning (today we say logical semantics) - in that it describes and Analysed objective contents of consciousness: Ideas, concepts, images, propositions, in short, ideal meanings of various types that serve as intentional contents, or noematic meanings, of various types of experience. These contents are shareable by different acts of consciousness, and in that sense they are objective, ideal meanings. Following Bolzano (and to some extent the platonistic logician Hermann Lotze), Husserl opposed any reduction of logic or mathematics or science to mere psychology, to how the public happens to think, and in the same spirit he distinguished phenomenology from mere psychology. For Husserl, phenomenology would study consciousness without reducing the objective and shareable meanings that inhabit experience to merely subjective happenstances. Ideal meaning would be the engine of intentionality in acts of consciousness.

A clear conception of phenomenology awaited Husserl's development of a clear model of intentionality. Indeed, phenomenology and the modern concept of intentionality emerged hand-in-hand in Husserl's Logical Investigations (1900-01). With theoretical foundations laid in the Investigations, Husserl would then promote the radical new science of phenomenology in Ideas I (1913). And alternative visions of phenomenology would soon follow.

Phenomenology matured and was nurtured through the works of Husserl, much as epistemology came about by means of its own nutrition but through Descartes study, and ontology or metaphysics came into its own with Aristotle on the heels of Plato. Yet phenomenology has been practised, with or without the name, for many centuries. When Hindu and Buddhist philosophers reflected on states of consciousness achieved in a variety of meditative states, they were practising phenomenology. When Descartes, Hume, and Kant characterized states of perception, thought, and imagination, they were practising phenomenology. When Brentano classified varieties of mental phenomena (defined by the directedness of consciousness), he was practising phenomenology. When William James appraised kinds of mental activity in the stream of consciousness (including their embodiment and their dependence on habit), he too was practising phenomenology. And when recent analytic philosophers of mind have addressed issues of consciousness and intentionality, they have often been practising phenomenology. Still, the discipline of phenomenology, its roots tracing back through the centuries, came full to flower in Husserl.

Husserl's work was followed by a flurry of Phenomenological writing in the first half of the 20th century. The diversity of traditional phenomenology is apparent in the Encyclopaedic of Phenomenology (Kluwer Academic Publishers, 1997, Dordrecht and Boston), which features separate articles on some seven types of phenomenology. (1) Transcendental constitutive phenomenology studies how objects are constituted in pure or transcendental consciousness, setting aside questions of any relation to the natural world around us. (2) Naturalistic constitutive phenomenology studies how consciousness constitutes or takes things in the world of nature, assuming with the natural attitude that consciousness is part of nature. (3) Existential phenomenology studies concrete human existence, including our experience of free choice or action in concrete situations. (4) Generative historicist phenomenology studies how meaning, as found in our experience, is generated in historical processes of collective experience over time. (5) Genetic phenomenology studies the genesis of meanings of things within one's own stream of experience. (6) Hermeneutical phenomenology studies interpretive structures of experience, how we understand and engage things around us in our human world, including ourselves and others. (7) Realistic phenomenology studies the structure of consciousness and intentionality, assuming it occurs in a real world that is largely external to consciousness and not somehow brought into being by consciousness.

The most famous of the classical phenomenologists were Husserl, Heidegger, Sartre, and Merleau-Ponty. In these four thinkers we find different conceptions of phenomenology, different methods, and different results. A brief sketch of their differences will capture both a crucial period in the history of phenomenology and a sense of the diversity of the field of phenomenology.

In his Logical Investigations (1900-01) Husserl outlined a complex system of philosophy, moving from logic to philosophy of language, to ontology (theory of universals and parts of wholes), to a Phenomenological theory of intentionality, and finally to a Phenomenological theory of knowledge. Then in Ideas I (1913) he focussed squarely on phenomenology itself. Husserl defined phenomenology as ‘the science of the essence of consciousness,’ entered on the defining trait of intentionality, approached explicitly ‘in the first person.’ In this spirit, we may say phenomenology is the study of consciousness - that is, conscious experience of various types - as experienced from the first-person point of view. In this discipline we study different forms of experience just as we experience them, from the perspective of the subject living through or performing them. Thus, we characterize experiences of seeing, hearing, imagining, thinking, feeling (i.e., emotion), wishing, desiring, willing, and acting, that is, embodied volitional activities of walking, talking, cooking, carpentering, etc. However, not just any characterization of an experience will do. Phenomenological analysis of a given type of experience will feature the ways in which we ourselves would experience that form of conscious activity. And the leading property of our familiar types of experience is their intentionality, their being a consciousness of or about something, something experienced or presented or engaged in a certain way. How I see or conceptualize or understand the object I am dealing with defines the meaning of that object in my current experience. Thus, phenomenology features a study of meaning, in a wide sense that includes more than what is expressed in language.

In Ideas I Husserl presented phenomenology with a transcendental turn. In part this means that Husserl took on the Kantian idiom of ‘transcendental idealism,’ looking for conditions of the possibility of knowledge, or of consciousness generally, and arguably turning away from any reality beyond phenomena. But Husserl's transcendental, turn also involved his discovery of the method of epoché (from the Greek skeptics' notion of abstaining from belief). We are to practice phenomenology, Husserl proposed, by ‘bracketing’ the question of the existence of the natural world around us. We thereby turn our attention, in reflection, to the structure of our own conscious experience. Our first key result is the observation that each act of consciousness is a consciousness of something, that is, intentional, or directed toward something. Consider my visual experience wherein I see a tree across the square. In Phenomenological reflection, we need not concern ourselves with whether the tree exists: my experience is of a tree whether or not such a tree exists. However, we do need to concern ourselves with how the object is meant or intended. I see a Eucalyptus tree, not a Yucca tree; I see that object as a referentially exposed Eucalyptus tree, with certain shape and with bark stripping off, etc. Thus, bracketing the tree itself, we turn our attention to my experience of the tree, and specifically to the content or meaning in my experience. This tree-as-perceived Husserl calls the noema or noematic sense of the experience.

Philosophers succeeding Husserl debated the proper characterization of phenomenology, arguing over its results and its methods. Adolf Reinach, an early student of Husserl's (who died in World War I), argued that phenomenology should remain merged with a total inference by some realistic ontologism, as in Husserl's Logical Investigations. Roman Ingarden, a Polish phenomenologist of the next generation, continued the resistance to Husserl's turn to transcendental idealism. For such philosophers, phenomenology should not bracket questions of being or ontology, as the method of epoché would suggest. And they were not alone. Martin Heidegger studied Husserl's early writings, worked as Assistant to Husserl in 1916, and in 1928, succeeded Husserl in the prestigious chair at the University of Freiburg. Heidegger had his own ideas about phenomenology.

In Being and Time (1927) Heidegger unfurled his rendition of phenomenology. For Heidegger, we and our activities are always ‘in the world,’ our being is being-in-the-world, so we do not study our activities by bracketing the world, rather we interpret our activities and the meaning things have for us by looking to our contextual relations to things in the world. Indeed, for Heidegger, phenomenology resolves into what he called ‘fundamental ontology.’ We must distinguish beings from their being, and we begin our investigation of the meaning of being in our own case, examining our own existence in the activity of ‘Dasein’ (that being whose being is in each case my own). Heidegger resisted Husserl's neo-Cartesian emphasis on consciousness and subjectivity, including how perception presents things around us. By contrast, Heidegger held that our more basic ways of relating to things are in practical activities like hammering, where the phenomenology reveals our situation in a context of equipment and in being-with-others.

In Being and Time Heidegger approached phenomenology, in a quasi-poetic idiom, through the root meanings of ‘logos’ and ‘phenomena,’ so that phenomenology is defined as the art or practice of ‘letting things show themselves.’ In Heidegger's inimitable linguistic play on the Greek roots, ‘phenomenology’ means . . . - to let that which shows itself to be seen from itself in the very way in which it shows itself from itself. Here Heidegger explicitly parodies Husserl's call, ‘To the things themselves,’ or ‘To the phenomena themselves!’ Heidegger went on to emphasize practical forms of comportment or better relating (Verhalten) as in hammering a nail, as opposed to representational forms of intentionality as in seeing or thinking about a hammer. Much, of Being and Time develops an existential interpretation of our modes of being including, famously, our being-toward-death.

In a very different style, in clear analytical prose, in the text of a lecture course called The Basic Problems of Phenomenology (1927), Heidegger traced the question of the meaning of being from Aristotle through many other thinkers into the issues of phenomenology. Our understanding of beings and their being comes ultimately through phenomenology. Here the connection with classical issues of ontology is more apparent, and consonant with Husserl's vision in the Logical Investigations (an early source of inspiration for Heidegger). One of Heidegger's most innovative ideas was his conception of the ‘ground’ of being, looking to modes of being more fundamental than the things around us (from trees to hammers). Heidegger questioned the contemporary concern with technology, and his writing might suggest that our scientific theories are historical artifacts that we use in technological practice, rather than systems of ideal truth (as Husserl had held). Our deep understanding of being, in our own case, comes rather from phenomenology, Heidegger held.

In the 1930s phenomenology migrated from Austrian and then German philosophy into French philosophy. The way had been paved in Marcel Proust's in Search of Lost Time, in which the narrator recounts in close detail his vivid recollections of experiences, including his famous associations with the smell of freshly baked madeleines. This sensibility to experience traces to Descartes' work, and French phenomenology has been an effort to preserve the central thrust of Descartes' insights while rejecting mind-body dualism. The experience of one's own body, or one's lived or living body, has been an important motif in many French philosophers of the 20th century

In the novel Nausea (1936) Jean-Paul Sartre described a bizarre course of experience in which the protagonist, writing in the first person, describes how ordinary objects lose their meaning until he encounters pure being at the foot of a chestnut tree, and in that moment recovers his sense of his own freedom. In Being and Nothingness (1943, written partly while a prisoner of war), Sartre developed his conception of Phenomenological ontology. Consciousness is a consciousness of objects, as Husserl had stressed. In Sartre's model of intentionality, the central player in consciousness is a phenomenon, and the occurrence of a phenomenon is just a consciousness-of-an-object. The chestnut tree I see is, for Sartre, such a phenomenon in my consciousness. Indeed, all things in the world, as we normally experience them, are phenomena, beneath or behind which lies their ‘being-in-itself.’ Consciousness, by contrast, has ‘being-for-itself,’ inasmuch as consciousness is not only a consciousness-of-its-object but also a pre-reflective consciousness-of-itself (conscience de soi). Yet for Sartre, unlike Husserl, that ‘I’ or self is nothing but a sequence of acts of consciousness, notably including radically free choices (like a Humean bundle of perceptions).

For Sartre, the practice of phenomenology proceeds by a deliberate reflection on the structure of consciousness. Sartre's method is in effect a literary style of interpretive description of different types of experience in relevant situations - a practice that does not really fit the methodological proposals of either Husserl or Heidegger, but makes use of Sartre's great literary skill. (Sartre wrote many plays and novels and was awarded the Nobel Prize in Literature.)

Sartre's phenomenology in Being and Nothingness became the philosophical foundation for his popular philosophy of existentialism, sketched in his famous lecture ‘Existentialism is a Humanism’ (1945). In Being and Nothingness Sartre emphasized the experience of freedom of choice, especially the project of choosing oneself, the defining pattern of one's past actions. Through vivid description of the ‘look’ of the Other, Sartre laid groundwork for the contemporary political significance of the concept of the Other (as in other groups or ethnicities). Indeed, in The Second Sex (1949) Simone de Beauvoir, Sartre's life-long companion, launched contemporary feminism with her nuance account of the perceived role of women as Other.

In 1940s Paris, Maurice Merleau-Ponty joined with Sartre and Beauvoir in developing phenomenology. In Phenomenology of Perception (1945) Merleau-Ponty developed a rich variety of phenomenology emphasizing the role of the body in human experience. Unlike Husserl, Heidegger, and Sartre, Merleau-Ponty looked to experimental psychology, analysing the reported experience of amputees who felt sensations in a phantom limb. Merleau-Ponty rejected both associationist psychology, focussed on correlations between sensation and stimulus, and intellectualist psychology, focussed on rational construction of the world in the mind. (Think of the behaviorist and computationalist models of mind in more recent decades of empirical psychology.) Instead, Merleau-Ponty focussed on the ‘body image,’ our experience of our own body and its significance in our activities. Extending Husserl's account of the lived body (as opposed to the physical body), Merleau-Ponty resisted the traditional Cartesian separation of mind and body. For the body image is neither in the mental realm nor in the mechanical-physical realm. Rather, my body is, as it were, me in my engaged action with things I perceive including other people.

The scope of Phenomenology of Perception is characteristic of the breadth of classical phenomenology, not least because Merleau-Ponty drew (with generosity) on Husserl, Heidegger, and Sartre while fashioning his own innovative vision of phenomenology. His phenomenology addressed the role of attention in the phenomenal field, the experience of the body, the spatiality of the body, the motility of the body, the body in sexual being and in speech, other selves, temporality, and the character of freedom so important in French existentialism. Near the end of a chapter on the Cogito (Descartes' ‘I think, therefore I am’), Merleau-Ponty succinctly captures his embodied, existential form of phenomenology, writing: Insofar as, when I reflect on the essence of subjectivity, I find it bound up with that of the body and that of the world, this is because my existence as subjectivity [= consciousness] is merely one with my existence as a body and with the existence of the world, and because the subject that I am, for when taken seriously, is inseparable from this body and this world. In short, consciousness is embodied (in the world), and equally body is infused with consciousness (with cognition of the world).

In the years since Hussserl, Heidegger, et al. wrote that phenomenologists have dug into all these classical issues, including intentionality, temporal awareness, intersubjectivity, practical intentionality, and the social and linguistic contexts of human activity. Interpretation of historical texts by Husserl et al. has played a prominent role in this work, both because the texts are rich and difficult and because the historical dimension is itself part of the practice of continental European philosophy. Since the 1960s, philosophers trained in the methods of analytic philosophy have also dug into the foundations of phenomenology, with an eye to 20th century work in philosophy of logic, language, and mind.

Phenomenology was already linked with logical and semantic theory in Husserl's Logical Investigations. Analytic phenomenology picks up on that connection. In particular, Dagfinn F¿llesdal and J. N. Mohanty have explored historical and conceptual relations between Husserl's phenomenology and Frége's logical semantics (in Frége's ‘On Sense and Reference,’ 1892). For Frége:: An expression refers to an object by way of a sense: Thus, two expressions (say, ‘the morning star’ and ‘the evening star’) may refer to the same object (Venus) but express different senses with different manners of presentation. For Husserl, similarly, an experience (or an act of consciousness) intends or refers to an object by way of a noema or noematic sense: Thus, two experiences may refer to the same object but have different noematic senses involving different ways of presenting the object (for example, in seeing the same object from different sides). Indeed, for Husserl, the theory of intentionality is a generalization of the theory of linguistic reference: as linguistic reference is mediated by sense, so intentional reference is mediated by noematic sense.

More recently, analytic philosophers of mind have rediscovered phenomenologically issues of mental representation, intentionality, consciousness, sensory experience, intentional content, and context-of-thought. Some of these analytic philosophers of mind hark back to William James and Franz Brentano at the origins of modern psychology, and some look to empirical research in today's cognitive neuroscience. Some researchers have begun to combine Phenomenological issues with issues of neuroscience and behavioural studies and mathematical modelling. Such studies will extend the methods of traditional phenomenology as the Zeitgeist moves on. We address philosophy of mind below.

The discipline of phenomenology forms one basic field in philosophy among others. How is phenomenology distinguished from, and related to, other fields in philosophy?

Traditionally, philosophy includes at least four core fields or disciplines: Ontology, epistemology, ethics, logic. Suppose phenomenology joins that list. Consider then these elementary definitions of field: (1) Ontology is the study of beings or their being - what is. (2) Epistemology is the study of knowledge - how we know. (3) Logic is the study of valid reasoning - how to reason. (4) Ethics is the study of right and wrong - how we should act. (5) Phenomenology is the study of our experience - how we experience. The domains of study in these five fields are clearly different, and they seem to call for different methods of study.

Philosophers have sometimes argued that one of these fields is ‘first philosophy,’ the most fundamental discipline, on which all philosophy or all knowledge or wisdom rests. Historically (it may be argued), Socrates and Plato put ethics first, then Aristotle put metaphysics or ontology first, then Descartes put epistemology first, then Russell put logic first, and then Husserl (in his later transcendental phase) put phenomenology first.

Consider epistemology. As we saw, phenomenology helps to define the phenomena on which knowledge claims rest, according to modern epistemology. On the other hand, phenomenology itself claims to achieve knowledge about the nature of consciousness, a distinctive description of the first-person knowledge. Through a form of intuition, consider logic, as a logical theory of meaning led Husserl into the theory of intentionality, the heart of phenomenology. On one account, phenomenology explicates the intentional or semantic force of ideal meanings, and propositional meanings are central to logical theory. But logical structure is expressed in language, either ordinary language or symbolic languages like those of predicate logic or mathematics or computer systems. It remains an important issue of debate where and whether language shapes specific forms of experience (thought, perception, emotion) and their content or meaning. So there is an important (if disputed) relation between phenomenology and logico-linguistic theory, especially philosophical logic and philosophy of language (as opposed to mathematical logic per se)

Consider ontology. Phenomenology studies (among other things) the nature of consciousness, which is a central issue in metaphysics or ontology, and one that leads into the traditional mind-body problem. Husserlian methodology would bracket the question of the existence of the surrounding world, thereby separating phenomenology from the ontology of the world. Yet Husserl's phenomenology presupposes theory about species and individuals (universals and particulars), relations of part and whole, and ideal meanings - all parts of ontology

Now consider ethics: Phenomenology might play a role in ethics by offering analyses of the structure of will, valuing, happiness, and care for others (in empathy and sympathy). Historically, though, ethics has been on the horizon of phenomenology. Husserl largely avoided ethics in his major works, though he featured the role of practical concerns in the structure of the life-world or of Geist (spirit, or culture, as in Zeitgeist). He once delivered a course of lectures giving ethics (like logic) a basic place in philosophy, indicating the importance of the phenomenology of sympathy in grounding ethics. In Being and Time Heidegger claimed not to pursue ethics while discussing phenomena ranging from care, conscience, and guilt to ‘falseness’ and ‘authenticity’ (all phenomena with theological echoes). In Being and Nothingness Sartre Analysed with subtlety the logical problem of ‘bad faith,’ yet he developed an ontology of value as produced by willing in good faith (which sounds like a revised Kantian foundation for morality). Beauvoir sketched an existentialist ethics, and Sartre left unpublished notebooks on ethics. However, an explicit Phenomenological approach to ethics emerged in the works of Emannuel Levinas, a Lithuanian phenomenologist who heard Husserl and Heidegger in Freiburg before moving to Paris. In Totality and Infinity (1961), modifying themes drawn from Husserl and Heidegger, Levinas focussed on the significance of the ‘face’ of the other, explicitly developing grounds for ethics in this range of phenomenology, writing an impressionistic style of prose with allusions to religious experience.

Allied with ethics are political and social philosophies. Sartre and Merleau-Ponty were politically engaged, in 1940s Paris and their existential philosophies (phenomenologically based) suggest a political theory based in individual freedom. Sartre later sought an explicit blend of existentialism with Marxism. Still, political theory has remained on the borders of phenomenology. Social theory, however, has been closer to phenomenology as such. Husserl Analysed the Phenomenological structure of the life-world and Geist generally, including our role in social activity. Heidegger stressed social practice, which he found more primordial than individual consciousness. Alfred Schutz developed a phenomenology of the social world. Sartre continued the Phenomenological appraisal of the meaning of the other, the fundamental social formation. Moving outward from Phenomenological issues, Michel Foucault studied the genesis and meaning of social institutions, from prisons to insane asylums. And Jacques Derrida has long practised a kind of phenomenology of language, pursuing sociologic meaning in the ‘deconstruction’ of wide-ranging texts. Aspects of French ‘poststructuralist’ theory are sometimes interpreted as broadly Phenomenological, but such issues are beyond the present purview.

Classical phenomenology, then, ties into certain areas of epistemology, logic, and ontology, and leads into parts of ethical, social, and political theory.

It ought to be obvious that phenomenology has a lot to say in the area called philosophy of mind. Yet the traditions of phenomenology and analytic philosophy of mind have not been closely joined, despite overlapping areas of interest. So it is appropriate to close this survey of phenomenology by addressing philosophy of mind, one of the most vigorously debated areas in recent philosophy.

The tradition of analytic philosophy began, early in the 20th century, with analyses of language, notably in the works of Gottlob Frége, Bertrand Russell, and Ludwig Wittgenstein. Then in The Concept of Mind (1949) Gilbert Ryle developed a series of analyses of language about different mental states, including sensation, belief, and will. Though Ryle is commonly deemed a philosopher of ordinary language, Ryle himself said The Concept of Mind could be called phenomenology. In effect, Ryle Analysed our Phenomenological understanding of mental states as reflected in ordinary language about the mind. From this linguistic phenomenology Ryle argued that Cartesian mind-body dualism involves a category mistake (the logic or grammar of mental verbs - ‘believe,’ ‘see,’ etc. - does not mean that we ascribe belief, sensation, etc., to ‘the ghost in the machine’). With Ryle's rejection of mind-body dualism, the mind-body problem was re-awakened: What is the ontology of mind/body, and how are mind and body related?

René Descartes, in his epoch-making Meditations on First Philosophy (1641), had argued that minds and bodies are two distinct kinds of being or substance with two distinct kinds of attributes or modes: Bodies are characterized by spatiotemporal physical properties, while minds are characterized by properties of thinking (including seeing, feeling, etc.). Centuries later, phenomenology would find, with Brentano and Husserl, that mental acts are characterized by consciousness and intentionality, while natural science would find that physical systems are characterized by mass and force, ultimately by gravitational, electromagnetic, and quantum fields. Where do we find consciousness and intentionality in the quantum-electromagnetic-gravitational field that, by hypothesis, orders everything in the natural world in which we humans and our minds exist? That is the mind-body problem today. In short, phenomenology by any other name lies at the heart of the contemporary, mind-body problem.

After Ryle, philosophers sought a more explicit and generally naturalistic ontology of mind. In the 1950s materialism was argued anew, urging that mental states are identical with states of the central nervous system. The classical identity theory holds that each token mental state (in a particular person's mind at a particular time) is identical with a token brain state (in that a person's brain at that time). The weaker of materialisms, holds instead, that each type of mental state is identical with a type of brain state. But materialism does not fit comfortably with phenomenology. For it is not obvious how conscious mental states as we experience them - sensations, thoughts, emotions - can simply be the complex neural states that somehow subserve or implement them. If mental states and neural states are simply identical, in token or in type, where in our scientific theory of mind does the phenomenology occur - is it not simply replaced by neuroscience? And yet experience is part of what is to be explained by neuroscience.

In the late 1960s and 1970s the computer model of mind set it, and functionalism became the dominant model of mind. On this model, mind is not what the brain consists in (electrochemical transactions in neurons in vast complexes). Instead, mind is what brains do: They are function of mediating between information coming into the organism and behaviour proceeding from the organism. Thus, a mental state is a functional state of the brain or of the human (or an animal) organism. More specifically, on a favourite variation of functionalism, the mind is a computing system: Mind is to brain as software is to hardware; Thoughts are just programs running on the brain's ‘NetWare.’ Since the 1970s the cognitive sciences - from experimental studies of cognition to neuroscience - have tended toward a mix of materialism and functionalism. Gradually, however, philosophers found that Phenomenological aspects of the mind pose problems for the functionalist paradigm too.

In the early 1970s Thomas Nagel argued in ‘What Is It Like to Be a Bat?’ (1974) that consciousness itself - especially the subjective character of what it is like to have a certain type of experience - escapes physical theory. Many philosophers pressed the case that sensory qualia - what it is like to feel pain, to see red, etc. - are not addressed or explained by a physical account of either brain structure or brain function. Consciousness has properties of its own. And yet, we know, it is closely tied to the brain. And, at some level of description, neural activities implement computation.

In the 1980s John Searle argued in Intentionality (1983) (and further in The Rediscovery of the Mind (1991)) that intentionality and consciousness are essential properties of mental states. For Searle, our brains produce mental states with properties of consciousness and intentionality, and this is all part of our biology, yet consciousness and intentionality require to ‘first-person’ ontology. Searle also argued that computers simulate but do not have mental states characterized by intentionality. As Searle argued, a computer system has a syntax (processing symbols of certain shapes) but has no semantics (the symbols lack meaning: we interpret the symbols). In this way Searle rejected both materialism and functionalism, while insisting that mind is a biological property of organisms like us: our brains ‘secrete’ consciousness

The analysis of consciousness and intentionality is central to phenomenology as appraised above, and Searle's theory of intentionality reads like a modernized version of Husserl's. (Contemporary logical theory takes the form of stating truth conditions for propositions, and Searle characterizes a mental state's intentionality by specifying its ‘satisfaction conditions’). However, there is an important difference in background theory. For Searle explicitly assumes the basic world view of natural science, holding that consciousness is part of nature. But Husserl explicitly brackets that assumption, and later phenomenologists - including Heidegger, Sartre, Merleau-Ponty - seem to seek a certain sanctuary for phenomenology beyond the natural sciences. And yet phenomenology itself should be largely neutral about further theories of how experience arises, notably from brain activity.

The philosophy or theory of mind overall may be factored into the following disciplines or ranges of theory relevant to mind: Phenomenology studies conscious experience as experienced, analysing the structure - the types, intentional forms and meanings, dynamics, and (certain) enabling conditions - of perception, thought, imagination, emotion, and volition and action.

Neuroscience studies the neural activities that serve as biological substrate to the various types of mental activity, including conscious experience. Neuroscience will be framed by evolutionary biology (explaining how neural phenomena evolved) and ultimately by basic physics (explaining how biological phenomena are grounded in physical phenomena). Here lie the intricacies of the natural sciences. Part of what the sciences are accountable for is the structure of experience, Analysed by phenomenology.

Cultural analysis studies the social practices that help to shape or serve as cultural substrate of the various types of mental activity, including conscious experience. Here we study the import of language and other social practices.

Ontology of mind studies the ontological type of mental activity in general, ranging from perception (which involves causal input from environment to experience) to volitional action (which involves causal output from volition to bodily movement).

This division of labour in the theory of mind can be seen as an extension of Brentano's original distinction between descriptive and genetic psychology. Phenomenology offers descriptive analyses of mental phenomena, while neuroscience (and wider biology and ultimately physics) offers models of explanation of what causes or gives rise to mental phenomena. Cultural theory offers analyses of social activities and their impact on experience, including ways language shapes our thought, emotion, and motivation. And ontology frames all these results within a basic scheme of the structure of the world, including our own minds.

Meanwhile, from an epistemological standpoint, all these ranges of theory about mind begin with how we observe and reason about and seek to explain phenomena we encounter in the world. And that is where phenomenology begins. Moreover, how we understand each piece of theory, including theory about mind, is central to the theory of intentionality, as it was, the semantics of thought and experience in general. And that is the heart of phenomenology.

The discipline of phenomenology may be defined as the study of structures of experience or consciousness. Literally. , Phenomenology is the

Study of ‘phenomena’: Appearances of things, or things as they appear in our experience, or the ways we experience things, thus the meaning’s things have in our experience. Phenomenology studies conscious experience as experienced from the subjective or first person point of view. This field of philosophy is then to be distinguished from, and related to, the other main fields of philosophy: ontology (the study of being or what is), epistemology (the study of knowledge), logic (the study of valid reasoning), ethics (the study of right and wrong action), etc.

The historical movement of phenomenology is the philosophical tradition launched in the first half of the 20th century by Edmund Husserl, Martin Heidegger, Maurice Merleau-Ponty, Jean-Paul Sartre. In that movement, the discipline of phenomenology was prized as the proper foundation of all philosophy - as opposed, say, to ethics or metaphysics or epistemology. The methods and characterization of the discipline were widely debated by Husserl and his successors, and these debates continue to the present day. (The definition of phenomenology offered above will thus is debatable, for example, by Heideggerians, but it remains the starting point in characterizing the discipline.)

In recent philosophy of mind, the term ‘phenomenology’ is often restricted to the characterization of sensory qualities of seeing, hearing, etc.: what it is like to have sensations of various kinds. However, our experience is normally much richer in content than mere sensation. Accordingly, in the Phenomenological tradition, phenomenology is given a much wider range, addressing the meaning things have in our experience, notably, the significance of objects, events, tools, the flow of time, the self, and others, as these things arise and are experienced in our ‘life-world.’

Phenomenology as a discipline has been central to the tradition of continental European philosophy throughout the 20th century, while philosophy of mind has evolved in the Austro-Anglo-American tradition of analytic philosophy that developed throughout the 20th century. Yet the fundamental character of our mental activity is pursued in overlapping ways within these two traditions. Accordingly, the perspective on phenomenology drawn in this article will accommodate both traditions. The main concern here will be to characterize the discipline of phenomenology, in contemporary views, while also highlighting the historical tradition that brought the discipline into its own.

Basically, phenomenology studies the structure of various types of experience ranging from perception, thought, memory, imagination, emotion, desire, and volition to bodily awareness, embodied action, and social activity, including linguistic activity. The structure of these forms of experience typically involves what Husserl called ‘intentionality,’ that is, the directedness of experience toward things in the world, the property of consciousness that it is a consciousness of or about something. According to classical Husserlian phenomenology, our experience remains directed towardly and represented or ‘intends’ - things only through particular concepts, thoughts, ideas, images, etc. These make up the meaning or content of a given experience, and are distinct from the things they present or mean.

The basic intentional structure of consciousness, we find in reflection or analysis, involves further forms of experience. Thus, phenomenology develops a complex account of temporal awareness (within the stream of consciousness), spatial awareness (notably in perception), attention (distinguishing focal and marginal or ‘horizonal’ awareness), awareness of one's own experience (self-consciousness, in one sense), self-awareness (awareness-of-oneself), the self in different roles (as thinking, acting, etc.), embodied action (including kinesthetic awareness of one's movement), purposive intention for its desire for action (more or less explicit), awareness of other persons (in empathy, intersubjectivity, collectivity), linguistic activity (involving meaning, communication, understanding others), social interaction (including collective action), and everyday activity in our surrounding life-world (in a particular culture).

Furthermore, in a different dimension, we find various grounds or enabling conditions -conditions of the possibility - of intentionality, including embodiment, bodily skills, cultural context, language and other social practices, social background, and contextual aspects of intentional activities. Thus, phenomenology leads from conscious experience into conditions that help to give experience its intentionality. Traditional phenomenology has focussed on subjective, practical, and social conditions of experience. Recent philosophy of mind, however, has focussed especially on the neural substrate of experience, on how conscious experience and mental representation or intentionality is grounded in brain activity. It remains a difficult question how much of these grounds of experience fall within the province of phenomenology as a discipline. Cultural conditions thus seem closer to our experience and to our familiar self-understanding than do the electrochemical workings of our brain, much less our dependence on quantum-mechanical states of physical systems to which we may belong. The cautious thing to say is that phenomenology leads in some ways into at least some background conditions of our experience.

Phenomenology studies structures of conscious experience as experienced from the first-person point of view, along with relevant conditions of experience. The central structure of an experience is its intentionality, the way it is directed through its content or meaning toward a certain object in the world.

We all experience various types of experience including perception, imagination, thought, emotion, desire, volition, and action. Thus, the domain of phenomenology is the range of experiences including these types (among others). Experience includes not only relatively passive experience as in vision or hearing, but also active experience as in walking or hammering a nail or kicking a ball. (The range will be specific to each species of being that enjoys consciousness; Our focus is on our own, human, experience. Not all conscious beings will, or will be able to, practice phenomenology, as we do.)

Conscious experiences have a unique feature: We experience them, we live through them or perform them. Other things in the world we may observe and engage. But we do not experience them, in the sense of living through or performing them. This experiential or first-person feature - that of being experienced -is an essential part of the nature or structure of conscious experience: as we say, ‘I see / think / desire / do . . .’ This feature is both a Phenomenological and an ontological feature of each experience: it is part of what it is for the experience to be experienced (Phenomenological) and part of what it is for the experience to be (ontological).

How shall we study conscious experience? We reflect on various types of experiences just as we experience them. That is to say, we proceed from the first-person point of view. However, we do not normally characterize an experience at the time we are performing it. In many cases we do not have that capability: a state of intense anger or fear, for example, consumes the entire focus at the time. Rather, we acquire a background of having lived through a given type of experience, and we look to our familiarity with that type of experience: While hearing a song, seeing the sun set, thinking about love, intending to jump a hurdle. The practice of phenomenology assumes such familiarity with the type of experiences to be characterized. Importantly, it is atypical of experience that phenomenology pursues, rather than a particular fleeting experience - unless its type is what interests us.

Classical phenomenologists practised some three distinguishable methods. (1) We describe a type of experience just as we find it in our own (past) experience. Thus, Husserl and Merleau-Ponty spoke of pure description of lived experience. (2) We interpret a type of experience by relating it to relevant features of context. In this vein, Heidegger and his followers spoke of hermeneutics, the art of interpretation in context, especially social and linguistic context. (3) We analyse the form of a type of experience. In the end, all the classical phenomenologists practised analysis of experience, factoring out notable features for further elaboration.

These traditional methods have been ramified in recent decades, expanding the methods available to phenomenology. Thus: (4) In a logico-semantic model of phenomenology, we specify the truth conditions for a type of thinking (say, where I think that dogs chase cats) or the satisfaction conditions for a type of intention (say, where I intend or will to jump that hurdle). (5) In the experimental paradigm of cognitive neuroscience, we design empirical experiments that tend to confirm or refute aspects of experience (say, where a brain scan shows electrochemical activity in a specific region of the brain thought to subserve a type of vision or emotion or motor control). This style of ‘neurophenomenology’ assumes that conscious experience is grounded in neural activity in embodied action in appropriate surroundings - mixing pure phenomenology with biological and physical science in a way that was not wholly congenial to traditional phenomenologists.

What makes an experience conscious is a certain awareness one has of the experience while living through or performing it. This form of inner awareness has been a topic of considerable debate, centuries after the issue arose with Locke's notion of self-consciousness on the heels of Descartes' sense of consciousness (conscience, co-knowledge). Does this awareness-of-experience consist in a kind of inner observation of the experience, as if one were doing two things at once? (Brentano argued no.) Is it a higher-order perception of one's mind's operation, or is it a higher-order thought about one's mental activity? (Recent theorists have proposed both.) Or is it a different form of inherent structure? (Sartre took this line, drawing on Brentano and Husserl.) These issues are beyond the scope of this article, but notice that these results of Phenomenological analysis shape the characterization of the domain of study and the methodology appropriate to the domain. For awareness-of-experience is a defining trait of conscious experience, the trait that gives experience a first-person, lived character. It is that a living characterization resembling its self, that life is to offer the experience through which allows a first-person perspective on the object of study, namely, experience, and that perspective is characteristic of the methodology of phenomenology.

Conscious experience is the starting point of phenomenology, but experience shades off into fewer overtly conscious phenomena. As Husserl and others stressed, we are only vaguely aware of things in the margin or periphery of attention, and we are only implicitly aware of the wider horizon of things in the world around us. Moreover, as Heidegger stressed, in practical activities like walking along, or hammering a nail, or speaking our native tongue, we are not explicitly conscious of our habitual patterns of action. Furthermore, as psychoanalysts have stressed, much of our intentional mental activity is not conscious at all, but may become conscious in the process of therapy or interrogation, as we come to realize how we feel or think about something. We should allow, then, that the domain of phenomenology - our own experience - spreads out from conscious experience into semiconscious and even unconscious mental activity, along with relevant background conditions implicitly invoked in our experience. (These issues are subject to debate; the point here is to open the door to the question of where to draw the boundary of the domain of phenomenology.)

To begin an elementary exercise in phenomenology, consider some typical experiences one might have in everyday life, characterized in the first person: (1) ‘I’ witnesses that fishing boat off the coast as dusk descends over the Pacific. (2) I hear that helicopter whirring overhead as it approaches the hospital. (3) I am thinking that phenomenology differs from psychology. (4) I wish that warm rain from Mexico were falling like last week. (5) I imagine a fearsome creature like that in my nightmare. (6) I intend to finish my writing by noon. (7) I walk carefully around the broken glass on the sidewalk. (8) I stroke a backhand cross-court with that certain underspin. (9) I am searching for the words to make my point in conversation.

Here are rudimentary characterizations of some familiar types of experience. Each sentence is a simple form of Phenomenological description, articulating in everyday English the structure of the type of experience so described. The subject term of ‘I,’ indicates the first-person structure of the experience: The intentionality proceeds from the subject. As the verb indicates, the type of intentional activity so described, as perception, thought, imagination, etc. Of central importance is the way that objects of awareness are presented or intended in our experiences, especially, the way we see or conceive or think about objects. The direct-object expression (‘that fishing boat off the coast’) articulates the mode of presentation of the object in the experience: The content or meaning of the experience, the core of what Husser called noema. In effect, the object-phrase expresses the noema of the act described, that is, to the extent that language has appropriate expressive power. The overall form of the given sentence articulates of a basic form of intentionality, in that of an experience has to its own subject-act-content-object.

Fruitful Phenomenological description or interpretation, as in Husserl or Merleau-Ponty, will far outrun such simple Phenomenological descriptions as above. But such simple descriptions bring out the basic form of intentionality. As we interpret the Phenomenological description further, we may assess the relevance of the context of experience. And we may turn to wider conditions of the possibility of that type of experience. In this way, in the practice of phenomenology, we classify, describe, interpret, and analyse structures of experiences in ways that answer to our own experience.

In such interpretive-descriptive analyses of experience, we immediately observe that we are analysing familiar forms of consciousness, conscious experience of or about this or that. Intentionality is thus the salient structure of our experience, and much of the phenomenology proceeds as the study of different aspects of intentionality. Thus, we explore structures of the stream of consciousness, the enduring self, the embodied self, and bodily action. Furthermore, as we reflect on how these phenomena work, we turn to the analysis of relevant conditions that enable our experiences to occur as they do, and to represent or intend as they do. Phenomenology then leads into analyses of conditions of the possibility of intentionality, conditions involving motor skills and habits, backgrounding to social practices, and often language, with its special place in human affairs. The Oxford English Dictionary presents the following definition: ‘Phenomenology. (i) The science of phenomena as distinct from being (ontology). (ii) That division of any science that describes and classifies its phenomena. From the Greek phainomenon, appearance.’ In philosophy, the term is used in the first sense, amid debates of theory and methodology. In physics and philosophy of science, the term is used in the second sense, even if only occasionally.

In its root meaning, then, phenomenology is the study of phenomena: Literally, appearances as opposed to reality. This ancient distinction launched philosophy as we emerged from Plato's cave. Yet the discipline of phenomenology did not blossom until the 20th century and remains poorly understood in many circles of contemporary philosophy. What is that discipline? How did philosophy move from a root concept of phenomena to the discipline of phenomenology?

Originally, in the 18th century, ‘phenomenology’ meant the theory of appearances fundamental to empirical knowledge, especially sensory appearances. The term seems to have been introduced by Johann Heinrich Lambert, a follower of Christian Wolff. Subsequently, Immanuel Kant used the term occasionally in various writings, as did Johann Gottlieb Fichte and G. W. F. Hegel. By 1889 Franz Brentano used the term to characterize what he called ‘descriptive psychology.’ From there Edmund Husserl took up the term for his new science of consciousness, and the rest is history.

Suppose we say phenomenology study’s phenomena: what appears to us - and its appearing? How shall we understand phenomena? The term has a rich history in recent centuries, in which we can see traces of the emerging discipline of phenomenology.

In a strict empiricist vein, what appears before the mind are sensory data or qualia: either patterns of one's own sensations (seeing red here now, feeling this ticklish feeling, hearing that resonant bass tone) or sensible patterns of worldly things, say, the looks and smells of flowers (what John Locke called secondary qualities of things). In a strict rationalist vein, by contrast, what appears before the mind are ideas, rationally formed ‘clear and distinct ideas’ (in René Descartes' ideal). In Immanuel Kant's theory of knowledge, fusing rationalist and empiricist aims, what appears to the mind are phenomena defined as things-as-they-appear or things-as-they-are-represented (in a synthesis of sensory and conceptual forms of objects-as-known). In Auguste Comte's theory of science, phenomena (phenomenes) are the facts (faits, what occurs) that a given science would explain.

In 18th and 19th century epistemology, then, phenomena are the starting points in building knowledge, especially science. Accordingly, in a familiar and still current sense, phenomena are whatever we observe (perceive) and seek to explain.

As the discipline of psychology emerged late in the 19th century, however, phenomena took on a somewhat different guise. In Franz Brentano's Psychology from an Empirical Standpoint (1874), phenomena are of what occurs in the mind: Mental phenomena are acts of consciousness (or their contents), and physical phenomena are objects of external perception starting with colours and shapes. For Brentano, physical phenomena exist ‘intentionally’ in acts of consciousness. This view revives a Medieval notion Brentano called ‘intentional in-existence. However, the ontology remains undeveloped (what is it to exist in the mind, and do physical objects exist only in the mind?). Moreover, phenomenons are whatever we are conscious of, as a phenomenon might that its events lay succumbantly around us, other people, ourselves. Even (in reflection) our own conscious experiences, as we experience these. In a certain technical sense, phenomena are things as they are given to our consciousness, whether in perception or imagination or thought or volition. This conception of phenomena would soon inform the new discipline of phenomenology.

Brentano distinguished descriptive psychology from genetic psychology. Where genetic psychology seeks the causes of various types of mental phenomena, descriptive psychology defines and classifies the various types of mental phenomena, including perception, judgment, emotion, etc. According to Brentano, every mental phenomenon, or act of consciousness, is directed toward some object, and only mental phenomena are so directed. This thesis of intentional directedness was the hallmark of Brentano's descriptive psychology. In 1889 Brentano used the term ‘phenomenology’ for descriptive psychology, and the way was paved for Husserl's new science of phenomenology.

Phenomenology as we know it was launched by Edmund Husserl in his Logical Investigations (1900-01). Two importantly different lines of theory came together in that monumental work: Psychological theory, on the heels of Franz Brentano (and William James, whose Principles of Psychology appeared in 1891 and greatly impressed Husserl); And logical or semantic theory, on the heels of Bernard Bolzano and Hussserl's contemporaries who founded modern logic, including Gottlob Frége.

Hussserl's Logical Investigations was inspired by Bolzano's ideal of logic, while taking up Brentano's conception of descriptive psychology. In his Theory of Science (1835) Bolzano distinguished between subjective and objective ideas or representations (Vorstellungen). In effect Bolzano criticized Kant and before him the classical empiricists and rationalists for failing to make this sort of distinction, thereby rendering phenomena merely subjective. Logic studies objective ideas, including propositions, which in turn make up objective theories as in the sciences. Psychology would, by contrast, study subjective ideas, the concrete contents (occurrences) of mental activities in particular minds at a given time. Husserl was after both, within a single discipline. So phenomena must be reconceived as objective intentional contents (sometimes called intentional objects) of subjective acts of consciousness. Phenomenology would then study this complex of consciousness and correlated phenomena. In Ideas I (Book One, 1913) Husserl introduced two Greek words to capture his version of the Bolzanoan distinction: noesis and noema (from the Greek verb noéaw, meaning to perceive, think, intend, from what place the noun nous or mind). The intentional process of consciousness is called noesis, while its ideal content is called noema. The noema of an act of consciousness Husserl characterized both as an ideal meaning and as ‘the object as intended.’ Thus the phenomenon, or object-as-it-appears, becomes the noema, or object-as-it-is-intended. The interpretations of Husserl's theory of noema have been several and amount to different developments of Husserl's basic theory of intentionality. (Is the noema an aspect of the object intended, or rather a medium of intention?)

For Husserl, then, phenomenology integrates a kind of psychology with a kind of logic. It develops a descriptive or analytic psychology in that it describes and analytical divisions of subjective mental activity or experience, in short, acts of consciousness. Yet it develops a kind of logic - a theory of meaning (today we say logical semantics) -by that, it describes and approves to analytical justification that an objective content of consciousness, brings forthwith the ideas, concepts, images, propositions, in short, ideal meanings of various types that serve as intentional contents, or noematic meanings, of various types of experience. These contents are shareable by different acts of consciousness, and in that sense they are objective, ideal meanings. Following Bolzano (and to some extent the platonistic logician Hermann Lotze), Husserl opposed any reduction of logic or mathematics or science to mere psychology, to how human beings happen to think, and in the same spirit he distinguished phenomenology from mere psychology. For Husserl, phenomenology would study consciousness without reducing the objective and shareable meanings that inhabit experience to merely subjective happenstances. Ideal meaning would be the engine of intentionality in acts of consciousness.

A clear conception of phenomenology awaited Husserl's development of a clear model of intentionality. Indeed, phenomenology and the modern concept of intentionality emerged hand-in-hand in Husserl's Logical Investigations (1900-01). With theoretical foundations laid in the Investigations, Husserl would then promote the radical new science of phenomenology in Ideas. And alternative visions of phenomenology would soon follow.

Phenomenology came into its own with Husserl, much as epistemology came into its own with Descartes, and ontology or metaphysics came into its own with Aristotle on the heels of Plato. Yet phenomenology has been practised, with or without the name, for many centuries. When Hindu and Buddhist philosophers reflected on states of consciousness achieved in a variety of meditative states, they were practising phenomenology. When Descartes, Hume, and Kant characterized states of perception, thought, and imagination, they were practising phenomenology. When Brentano classified varieties of mental phenomena (defined by the directedness of consciousness), he was practising phenomenology. When William James appraised kinds of mental activity in the stream of consciousness (including their embodiment and their dependence on habit), he too was practising phenomenology. And when recent analytic philosophers of mind have addressed issues of consciousness and intentionality, they have often been practising phenomenology. Still, the discipline of phenomenology, its roots tracing back through the centuries, came full to flower in Husserl.

Husserl's work was followed by a flurry of Phenomenological writing in the first half of the 20th century. The diversity of traditional phenomenology is apparent in the Encyclopaedia of Phenomenology (Kluwer Academic Publishers, 1997, Dordrecht and Boston), which features separate articles on some seven types of phenomenology. (1) Transcendental constitutive phenomenology studies how objects are constituted in pure or transcendental consciousness, setting aside questions of any relation to the natural world around us. (2) Naturalistic constitutive phenomenology studies how consciousness constitutes or takes things in the world of nature, assuming with the natural attitude that consciousness is part of nature. (3) Existential phenomenology studies concrete human existence, including our experience of free choice or action in concrete situations. (4) Generative historicist phenomenology studies how meaning, as found in our experience, is generated in historical processes of collective experience over time. (5) Genetic phenomenology studies the genesis of meanings of things within one's own stream of experience. (6) Hermeneutical phenomenology studies interpretive structures of experience, how we understand and engage things around us in our human world, including ourselves and others. (7) Realistic phenomenology studies the structure of consciousness and intentionality, assuming it occurs in a real world that is largely external to consciousness and not somehow brought into being by consciousness.

The most famous of the classical phenomenologists were Husserl, Heidegger, Sartre, and Merleau-Ponty. In these four thinkers we find different conceptions of phenomenology, different methods, and different results. A brief sketch of their differences will capture both a crucial period in the history of phenomenology and a sense of the diversity of the field of phenomenology.

In his Logical Investigations (1900-01) Husserl outlined a complex system of philosophy, moving from logic to philosophy of language, to ontology (theory of universals and parts of wholes), to a Phenomenological theory of intentionality, and finally to a Phenomenological theory of knowledge. Then in Ideas I (1913) he focussed squarely on phenomenology itself. Husserl defined phenomenology as ‘the science of the essence of consciousness,’ entered on the defining trait of intentionality, approached explicitly ‘in the first person.’ In this spirit, we may say phenomenology is the study of consciousness - that is, conscious experience of various types - as experienced from the first-person point of view. In this discipline we study different forms of experience just as we experience them, from the perspective of its topic for living through or performing them. Thus, we characterize experiences of seeing, hearing, imagining, thinking, feeling (i.e., emotion), wishing, desiring, willing, and acting, that is, embodied volitional activities of walking, talking, cooking, carpentering, etc. However, not just any characterization of an experience will do. Phenomenological analysis of a given type of experience will feature the ways in which we ourselves would experience that form of conscious activity. And the leading property of our familiar types of experience is their intentionality, their being a consciousness of or about something, something experienced or presented or engaged in a certain way. How I see or conceptualize or understand the object I am dealing with defines the meaning of that object in my current experience. Thus, phenomenology features a study of meaning, in a wide sense that includes more than what is expressed in language.

In Ideas, Husserl presented phenomenology with a transcendental turn. In part this means that Husserl took on the Kantian idiom of ‘transcendental idealism,’ looking for conditions of the possibility of knowledge, or of consciousness generally, and arguably turning away from any reality beyond phenomena. But Hussserl's transcendental, and turns to involve his discovery of the method of epoché (from the Greek skeptics' notion of abstaining from belief). We are to practice phenomenology, Husserl proposed, by ‘bracketing’ the question of the existence of the natural world around us. We thereby turn our attention, in reflection, to the structure of our own conscious experience. Our first key result is the observation that each act of consciousness is a consciousness of something, that is, intentional, or directed toward something. Consider my visual experience wherein I see a tree across the square. In Phenomenological reflection, we need not concern ourselves with whether the tree exists: my experience is of a tree whether or not such a tree exists. However, we do need to concern ourselves with how the object is meant or intended. I see a Eucalyptus tree, not a Yucca tree; I see the object as a Eucalyptus tree, with a certain shape, with bark stripping off, etc. Thus, bracketing the tree itself, we turn our attention to my experience of the tree, and specifically to the content or meaning in my experience. This tree-as-perceived Husserl calls the noema or noematic sense of the experience.

Philosophers succeeding Husserl debated the proper characterization of phenomenology, arguing over its results and its methods. Adolf Reinach, an early student of Husserl's (who died in World War I), argued that phenomenology should remain cooperatively affiliated within there be of the view that finds to some associative values among the finer qualities that have to them the realist’s ontology, as in Husserl's Logical Investigations. Roman Ingarden, a Polish phenomenologist of the next generation, continued the resistance to Hussserl's turn to transcendental idealism. For such philosophers, phenomenology should not bracket questions of being or ontology, as the method of epoché would suggest. And they were not alone. Martin Heidegger studied Hussserl's early writings, worked as Assistant to Husserl in 1916, and in 1928 Husserl was to succeed in the prestigious chair at the University of Freiburg. Heidegger had his own ideas about phenomenology.

In Being and Time (1927) Heidegger unfurled his rendition of phenomenology. For Heidegger, we and our activities are always ‘in the world,’ our being is being-in-the-world, so we do not study our activities by bracketing the world, rather we interpret our activities and the meaning things have for us by looking to our contextual relations to things in the world. Indeed, for Heidegger, phenomenology resolves into what he called ‘fundamental ontology.’ We must distinguish beings from their being, and we begin our investigation of the meaning of being in our own case, examining our own existence in the activity of ‘Dasein’ (that being whose being is in each case my own). Heidegger resisted Husserl's neo-Cartesian emphasis on consciousness and subjectivity, including how perception presents things around us. By contrast, Heidegger held that our more basic ways of relating to things are in practical activities like hammering, where the phenomenology reveals our situation in a context of equipment and in being-with-others

In Being and Time Heidegger approached phenomenology, in a quasi-poetic idiom, through the root meanings of ‘logos’ and ‘phenomena,’ so that phenomenology is defined as the art or practice of ‘letting things show themselves.’ In Heidegger's inimitable linguistic play on the Greek roots, ‘phenomenology’ means, . . . to let that which shows itself be seen from themselves in the very way in which it shows itself from itself. Here Heidegger explicitly parodies Hussserl's call, ‘To the things themselves!’, or ‘To the phenomena themselves!’ Heidegger went on to emphasize practical forms of comportment or better relating (Verhalten) as in hammering a nail, as opposed to representational forms of intentionality as in seeing or thinking about a hammer. Being and Time developed an existential interpretation of our modes of being including, famously, our being-toward-death.

In a very different style, in clear analytical prose, in the text of a lecture course called The Basic Problems of Phenomenology (1927), Heidegger traced the question of the meaning of being from Aristotle through many other thinkers into the issues of phenomenology. Our understanding of beings and their being comes ultimately through phenomenology. Here the connection with classical issues of ontology is more apparent, and consonant with Hussserl's vision in the Logical Investigations (an early source of inspiration for Heidegger). One of Heidegger's most innovative ideas was his conception of the ‘ground’ of being, looking to modes of being more fundamental than the things around us (from trees to hammers). Heidegger questioned the contemporary concern with technology, and his writing might suggest that our scientific theories are historical artifacts that we use in technological practice, rather than systems of ideal truth (as Husserl had held). Our deep understanding of being, in our own case, comes rather from phenomenology, Heidegger held.

In the 1930s phenomenology migrated from Austrian and then German philosophy into French philosophy. The way had been paved in Marcel Proust's In Search of Lost Time, in which the narrator recounts in close detail his vivid recollections of experiences, including his famous associations with the smell of freshly baked madeleines. This sensibility to experience traces to Descartes' work, and French phenomenology has been an effort to preserve the central thrust of Descartes' insights while rejecting mind-body dualism. The experience of one's own body, or one's lived or living body, has been an important motif in many French philosophers of the 20th century.

In the novel Nausea (1936) Jean-Paul Sartre described a bizarre course of experience in which the protagonist, writing in the first person, describes how ordinary objects lose their meaning until he encounters pure being at the foot of a chestnut tree, and in that moment recovers his sense of his own freedom. In Being and Nothingness (1943, written partly while a prisoner of war), Sartre developed his conception of Phenomenological ontology. Consciousness is a consciousness of objects, as Husserl had stressed. In Sartre's model of intentionality, the central player in consciousness is a phenomenon, and the occurrence of a phenomenon is just a consciousness-of-an-object. The chestnut tree I see is, for Sartre, such a phenomenon in my consciousness. Indeed, all things in the world, as we normally experience them, are phenomena, beneath or behind which lies their ‘being-in-itself.’ Consciousness, by contrast, has ‘being-for-itself,’ since everything conscious is not only a consciousness-of-its-object but also a pre-reflective consciousness-of-itself (conscience). Yet for Sartre, unlike Husserl, the formal ‘I’ or self is nothing but a sequence of acts of consciousness, notably including radically free choices (like a Humean bundle of perceptions).

For Sartre, the practice of phenomenology proceeds by a deliberate reflection on the structure of consciousness. Sartre's method is in effect a literary style of interpretive description of different types of experience in relevant situations - a practice that does not really fit the methodological proposals of either Husserl or Heidegger, but makes benefit from Sartre's great literary skill. (Sartre wrote many plays and novels and was awarded the Nobel Prize in Literature.)

Sartre's phenomenology in Being and Nothingness became the philosophical foundation for his popular philosophy of existentialism, sketched in his famous lecture ‘Existentialism is a Humanism’ (1945). In Being and Nothingness Sartre emphasized the experience of freedom of choice, especially the project of choosing oneself, the defining pattern of one's past actions. Through vivid description of the ‘look’ of the Other, Sartre laid groundwork for the contemporary political significance of the concept of the Other (as in other groups or ethnicities). Indeed, in The Second Sex (1949) Simone de Beauvoir, Sartre's life-long companion, launched contemporary feminism with her nuance account of the perceived role of women as Other.

In 1940s Paris, Maurice Merleau-Ponty joined with Sartre and Beauvoir in developing phenomenology. In Phenomenology of Perception (1945) Merleau-Ponty developed a rich variety of phenomenology emphasizing the role of the body in human experience. Unlike Husserl, Heidegger, and Sartre, Merleau-Ponty looked to experimental psychology, analysing the reported experience of amputees who felt sensations in a phantom limb. Merleau-Ponty rejected both associationist psychology, focussed on correlations between sensation and stimulus, and intellectualist psychology, focussed on rational construction of the world in the mind. (Think of the behaviorist and computationalist models of mind in more recent decades of empirical psychology.) Instead, Merleau-Ponty focussed on the ‘body image,’ our experience of our own body and its significance in our activities. Extending Hussserl's account of the lived body (as opposed to the physical body), Merleau-Ponty resisted the traditional Cartesian separation of mind and body. For the body image is neither in the mental realm nor in the mechanical-physical realm. Rather, my body is, as it were, me in my engaged action with things I perceive including other people.

The scope of Phenomenology of Perception is characteristic of the breadth of classical phenomenology, not least because Merleau-Ponty drew (with generosity) on Husserl, Heidegger, and Sartre while fashioning his own innovative vision of phenomenology. His phenomenology addressed the role of attention in the phenomenal field, the experience of the body, the spatiality of the body, the motility of the body, the body in sexual being and in speech, other selves, temporality, and the character of freedom so important in French existentialism. Near the end of a chapter on the Cogito (Descartes' ‘I think, therefore I am’), Merleau-Ponty succinctly captures his embodied, existential form of phenomenology, writing: Insofar as, when I reflect on the essence of subjectivity, I find it bound up with that of the body and that of the world, this is because my existence as subjectivity [= consciousness] is merely one with my existence as a body and with the existence of the world, and because the subject that I am, when appropriated concrete, it is inseparable from this body and this world.

In short, consciousness is embodied (in the world), and equally body is infused with consciousness (with cognition of the world).

In the years since Hussserl, Heidegger, et al, wrote that its topic or ways of conventional study are to phenomenologists of having in accord dug into all these classical disseminations that include, intentionality, temporal awareness, intersubjectivity, practical intentionality, and the social and linguistic contexts of human activity. Interpretation of historical texts by Husserl et al. has played a prominent role in this work, both because the texts are rich and difficult and because the historical dimension is itself part of the practice of continental European philosophy. Since the 1960s, philosophers trained in the methods of analytic philosophy have also dug into the foundations of phenomenology, with an eye to 20th century work in philosophy of logic, language, and mind.

Phenomenology was already linked with logical and semantic theory in Husserl's Logical Investigations. Analytic phenomenology picks up on that connection. In particular, Dagfinn F¿llesdal and J. N. Mohanty have explored historical and conceptual relations between Husserl's phenomenology and Frége's logical semantics (in Frége's ‘On Sense and Reference,’ 1892). For Frége, an expression refers to an object by way of a sense: Thus, two expressions (say, ‘the morning star’ and ‘the evening star’) may refer to the same object (Venus) but express different senses with different manners of presentation. For Husserl, similarly, an experience (or act of consciousness) intends or refers to an object by way of a noema or noematic sense: Consequently, two experiences may refer to the same object but have different noematic senses involving different ways of presenting the object (for example, in seeing the same object from different sides). Indeed, for Husserl, the theory of intentionality is a generalization of the theory of linguistic reference: as linguistic reference is mediated by sense, so intentional reference is mediated by noematic sense.

More recently, analytic philosophers of mind have rediscovered Phenomenological issues of mental representation, intentionality, consciousness, sensory experience, intentional content, and context-of-thought. Some of these analytic philosophers of mind hark back to William James and Franz Brentano at the origins of modern psychology, and some look to empirical research in today's cognitive neuroscience. Some researchers have begun to combine Phenomenological issues with issues of neuroscience and behavioural studies and mathematical modelling. Such studies will extend the methods of traditional phenomenology as the Zeitgeist moves on.

The discipline of phenomenology forms one basic field in philosophy among others. How is phenomenology distinguished from, and related to, other fields in philosophy?

Traditionally, philosophy includes at least four core fields or disciplines: Ontology, epistemology, ethics, logic presupposes phenomenology as it joins that list. Consider then these elementary definitions of field: (1) Ontology is the study of beings or their being - what is. (2) Epistemology is the study of knowledge - how we know. (3) Logic is the study of valid reasoning - how to reason. (4) Ethics is the study of right and wrong - how we should act. (5) Phenomenology is the study of our experience - how we experience.

The domains of study in these five fields are clearly different, and they seem to call for different methods of study.

Philosophers have sometimes argued that one of these fields is ‘first philosophy,’ the most fundamental discipline, on which all philosophy or all knowledge or wisdom rests. Historically (it may be argued), Socrates and Plato put ethics first, then Aristotle put metaphysics or ontology first, then Descartes put epistemology first, then Russell put logic first, and then Husserl (in his later transcendental phase) put phenomenology first.

Consider epistemology. As we saw, phenomenology helps to define the phenomena on which knowledge claims rest, according to modern epistemology. On the other hand, phenomenology itself claims to achieve knowledge about the nature of consciousness, a distinctive description of first-person knowledge, through a form of intuition.

Consider logic saw being a logical theory of meaning, in that this had persuaded Husserl into the theory of intentionality, the heart of phenomenology. On one account, phenomenology explicates the intentional or semantic force of ideal meanings, and propositional meanings are central to logical theory. But logical structure is expressed in language, either ordinary language or symbolic languages like those of predicate logic or mathematics or computer systems. It remains an important issue of debate where and whether language shapes specific forms of experience (thought, perception, emotion) and their content or meaning. So there is an important (if disputed) relation between phenomenology and logico-linguistic theory, especially philosophical logic and philosophy of language (as opposed to mathematical logic per se).

Consider ontology. Phenomenology studies (among other things) the nature of consciousness, which is a central issue in metaphysics or ontology, and one that lead into the traditional mind-body problem. Husserlian methodology would bracket the question of the existence of the surrounding world, thereby separating phenomenology from the ontology of the world. Yet Husserl's phenomenology presupposes theory about species and individuals (universals and particulars), relations of part and whole, and ideal meanings - all parts of ontology.

Now consider ethics. Phenomenology might play a role in ethics by offering analyses of the structure of will, valuing, happiness, and care for others (in empathy and sympathy). Historically, though, ethics has been on the horizon of phenomenology. Husserl largely avoided ethics in his major works, though he featured the role of practical concerns in the structure of the life-world or of Geist (spirit, or culture, as in Zeitgeist). He once delivered a course of lectures giving ethics (like logic) a basic place in philosophy, indicating the importance of the phenomenology of sympathy in grounding ethics. In Being and Time Heidegger claimed not to pursue ethics while discussing phenomena ranging from care, conscience, and guilt to ‘falseness’ and ‘authenticity’ (all phenomena with theological echoes). In Being and Nothingness Sartre analysed with subtlety the logical problem of ‘bad faith,’ yet he developed an ontology of value as produced by willing in good faith (which sounds like a revised Kantian foundation for morality). Beauvoir sketched an existentialist ethics, and Sartre left unpublished notebooks on ethics. However, an explicit Phenomenological approach to ethics emerged in the works of Emannuel Levinas, a Lithuanian phenomenologist who heard Husserl and Heidegger in Freiburg before moving to Paris. In Totality and Infinity (1961), modifying themes drawn from Husserl and Heidegger, Levinas focussed on the significance of the ‘face’ of the other, explicitly developing grounds for ethics in this range of phenomenology, writing an impressionistic style of prose with allusions to religious experience.

Allied with ethics that on the same line, signify political and social philosophy. Sartre and Merleau-Ponty were politically captivated in 1940s Paris, and their existential philosophies (phenomenologically based) suggest a political theory based in individual freedom. Sartre later sought an explicit blend of existentialism with Marxism. Still, political theory has remained on the borders of phenomenology. Social theory, however, has been closer to phenomenology as such. Husserl analysed the Phenomenological structure of the life-world and Geist generally, including our role in social activity. Heidegger stressed social practice, which he found more primordial than individual consciousness. Alfred Schutz developed a phenomenology of the social world. Sartre continued the Phenomenological appraisal of the meaning of the other, the fundamental social formation. Moving outward from Phenomenological issues, Michel Foucault studied the genesis and meaning of social institutions, from prisons to insane asylums. And Jacques Derrida has long practised a kind of phenomenology of language, seeking socially meaning in the ‘deconstruction’ of wide-ranging texts. Aspects of French ‘poststructuralist’ theory are sometimes interpreted as broadly Phenomenological, but such issues are beyond the present purview.

Classical phenomenology, then, ties into certain areas of epistemology, logic, and ontology, and leads into parts of ethical, social, and political theory.

It ought to be obvious that phenomenology has a lot to say in the area called philosophy of mind. Yet the traditions of phenomenology and analytic philosophy of mind have not been closely joined, despite overlapping areas of interest. So it is appropriate to close this survey of phenomenology by addressing philosophy of mind, one of the most vigorously debated areas in recent philosophy.

The tradition of analytic philosophy began, early in the 20th century, with analyses of language, notably in the works of Gottlob Frége, Bertrand Russell, and Ludwig Wittgenstein. Then in The Concept of Mind (1949) Gilbert Ryle developed a series of analyses of language about different mental states, including sensation, belief, and will. Though Ryle is commonly deemed a philosopher of ordinary language, Ryle himself said The Concept of Mind could be called phenomenology. In effect, Ryle analysed our Phenomenological understanding of mental states as reflected in ordinary language about the mind. From this linguistic phenomenology Ryle argued that Cartesian mind-body dualism involves a category mistake (the logic or grammar of mental verbs - ‘believe,’ ‘see,’ etc. -does not mean that we ascribe belief, sensation, etc., to ‘the ghost in the machine’). With Ryle's rejection of mind-body dualism, the mind-body problem was re-awakened: What is the ontology of mind/body, and how are mind and body related?

René Descartes, in his epoch-making Meditations on First Philosophy (1641), had argued that minds and bodies are two distinct kinds of being or substance with two distinct kinds of attributes or modes: bodies are characterized by spatiotemporal physical properties, while minds are characterized by properties of thinking (including seeing, feeling, etc.). Centuries later, phenomenology would find, with Brentano and Husserl, that mental acts are characterized by consciousness and intentionality, while natural science would find that physical systems are characterized by mass and force, ultimately by gravitational, electromagnetic, and quantum fields. Where do we find consciousness and intentionality in the quantum-electromagnetic-gravitational field that, by hypothesis, orders everything in the natural world in which we humans and our minds exist? That is the mind-body problem today. In short, phenomenology by any other name lies at the heart of the contemporary, mind-body problem.

After Ryle, philosophers sought a more explicit and generally naturalistic ontology of mind. In the 1950s materialism was argued anew, urging that mental states are identical with states of the central nervous system. The classical identity theory holds that each token mental state (in a particular person's mind at a particular time) is identical with a token brain state (in that person's brain at that time). A weaker materialism holds, instead, that each type of mental state is identical with a type of brain state. But materialism does not fit comfortably with phenomenology. For it is not obvious how conscious mental states as we experience them - sensations, thoughts, emotions - can simply be the complex neural states that somehow subserve or implement them. If mental states and neural states are simply identical, in token or in type, where in our scientific theory of mind does the phenomenology occur - is it not simply replaced by neuroscience? And yet experience is part of what is to be explained by neuroscience.

In the late 1960s and 1970s the computer model of mind set it, and functionalism became the dominant model of mind. On this model, mind is not what the brain consists in (electrochemical transactions in neurons in vast complexes). Instead, mind is what brains do: They are function of mediating between information coming into the organism and behaviour proceeding from the organism. Thus, a mental state is a functional state of the brain or of the human or an animal organism. More specifically, on a favourite variation of functionalism, the mind is a computing system: Mind is to brain as software is to hardware; Thoughts are just programs running on the brain's ‘NetWare.’ Since the 1970s the cognitive sciences - from experimental studies of cognition to neuroscience - have tended toward a mix of materialism and functionalism. Gradually, however, philosophers found that Phenomenological aspects of the mind pose problems for the functionalist paradigm too.

In the early 1970s Thomas Nagel argued in ‘What Is It Like to Be a Bat?’ (1974) that consciousness itself - especially the subjective character of what it is like to have a certain type of experience - escapes physical theory. Many philosophers pressed the case that sensory qualia - what it is like to feel pain, to see red, etc. - are not addressed or explained by a physical account of either brain structure or brain function. Consciousness has properties of its own. And yet, we know, it is closely tied to the brain. And, at some level of description, neural activities implement computation.

In the 1980s John Searle argued in Intentionality (1983) (and further in The Rediscovery of the Mind (1991)) that intentionality and consciousness are essential properties of mental states. For Searle, our brains produce mental states with properties of consciousness and intentionality, and this is all part of our biology, yet consciousness and intentionality require the ‘first-person’ ontology. Searle also argued that computers simulate but do not have mental states characterized by intentionality. As Searle argued, a computer system has of the syntax (processing symbols of certain shapes) but has no semantics (the symbols lack meaning: We interpret the symbols). In this way Searle rejected both materialism and functionalism, while insisting that mind is a biological property of organisms like us: Our brains ‘secrete’ consciousness.

The analysis of consciousness and intentionality is central to phenomenology as appraised above, and Searle's theory of intentionality reads like a modernized version of Husserl's. (Contemporary logical theory takes the form of stating truth conditions for propositions, and Searle characterizes a mental state's intentionality by specifying its ‘satisfaction conditions’). However, there is an important difference in background theory. For Searle explicitly assumes the basic world view of natural science, holding that consciousness is part of nature. But Husserl explicitly brackets that assumption, and later phenomenologists - including Heidegger, Sartre, Merleau-Ponty - seem to seek a certain sanctuary for phenomenology beyond the natural sciences. And yet phenomenology itself should be largely neutral about further theories of how experience arises, notably from brain activity.

The philosophy or theory of mind overall may be factored into the following disciplines or ranges of theory relevant to mind: Phenomenology studies conscious experience as experienced, analysing the structure - the types, intentional forms and meanings, dynamics, and (certain) enabling conditions - of perception, thought, imagination, emotion, and volition and action.

Neuroscience studies the neural activities that serve as biological substrate to the various types of mental activity, including conscious experience. Neuroscience will be framed by evolutionary biology (explaining how neural phenomena evolved) and ultimately by basic physics (explaining how biological phenomena are grounded in physical phenomena). Here lie the intricacies of the natural sciences. Part of what the sciences are accountable for is the structure of experience, analysed by phenomenology.

Cultural analysis studies the social practices that help to shape or serve as cultural substrate of the various types of mental activity, including conscious experience. Here we study the import of language and other social practices. Ontology of mind studies the ontological type of mental activity in general, ranging from perception (which involves causal input from environment to experience) to volitional action (which involves causal output from volition to bodily movement).

This division of labour in the theory of mind can be seen as an extension of Brentano's original distinction between descriptive and genetic psychology. Phenomenology offers descriptive analyses of mental phenomena, while neuroscience (and wider biology and ultimately physics) offers models of explanation of what causes or gives rise to mental phenomena. Cultural theory offers analyses of social activities and their impact on experience, including ways language shapes our thought, emotion, and motivation. And ontology frames all these results within a basic scheme of the structure of the world, including our own minds.

Meanwhile, from an epistemological standpoint, all these ranges of theory about mind begin with how we observe and reason about and seek to explain phenomena we encounter in the world. And that is where phenomenology begins. Moreover, how we understand each piece of theory, including theory about mind, is central to the theory of intentionality, as it were, the semantics of thought and experience in general. And that is the heart of phenomenology.

There is potentially a rich and productive interface between neuroscience/cognitive science. The two traditions, however, have evolved largely independent, based on differing sets of observations and objectives, and tend to use different conceptual frameworks and vocabulary representations. The distributive contributions to each their dynamic functions of finding a useful common reference to further exploration of the relations between neuroscience/cognitive science and psychoanalysis/psychotherapy.

Forthwith, is the existence of a historical gap between neuroscience/cognitive science and psychotherapy is being productively closed by, among other things, the suggestion that recent understandings of the nervous system as a modeler and predictor bear a close and useful similarity to the concepts of projection and transference. The gap could perhaps be valuably narrowed still further by a comparison in the two traditions of the concepts of the ‘unconscious’ and the ‘conscious’ and the relations between the two. It is suggested that these be understood as two independent ‘story generators’ - each with different styles of function and both operating optimally as reciprocal contributors to each others' ongoing story evolution. A parallel and comparably optimal relation might be imagined for neuroscience/cognitive science and psychotherapy.

For the sake of argument, imagine that human behaviour and all that it entails (including the experience of being a human and interacting with a world that includes other humans) is a function of the nervous system. If this were so, then there would be lots of different people who are making observations of (perhaps different) aspects of the same thing, and telling (perhaps different) stories to make sense of their observations. The list would include neuroscientists and cognitive scientists and psychologists. It would include as well psychoanalysts, psychotherapists, psychiatrists, and social workers. If we were not too fussy about credentials, it should probably include as well educators, and parents and . . . babies? Arguably, all humans, from the time they are born, spend significant measures of their time making observations of how people (others and themselves) behave and why, and telling stories to make sense of those observations.

The stories, of course, all differ from one another to greater or lesser degrees. In fact, the notion that ‘human behaviour and all that it entails . . . is a function of the nervous system’ is itself a story used to make sense of observations by some people and not by others. It is not my intent here to try to defend this particular story, or any other story for that matter. Very much to the contrary, is to explore the implications and significance of the fact that there ARE different stories and that they might be about the same (some)thing

In so doing, I want to try to create a new story that helps to facilitate an enhanced dialogue between neuroscience/cognitive science, on the one hand, and psychotherapy, on the other. That new story is itself is a story of conflicting stories within . . . what is called the ‘nervous system’ but others are free to call the ‘self,’ ‘mind,’ ‘soul,’ or whatever best fits their own stories. What is important is the idea that multiple things, evident by their conflicts, may not in fact be disconnected and adversarial entities but could rather be fundamentally, understandably, and valuably interconnected parts of the same thing.

Many practising psychoanalysts (and psychotherapists too, I suspect) feel that the observations/stories of neuroscience/cognitive science are for their own activities, least of mention, are at primes of irrelevance, and at worst destructive, and the same probable holds for many neuroscientists/cognitive scientists. Pally clearly feels otherwise, and it is worth exploring a bit why this is so in her case. A general key, I think, is in her line ‘In current paradigms, the brain has intrinsic activity, is highly integrated, is interactive with the environment, and is goal-oriented, with predictions operating at every level, from lower systems to . . . the highest functions of abstract thought.’ Contemporary neuroscience/cognitive science has indeed uncovered an enormous complexity and richness in the nervous system, ‘making it not so different from how psychoanalysts (or most other people) would characterize the self, at least not in terms of complexity, potential, and vagary.’ Given this complexity and richness, there is substantially less reason than there once was to believe psychotherapists and neuroscientists/cognitive scientists are dealing with two fundamentally different thing’s ally suspects, more aware of this than many psychotherapists because she has been working closely with contemporary neuroscientists who are excited about the complexity to be found in the nervous system. And that has an important lesson, but there is an additional one at least as important in the immediate context. In 1950, two neuroscientists wrote: ‘The sooner we realize that not to expect of expectation itself, which we would recognize the fact that the complex and higher functional Gestalts that leave the reflex physiologist dumfounded in fact send roots down to the simplest basal functions of the CNS, the sooner we will see that the previously terminologically insurmountable barrier between the lower levels of neurophysiology and higher behavioural theory simply dissolves away.’

And in 1951 another said, ‘I am becoming subsequently forwarded by the conviction that the rudiments of every behavioural mechanism will be found far down in the evolutionary scale and represented in primitive activities of the nervous system.’

Neuroscience (and what came to be cognitive science) was engaged from very early on in an enterprise committed to the same kind of understanding sought by psychotherapists, but passed through a phase (roughly from the 1950's to the 1980's) when its own observations and stories were less rich in those terms. It was a period that gave rise to the notion that the nervous system was ‘simple’ and ‘mechanistic,’ which in turn made neuroscience/cognitive science seem less relevant to those with broader concerns, perhaps even threatening and apparently adversarial if one equated the nervous system with ‘mind,’ or ‘self,’ or ‘soul,’ since mechanics seemed degrading to those ideas. Arguably, though, the period was an essential part of the evolution of the contemporary neuroscience/cognitive science story, one that laid needed groundwork for rediscovery and productive exploration of the richness of the nervous system. Psychoanalysis/psychotherapy of course went through its own story evolution over this time. That the two stories seemed remote from one another during this period was never adequate evidence that they were not about the same thing but only an expression of their needed independent evolutions.

An additional reason that Pally is comfortable with the likelihood that psychotherapists and neuroscientists/cognitive scientists are talking about the same thing is her recognition of isomorphism (or congruities, Pulver 2003) between the two sets of stories, places where different vocabularies in fact seem to be representing the same (or quite similar) things. I am not sure I am comfortable calling these ‘shared assumptions’ (as Pally does) since they are actually more interesting and probably more significant if they are instead instances of coming to the same ideas from different directions (as I think they are). In this case, the isomorphism tend to imply that, rephrasing Gertrude Stein, ‘that there exists an actual there.’ Regardless, Pally has entirely appropriately and, I think, usefully called attention to an important similarity between the psychotherapeutic concept of ‘transference’ and an emerging recognition within neuroscience/cognitive science that the nervous system does not so much collect information about the world as generate a model of it, act in relation to that model, and then check incoming information against the predictions of that model. Pally's suggestion that this model reflects in part early interpersonal experiences, can be largely ‘unconscious,’ and so may cause inappropriate and troubling behaviour in current time seems entirely reasonable. So too is she to think of thoughts that there is an interaction with the analyst, and this can be of some help by bringing the model to ‘consciousness’ through the intermediary of recognizing the transference onto the analyst.

The increasing recognition of substantial complexity in the nervous system together with the presence of identifiable isomorphism provides a solid foundation for suspecting that psychotherapists and neuroscientists/cognitive scientists are indeed talking about the same thing. But the significance of different stories for better understanding a single thing lies as much in the differences between the stories as it does in their similarities/isomorphism, in the potential for differing and not obviously isomorphic stories productively to modify each other, yielding a new story in the process. With this thought in mind, I want to call attention to some places where the psychotherapeutic and the neuroscientific/cognitive scientific stories have edges that rub against one another rather than smoothly fitting together. And perhaps to ways each could be usefully further evolved in response to those non-isomorphism.

Unconscious stories and ‘reality.’ Though her primary concern is with interpersonal relations, Pally clearly recognizes that transference and related psychotherapeutic phenomena are one (actually relatively small) facet of a much more general phenomenon, the creation, largely unconsciously, of stories that are understood to be that of what are not necessarily reflective of the ‘real world.’ Ambiguous figures illustrate the same general phenomenon in a much simpler case, that of visual perception. Such figures may be seen in either of two ways; They represent two ‘stories’ with the choice between them being, at any given time, largely unconscious. More generally, a serious consideration of a wide array of neurobiological/cognitive phenomena clearly implies that, as Pally said, that if we could ever see ‘reality,’ but only have stories to describe it that result from processes of which we are not consciously aware.

All of this raises some quite serious philosophical questions about the meaning and usefulness of the concept of ‘reality.’ In the present context, what is important is that it is a set of questions that sometimes seem to provide an insurmountable barrier between the stories of neuroscientists/cognitive scientists, who by and large think they are dealing with reality, and psychotherapists, who feel more comfortable in more idiosyncratic and fluid spaces. In fact, neuroscience and cognitive science can proceed perfectly well in the absence of a well-defined concept of ‘reality’ and, without being fully conscious of it, does in fact do so. And psychotherapists actually make more use of the idea of ‘reality’ than is entirely appropriate. There is, for example, a tendency within the psychotherapeutic community to presume that unconscious stories reflect ‘traumas’ and other historically verifiable events, while the neurobiological/cognitive science story says quite clearly that they may equally reflect predispositions whose origins reflect genetic information and hence bear little or no relation to ‘reality’ in the sense usually meant. They may, in addition, reflect random ‘play’ (Grobstein, 1994), putting them even further out of reach of easy historical interpretation. In short, with regard to the relation between ‘story’ and ‘reality,’ each set of stories could usefully be modified by greater attention to the other. Differing concepts of ‘reality’ (perhaps the very concept itself) gets in the way of usefully sharing stories. The neurobiologists and/or/cognitive scientists' preoccupation with ‘reality’ as an essential touchstone could valuably be lessened, and the therapist's sense of the validation of story in terms of personal and historical idiosyncrasies could be helpfully adjusted to include a sense of actual material underpinnings.

The Unconscious and the Conscious. Pally appropriately makes a distinction between the unconscious and the conscious, one that has always been fundamental to psychotherapy. Neuroscience/cognitive science has been slower to make a comparable distinction but is now rapidly beginning to catch up. Clearly some neural processes generate behaviour in the absence of awareness and intent and others yield awareness and intent with or without accompanying behaviour. An interesting question however, raised at a recent open discussion of the relations between neuroscience and psychoanalysis, is whether the ‘neurobiological unconscious’ is the same thing as the ‘psychotherapeutic unconscious,’ and whether the perceived relations between the ‘unconscious’ and the’conscious’ are the same in the two sets of stories. Is this a case of an isomorphism or, perhaps more usefully, a masked difference?

An oddity of Pally's article is that she herself acknowledges that the unconscious has mechanisms for monitoring prediction errors and yet implies, both in the title of the paper, and in much of its argument, that there is something special or distinctive about consciousness (or conscious processing) in its ability to correct prediction errors. And here, I think, there is evidence of a potentially useful ‘rubbing of edges’ between the neuroscientific/cognitive scientific tradition and the psychotherapeutic one. The issue is whether one regards consciousness (or conscious processing) as somehow ‘superior’ to the unconscious (or unconscious processing). There is a sense in Pally of an old psychotherapeutic perspective of the conscious as a mechanism for overcoming the deficiencies of the unconscious, of the conscious as the wise father/mother and the unconscious as the willful child. Actually, Pally does not quite go this far, but there is enough of a trend to illustrate the point and, without more elaboration, I do not think many neuroscientists/cognitive scientists will catch Pally's more insightful lesson. I think Pally is almost certainly correct that the interplay of the conscious and the unconscious can achieve results unachievable by the unconscious alone, but think also that neither psychotherapy nor neuroscience/cognitive science are yet in a position to say exactly why this is so. So let me take a crack here at a new, perhaps bi-dimensional story that could help with that common problem and perhaps both traditions as well.

A major and surprising lesson of comparative neuroscience, supported more recently by neuropsychology (Weiskrantz, 1986) and, more recently still, by artificial intelligence is that an extraordinarily rich repertoire of adaptive behaviour can occur unconsciously, in the absence of awareness of intent (be supported by unconscious neural processes). It is not only modelling of the world and prediction and error correction that can occur this way but virtually (and perhaps literally) the entire spectrum of behaviour externally observed, including fleeing from threat, approaching good things, generating novel outputs, learning from doing so, and so on.

This extraordinary terrain, discovered by neuroanatomists, electrophysiologists, neurologists, behavioural biologists, and recently extended by others using more modern techniques, is the unconscious of which the neuroscientists/cognitive scientist speaks. It is a terrain so surprisingly rich that it creates, for some people, the inpuzzlement about whether there is anything else at all. Moreover, it seems, at first glance, to be a totally different terrain from that of the psychotherapist, whose clinical experience reveals a territory occupied by drives, unfulfilled needs, and the detritus with which the conscious would prefer not to deal.

As indicated earlier, it is one of the great strengths of Pally's article to suggest that the two terrains may in fact turns out to be the same in many ways, but if they are of the same line, it then becomes the question of whether or not it feels in what way nature resembles the ‘unconscious’ and the ‘conscious’ different? Where now are the ‘two stories?’ Pally touches briefly on this point, suggesting that the two systems differ not so much (or at all?) in what they do, but rather in how they do it. This notion of two systems with different styles seems to me worth emphasizing and expanding. Unconscious processing is faster and handles many more variables simultaneously. Conscious processing is slower and handles several variables at one time. It is likely that there appear to a host of other differences in style as well, in the handling of number for example, and of time.

In the present context, however, perhaps the most important difference in style is one that Lacan called attention to from a clinical/philosophical perspective - the conscious (conscious processing) that has in resemblance to some objective ‘coherence,’ that is, it attempts to create a story that makes sense simultaneously of all its parts. The unconscious, on the other hand, is much more comfortable with bits and pieces lying around with no global order. To a neurobiologist/cognitive scientist, this makes perfectly good sense. The circuitry includes the unconscious (sub-cortical circuitry?) assembly of different parts organized for a large number of different specific purposes, and only secondarily linked together to try to assure some coordination? The circuitry preserves the conscious precessing (neo-cortical circuitry?), that, on the other hand, seems to both be more uniform and integrated and to have an objective for which coherence is central.

That central coherence is well-illustrated by the phenomena of ‘positive illusions,’ exemplified by patients who receive a hypnotic suggestion that there is an object in a room and subsequently walk in ways that avoid the object while providing a variety of unrelated explanations for their behaviour. Similar ‘rationalization’ is, of course, seen in schizophrenic patients and in a variety of fewer dramatic forms in psychotherapeutic settings. The ‘coherent’ objective is to make a globally organized story out of the disorganized jumble, a story of (and constituting) the ‘self.’

What this thoroughly suggests is that the mind/brain be actually organized to be constantly generating at least two different stories in two different styles. One, written by conscious processes in simpler terms, is a story of/about the ‘self’ and experienced as such, for developing insights into how such a story can be constructed using neural circuitry. The other is an unconscious ‘story’ about interactions with the world, perhaps better thought of as a series of different ‘models’ about how various actions relate to various consequences. In many ways, the latter is the grist for the former.

In this sense, we are safely back to the two story ideas that has been central to psychotherapy, but perhaps with some added sophistication deriving from neuroscience/cognitive science. In particular, there is no reason to believe that one story is ‘better’ than the other in any definitive sense. They are different stories based on different styles of story telling, with one having advantages in certain sorts of situations (quick responses, large numbers of variables, more direct relation to immediate experiences of pain and pleasure) and the other in other sorts of situations (time for more deliberate responses, challenges amenable to handling using smaller numbers of variables, more coherent, more able to defer immediate gratification/judgment.

In the clinical/psychotherapeutic context, an important implication of the more neutral view of two story-tellers outlined above is that one ought not to over-value the conscious, nor to expect miracles of the process of making conscious what is unconscious. In the immediate context, the issue is if the unconscious is capable of ‘correcting prediction errors,’ then why appeal to the conscious to achieve this function? More generally, what is the function of that persistent aspect of psychotherapy that aspires to make the unconscious conscious? And why is it therapeutically effective when it is? Here, it is worth calling special attention to an aspect of Pally's argument that might otherwise get a bit lost in the details of her article: . . . the therapist encourages the wife to stop consciously and consider her assumption that her husband does not properly care about her, and to effort fully consider an alternative view and inhibit her impulse to reject him back. This, in turn, creates a new type of experience, one in which he is indeed more loving, such that she can develop new predictions.’

It is not, as Pally describes it, the simple act of making something conscious that is therapeutically effective. What is necessary is too consciously decompose the story (something that is made possible by its being a story with a small number of variables) and, even more important, to see if the story generates a new ‘type of experience’ that in turn causes the development of ‘new predictions.’ The latter, I suggest, is an effect of the conscious on the unconscious, an alteration of the unconscious brought about by hearing, entertaining, and hence acting on a new story developed by the conscious. It is not ‘making things conscious’ that is therapeutically effective; it is the exchange of stories that encourages the creation of a new story in the unconscious.

For quite different reasons, Grey (1995) earlier made a suggestion not dissimilar to Pally's, proposing that consciousness was activated when an internal model detected a prediction failure, but acknowledged he could see no reason ‘why the brain should generate conscious experience of any kind at all.’ It seems to me that, despite her title, it is not the detection of prediction errors that is important in Pally's story. Instead, it is the detection of mismatches between two stories, one unconscious and the other conscious, and the resulting opportunity for both to shape a less trouble-making new story. That, in brief, is it to why the brain ‘should generate conscious experience,’ and reap the benefits of having a second story teller with which a different style of paraphrasing Descartes, one might know of another in what one might say ‘I am, and I can think, therefore I can change who I am.’ It is not only the neurobiological ‘conscious’ that can undergo change; it is the neurobiological ‘unconscious’ as well.

More generally, the most effective psychotherapy requires the recognitions that assume their responsibility is rapidly emerging from neuroscience/cognitive science, that the brain/mind has evolved with two (or more) independent story tellers and has done so precisely because there are advantages to having independent story tellers that generate and exchange different stories. The advantage is that each can learn from the other, and the mechanisms to convey the stories and forth and for each story teller to learn from the stories of the other are a part of our evolutionary endowment as well. The problems that bring patients into a therapist's office are problems in the breakdown of story exchange, for any of a variety of reasons, and the challenge for the therapist is to reinstate the confidence of each story teller in the value of the stories created by the other. Neither the conscious nor the unconscious is primary; they function best as an interdependent loop with each developing its own story facilitated by the semi-independent story of the other. In such an organization, there is not only no ‘real,’ and no primacy for consciousness, there is only the ongoing development and, ideally, effective sharing of different stories.

There are, in the story I am outlining, implications for neuroscience/cognitive science as well. The obvious key questions are what does one mean (in terms of neurons and neuronal assemblies) by ‘stories,’ and in what ways are their construction and representation different in unconscious and conscious neural processing. But even more important, if the story I have outlined makes sense, what are the neural mechanisms by which unconscious and conscious stories are exchanged and by which each kind of story impacts on the other? And why (again in neural terms) does the exchange sometimes break down and fail in a way that requires a psychotherapist - an additional story teller - to be repaired?

Just as the unconscious and the conscious are engaged in a process of evolving stories for separate reasons and using separate styles, so too have been and will continue to be neuroscience/cognitive science and psychotherapy. And it is valuable that both communities continue to do so. But there is every reason to believe that the different stories are indeed about the same thing, not only because of isomorphism between the differing stories but equally because the stories of each can, if listened to, be demonstrably of value to the stories of the other. When breakdowns in story sharing occur, they require people in each community who are daring enough to listen and be affected by the stories of the other community. Pally has done us all a service as such a person. I hope my reactions to her article will help further to construct the bridge she has helped to lay, and that others will feel inclined to join in an act of collective story telling that has enormous intellectual potential and relates as well very directly to a serious social need in the mental health arena. Indeed, there are reasons to believe that an enhanced skill at hearing, respecting, and learning from differing stories about similar things would be useful in a wide array of contexts.

There is now a more satisfactory range of ideas available [in the field of consciousness studies] . . . They involve mostly quantum objects called Bose-Einstein condensates that may be capable of forming ephemeral but extended structures in the brain (Pessa). Marshall's original idea (based on the work of Frölich) was that the condensates that comprise the physical basis of mind, form from activity of vibrating molecules (dipoles) in nerve cell membranes. One of us (Clarke) has found theoretical evidence that the distribution of energy levels for such arrays of molecules prevents this happening in the way that Marshall first thought. However, the occurrence of similar condensates centring around the microtubules that are an important part of the structure of every cell, including nerve cells, remains a theoretical possibility (del Giudice et al.). Hameroff has pointed out that single-cell organisms such as 'paramecium' can perform quite complicated actions normally thought to need a brain. He suggests that their 'brain' be in their microtubules. Shape changes in the constituent proteins (tubulin) could subserve computational functions and would involve quantum phenomena of the sort envisaged by del Giudice. This raises the intriguing possibility that the most basic cognitive unit is provided, not by the nerve cell synapse as is usually supposed, but by the microtubular structure within cells. The underlying intuition is that the structures formed by Bose-Einstein condensates are the building Forms of mental life; in relation to perception they are models of the world, transforming a pleasant view, say, into a mental structure that represents some of the inherent qualities of that view.

We thought that, if there is anything to ideas of this sort, the quantum nature of awareness should be detectable experimentally. Holism and non-locality are features of the quantum world with no precise classical equivalents. The former presupposes that the interacting systems have to be considered as wholes - you cannot deal with one part in isolation from the rest. Non-locality means, among other things, that spatial separation between its parts does not alter the requirement to deal with an interacting system holistically. If we could detect these in relation to awareness, we would show that consciousness cannot be understood solely in terms of classical concepts.

Generative thought and words are the attempts to discover the relation between thought and speech at the earliest stages of phylogenetic and ontogenetic development. We found no specific interdependence between the genetic roots of thought and of word. It became plain that the inner relationship we were looking for was not a prerequisite for, but rather a product of, the historical development of human consciousness.

In animals, even in anthropoids whose speech is phonetically like human speech and whose intellect is akin to man’s, speech and thinking are not interrelated. A prelinguistic epoché through which times interval in thought and a preintellectual period in speech undoubtedly exist also in the development of the child. Thought and word are not connected by a primary bond. A connection originates, changes, and grows in the course of the evolution of thinking and speech.

It would be wrong, however, to regard thought and speech as two unrelated processes either parallel or crossing at certain points and mechanically influencing each other. The absence of a primary bond does not mean that a connection between them can be formed only in a mechanical way. The futility of most of the earlier investigations was largely due to the assumption that thought and word were isolated, independent elements, and verbal thought the fruit of their external union.

The method of analysis based on this conception was bound to fail. It sought to explain the properties of verbal thought by breaking it up into its component elements, thought and word, neither of which, taken separately, possessed the properties of the whole. This method is not true analysis helpful in solving concrete problems. It leads, rather, to generalisation. We compared it with the analysis of water into hydrogen and oxygen - which can result only in findings applicable to all water existing in nature, from the Pacific Ocean to a raindrop. Similarly, the statement that verbal thought is composed of intellectual processes and speech is functionally proper applications to all verbal thought and all its manifestations and explains none of the specific problems facing the student of verbal thought.

We tried a new approach to the subject and replaced analysis into elements by analysis into units, each of which retains in simple form all the properties of the whole. We found this unit of verbal thought in word meaning.

The meaning of a word represents such a close amalgam of thought and language that it is hard to tell whether it is a phenomenon of speech or a phenomenon of thought. A word without meaning is an empty sound; meaning, therefore, is a criterion of ‘word,’ its indispensable component. It would seem, then, that it may be regarded as a phenomenon of speech. But from the point of view of psychology, the meaning of every word is a generalisation or a concept. And since generalisations and concepts are undeniably acts of thought, but we may regard meaning as a phenomenon of thinking. It does not follow, however, that meaning formally belongs in two different spheres of psychic life. Word meaning is a phenomenon of thought only insofar as thought is embodied in speech, and of speech only insofar as speech is connected with thought and illumined by it. It is a phenomenon of verbal thought, or meaningful speech - a union of word and thought.

Our experimental investigations fully confirm this basic thesis. They not only proved that concrete study of the development of verbal thought is made possible by the use of word meaning as the analytical unit but they also led to a further thesis, which we consider the major result of our study and which issues directly from the further thesis that word meanings develop. This insight must replace the postulate of the immutability of word meanings.

From the point of view of the old schools of psychology, the bond between word and meaning is an associative bond, established through the repeated simultaneous perception of a certain sound and a certain object. A word calls to mind its content as the overcoat of a friend reminds us of that friend, or a house of its inhabitants. The association between word and meaning may grow stronger or weaker, be enriched by linkage with other objects of a similar kind, spread over a wider field, or become more limited, i.e., it may undergo quantitative and external changes, but it cannot change its psychological nature. To do that, it would have to cease being an association. From that point of view, any development in word meanings is inexplicable and impossible - an implication that impeded linguistics as well as psychology. Once having committed itself to the association theory, semantics persisted in treating word meaning as an association between a word’s sound and its content. All words, from the most concrete to the most abstract, appeared to be formed in the same manner in regard to meaning, and to contain nothing peculiar to speech as such; a word made us think of its meaning just as any object might remind us of another. It is hardly surprising that semantics did not even pose the larger question of the development of word meanings. Development was reduced to changes in the associative connections between single words and single objects: A word brawn to denote at first one object and then become associated with another, just as an overcoat, having changed owners, might remind us first of one person and later of another. Linguistics did not realize that in the historical evolution of language the very structure of meaning and its psychological nature also change. From primitive generalisations, verbal thought rises to the most abstract concepts. It is not merely the content of a word that changes, but the way in which reality is generalised and reflected in a word.

Equally inadequate is the association theory in explaining the development of word meanings in childhood. Here, too, it can account only for the pure external, quantitative changes in the bonds uniting word and meaning, for their enrichment and strengthening, but not for the fundamental structural and psychological changes that can and do occur in the development of language in children.

Oddly enough, the fact that associationism in general had been abandoned for some time did not seem to affect the interpretation of word and meaning. The Wuerzburg school, whose main object was to prove the impossibility of reducing thinking to a mere play of associations and to demonstrate the existence of specific laws governing the flow of thought, did not revise the association theory of word and meaning, or even recognise the need for such a revision. It freed thought from the fetters of sensation and imagery and from the laws of association, and turned it into a purely spiritual act. By so doing, it went back to the prescientific concepts of St. Augustine and Descartes and finally reached extreme subjective idealism. The psychology of thought was moving toward the ideas of Plato. Speech, at the same time, was left at the mercy of association. Even after the work of the Wuerzburg school, the connection between a word and its meaning was still considered a simple associative bond. The word was seen as the external concomitant of thought, its attire only, having no influence on its inner life. Thought and speech had never been as widely separated as during the Wuerzburg period. The overthrow of the association theory in the field of thought actually increased its sway in the field of speech.

The work of other psychologists further reinforced this trend. Selz continued to investigate thought without considering its relation to speech and came to the conclusion that man’s productive thinking and the mental operations of chimpanzees were identical in nature – so completely did he ignore the influence of words on thought.

Even Ach, who made a special studies in phraseology, by the meaning of who tried to overcome the correlation in his theory of concepts, did not go beyond assuming the presence of ‘determining tendencies’ operative, along with associations, in the process of concept formation. Hence, the conclusions he reached did not change the old understanding of word meaning. By identifying concept with meaning, he did not allow for development and changes in concepts. Once established, the meaning of a word was set forever; Its development was completed. The same principles were taught by the very psychologists Ach attacked. To both sides, the starting point was also the end of the development of a concept; the disagreement concerned only the way in which the formation of word meanings began.

In Gestalt psychology, the situation was not very different. This school was more consistent than others in trying to surmount the general principle of a collective associationism. Not satisfied with a partial solution of the problem, it tried to liberate thinking and speech from the rule of association and to put both under the laws of structure formation. Surprisingly, even this most progressive of modern psychological schools made no progress in the theory of thought and speech.

For one thing, it retained the complete separation of these two functions. In the light of Gestalt psychology, the relationship between thought and word appears as a simple analogy, a reduction of both to a common structural denominator. The formation of the first meaningful words of a child is seen as similar to the intellectual operations of chimpanzees in Koehler’s experiments. Words that filter through the structure of things and acquire a certain functional meaning, in much the same way as the stick, to the chimpanzee, becomes part of the structure of obtaining the fruit and acquires the functional meaning of tool, that the connection between word and meaning is no longer regarded as a matter of simple association but as a matter of structure. That seems like a step forward. But if we look more closely at the new approach, it is easy to see that the step forward is an illusion and that we are still standing in the same place. The principle of structure is applied to all relations between things in the same sweeping, undifferentiated way as the principle of association was before it. It remains impossible to deal with the specific relations between word and meaning.

They are from the outset accepted as identical in principle with any and all other relations between things. All cats are as grey in the dusk of Gestalt psychology as in the earlier plexuities that assemble in universal associationism.

While Ach sought to overcome the associationism with ‘determining tendencies,’ Gestalt psychology combatted it with the principle of structure - retaining, however, the two fundamental errors of the older theory: the assumption of the identical nature of all connections and the assumption that word meanings do not change. The old and the new psychology both assume that the development of a word’s meaning is finished as soon as it emerges. The new trends in psychology brought progress in all branches except in the study of thought and speech. Here the new principles resemble the old ones like twins.

If Gestalt psychology is at a standstill in the field of speech, it has made a big step backward in the field of thought. The Wuerzburg school at least recognised that thought had laws of its own. Gestalt psychology denies their existence. By reducing to a common structural denominator the perceptions of domestic fowl, the mental operations of chimpanzees, the first meaningful words of the child, and the conceptual thinking of the adult, it obliterates every distinction between the most elementary perception and the highest forms of thought.

This may be summed up as follows: All the psychological schools and trends overlook the cardinal point that every thought is a generalisation. They all study word and meaning without any reference to development. As long as these two conditions persist in the successive trends, there cannot be much difference in the treatment of the problem.

The discovery that word meanings evolve leads the study of thought and speech out of a blind alley. Word meanings are dynamic rather than static formations. They change as the child develops; they change also with the various ways in which thought functions.

If word meanings change in their inner nature, then the relation of thought to word also changes. To understand the dynamics of that relationship, we must supplement the genetic approach of our main study by functional analysis and examine the role of word meaning in the process of thought.

Let us consider the process of verbal thinking from the first dim stirring of a thought to its formulation. What we want to show now is not how meanings develop over long periods of time but the way they function in the live process of verbal thought. On the basis of such a functional analysis, we will be able to show also that each stage in the development of word meaning has its own particular relationship between thought and speech. Since functional problems are most readily solved by examining the highest form of a given activity, we will, for a while, put aside the problem of development and consider the relations between thought and word in the mature mind.

The leading idea in the following discussion can be reduced to this formula: The relation of thought to word is not a thing but a process, a continual movement back and forth from thought to word and from word to thought. In that process the relation of thought to word undergoes changes that they may be regarded as development in the functional sense. Thought is not merely expressed in words; it comes into existence through them. Every thought tends to connect something with something else, to establish a relationship between things. Every thought moves, grows and develops, fulfils a function, solves a problem. This flow of thought occurs as an inner movement through a series of planes. An analysis of the interaction of thought and word must begin with an investigation of the different phases and planes a thought traverses before it is embodied in words.

The first thing such a study reveals is the need to distinguish between two planes of speech. Both the inner, meaningful, semantic aspect of speech and the external, phonetic aspects, though forming a true unity, have their own laws of movement. The unity of speech is a complex, not a homogeneous, unity. A number of facts in the linguistic development of the child indicate independent movement in the phonetic and the semantic spheres. We will point out two of the most important of these facts.

In mastering external speech, the child starts from one word, then connects two or three words; a little later, he advances from simple sentences to more complicated ones, and finally to coherent speech made up of series of such sentences; in other words, he proceeds from a part to the whole. In regard to meaning on the other hand, the first word of the child is a whole sentence. Semantically, the child starts from the whole, from a meaningful complex, and only later begins to master the separate semantic units, the meanings of words, and to divide his formerly undifferentiated thought into those units. The external and the semantic aspects of speech develop in opposite directions – one from the particular to the whole, from word to sentence, and the other from the whole to the particular, from sentence to word.

This in itself suffices to show how important it is to distinguish between the vocal and the semantic aspects of speech. Since they move in reverse directions, their development does not coincide, but that does not mean that they are independent of each other. On the contrary, their difference is the first stage of a close union. In fact, our example reveals their inner relatedness as clearly as it does their distinction. A child’s thought, precisely because it is born as a dim, amorphous whole, must find expression in a single word. As his thought becomes more differentiated, the child is less apt to express it in single words but constructs a composite whole. Conversely, progress in speech to the differentiated whole of a sentence helps the child’s thoughts to progress from a homogeneous whole to well-defined parts. Thought and word are not cut from one pattern. In a sense, there are more differences than likenesses between them. The structure of speech does not simply mirror the structure of thought that is why words cannot be put on by thought like a ready-made garment. Thought undergoes many changes as it turns into speech. It does not merely find expression in speech; It finds its reality and form. The semantic and the phonetic developmental processes are essentially one, precisely because of their reverse directions.

The second, equally important fact emerges at a later period of development. Piaget demonstrated that the child uses subordinate clauses with because, although, etc., long before he grasps the structures of meaning corresponding to these syntactic forms. Grammar precedes logic. Here, too, as in our previous example, the discrepancy does not exclude union but is, in fact, necessary for union.

In adults the divergence between the semantic and the phonetic aspects of speech is even more striking. Modern, psychologically oriented linguistics is familiar with this phenomenon, especially in regard too grammatical and psychological subject and predicate. For example, in the sentence ‘The clock fell,’ emphasis and meaning may change in different situations. Suppose I notice that the clock has stopped and ask how this happened. The answer is, ‘The clock fell.’ Grammatical and psychological subject coincide: ‘The clock’ is the first idea in my consciousness; ‘fell’ is what is said about the clock. But if I hear a crash in the next room and inquire what happened, and get the same answer, subject and predicate are psychologically reversed. I knew something had fallen – that is what we are talking about. ‘The clock’ completes the idea. The sentence could be changed to: ‘What has fallen is the clock’; Then the grammatical and the psychological subject would coincide. In the prologue to his play Duke Ernst von Schwaben, Uhland says: ‘Grim scenes will pass before you.’ Psychologically, ‘will pass’ is the subject. The spectator knows he will see events unfold the additional idea, the predicate, remains in ‘grim scenes.’ Uhland meant, ‘What will pass before your eyes are a tragedy.’ Any part of a sentence may become the psychological predicate, the carrier of topical emphasis: on the other hand, entirely different meanings may lie hidden behind one grammatical structure. Accord between syntactical and psychological organisation is not as prevalent as we tend to assume – rather, it is a requirement that is seldom met. Not only subject and predicate, but grammatical gender, number, case, tense, degree, etc. has their psychological doubles. A spontaneous utterance wrong from the point of view of grammar, may have charm and aesthetic value. Absolute correctness is achieved only beyond natural language, in mathematics. Our daily speech continually fluctuates between the ideals of mathematical and of imaginative harmony.

We will illustrate the interdependence of the semantic and the grammatical aspects of language by citing two examples that show that changes in formal structure can entail far-reaching changes in meaning.

In translating the fable ‘La Cigale et la Fourmi,’ Krylov substituted a dragonfly for La Fontaine’s grasshopper. In French Grasshopper is feminine and therefore well suited to symbolise a light-hearted, carefree attitude. The nuance would be lost in a literal translation, since in Russian Grasshopper is masculine. When he settled for dragonflies, which is feminine in Russian, Krylov disregarded the literal meaning in favour of the grammatical form required to render La Fontaine’s thought.

Tjutchev did the same in his translation of Heine’s poem about a fir and a palm. In German fir is masculine and palm feminine, and the poem suggests the love of a man for a woman. In Russian, both trees are feminine. To retain the implication, Tjutchev replaced the fir by a masculine cedar. Lermontov, in his more literal translation of the same poem, deprived it of these poetic overtones and gave it an essentially different meaning, more abstract and generalised. One grammatical detail may, on occasion, change the whole of which is to purport of what is said.

Behind words, there is the independent grammar of thought, the syntax of word meanings. The simplest utterance, far from reflecting a constant, rigid correspondence between sound and meaning, is really a process. Verbal expressions cannot emerge fully formed but must develop gradually. This complex process of transition from meaning to sound must itself be developed and perfected. The child must learn to distinguish between semantics and phonetics and understand the nature of the difference. At first he uses verbal forms and meanings without being conscious of them as separate. The word, to the child, is an integral part of the object it denotes. Such a conception seems to be characteristic of primitive linguistic consciousness. We all know the old story about the rustic who said he wasn’t surprised that savants with all their instruments could figure out the size of stars and their course – what baffled him was how they found out their names. Simple experiments show that preschool children ‘explain’ the names of objects by their attributes. According to them, an animal is called ‘cow’ because it has horns, ‘calves’ because its horns are still small, ‘dog’ because it is small and has no horns; an object is called ‘car’ because it is not an animal. When asked whether one could interchange the names of objects, for instance call a cow ‘ink,’ and ink ‘cow,’ children will answer no, ‘because ink is used for writing, and the cow gives milk.’ An exchange of names would mean an exchange of characteristic features, so inseparable is the connection between them in the child’s mind. In one experiment, the children were told that in a game a dog would be called ‘cow.’ Here is a typical sample of questions and answers: Does a cow have horns? ‘Yes.’ ‘But do you not remember that the cow is really a dog? Come now, does a dog have horns? ‘Sure, if it is a cow, if it is called cow, it has horns. That kind of dog has to have little horns.

We can see how difficult it is for children to separate the name of an object from its attributes, which cling to the name when it is transferred like possessions following their owner.

The fusion of the two planes of speech, semantic and vocal begins to break down as the child grows older, and the distance between them gradually increases. Each stage in the development of word meanings has its own specific interrelation of the two planes. A child’s ability to communicate through language is directly related to the differentiation of word meanings in his speech and consciousness.

To understand this, we must remember a basic characteristic of the structure of word meanings. In the semantic structure of a word, we distinguish between referent and meaning correspondingly, we distinguish a word’s nominative from its significative function. When we compare these structural and functional relations at the earliest, middle, and advanced stages of development, we find the following genetic regularity: In the beginning, only the nominative functions exist, and semantically, only the unbiased objective becomes the reference, and independent of naming, and meaning independent of reference, appear later and develop along the paths we have attempted to trace and describe.

Only when this development is completed does the child become fully able to formulate his own thought and to understand the speech of others. Until then, his usage of words coincides with that of adults in its objective reference but not in its meaning.

We must probe still deeper and explore the plane of inner speech lying beyond the semantic plane. We will discuss here some of the data of the special investigation we have made of it. The relationship of thought and word cannot be understood in all its complexity without a clear understanding of the psychological nature of inner speech. Yet, of all the problems connected with thought and language, this is perhaps the most complicated, beset as it is with terminological and other misunderstandings.

The term inner speech, or endophasy, has been applied to various phenomena, and authors argue about different things that they call by the same name. Originally, inner speech seems to have been understood as verbal memory. An example would be the silent recital of a poem known by heart. In that case, inner speech differs from vocal speech only as the idea or image of an object differs from the real object. It was in this sense that inner speech was understood by the French authors who tried to find out how words were reproduced in memory – whether as auditory, visual, motor, or synthetic images. We will see that word memory is indeed one of the constituent elements of inner speech but not all of it.

In a second interpretation, inner speech is seen as truncated external speech - as ‘speech minus sound’ (Mueller) or ‘sub-vocal speech’ (Watson). Bekhterev defined it as a speech reflex inhibited in its motor part. Such an explanation is by no measure of sufficiency. Silent ‘pronouncing’ of words is not equivalent to the total process of inner speech.

The third definition is, on the contrary, too broad. To Goldstein, the term covers everything that precedes the motor act of speaking, including Wundt’s ‘motives of speech’ and the indefinable, non-sensory and non-motor specific speech experience -, i.e., the whole interior aspect of any speech activity. It is hard to accept the equation of inner speech with an inarticulate inner experience in which the separate identifiable structural planes are dissolved without trace. This central experience is common to all linguistic activity, and for this reason alone Goldstein’s interpretation does not fit that specific, unique function that alone deserves the name of inner speech. Logically developed, Goldstein’s view must lead to the thesis that inner speech is not speech at all but rather an intellectual and affective-volitional activity, since it includes the motives of speech and the thought that is expressed in words.

To get a true picture of inner speech, one must embark upon that which is a specific formation, with its own laws and complex relations to the other forms of speech activity. Before we can study its relation to thought, on the one hand, and to speech, on the other, we must determine its special characteristics and function.

Inner speech allows one to speak for one’s external oration, for which of the others would be surprising, if such a difference in function did not affect the structure of the two kinds of speech. Absence of vocalisation per se is only a consequence of the specific nature of inner speech, which is neither an antecedent of external speech nor its reproduction in memory but is, in a sense, the opposite of external speech. The latter is the turning of thought into words, its materialisation and objectification. With inner speech, the process is reversed: Speech turns into inward thought. Consequently, their structures must differ.

The area of inner speech is one of the most difficult to investigate. It remained almost inaccessible to experiments until ways were found to apply the genetic method of experimentation. Piaget was the first to pay attention to the child’s egocentric speech and to see its theoretical significance, but he remained blind to the most important trait of egocentric speech - its genetic connection with inner speech – and this warped his interpretation of its function and structure. We made that relationship the central problem of our study and thus were able to investigate the nature of inner speech with unusual completeness. A number of considerations and observations led us to conclude that egocentric speech is a stage of development preceding inner speech: Both fulfil intellectual functions; Their structures are similar; egocentric speech disappears at school age, when inner speech begins to develop. From all this we infer that one change into the other.

If this transformation does take place, then egocentric speech provides the key to the study of inner speech. One advantage of approaching inner speech through egocentric speech is its accessibility to experimentation and observation. It is still vocalised, audible speech, i.e., external in its mode of expression, but at the same time inner speech in function and structure. To study an internal process, in that it is necessary to externalise it experimentally, by connecting it with some outer activity; barely then is objective functional analysis possible. Egocentric speech is, in fact, a natural experiment of this type.

This method has another great advantage: Since egocentric speech can be studied at the time when some of its characteristics are waning and new ones forming, we are able to judge which traits are essential to inner speech and which are only temporary, and thus to determine the goal of this movement from egocentric to inner speech -, i.e., the nature of inner speech.

Before we go on to the results obtained by this method, we will briefly discuss the nature of egocentric speech, stressing the differences between our theory and Piaget’s. Piaget contends that the child’s egocentric speech is a direct expression of the egocentrism of his thought, which in turn is a compromise between the primary autism of his thinking and its gradual socialisation. As the child grows older, and as autism overturns the associative remembers affiliated to socialisation progresses, leading to the waning of egocentrism in his thinking and speech.

In Piaget’s conception, the child in his egocentric speech does not adapt himself to the thinking of adults. His thought remains entirely egocentric; This makes his talk incomprehensibly to others. Egocentric speech has no function in the child’s realistic thinking or activity, but it merely accompanies them. And since it is an expression of egocentric thought, it disappears together with the child’s egocentrism. From its climax at the beginning of the child’s development, egocentric speech drops to zero on the threshold of school age. Its history is one of involution rather than evolution. It has no future.

In our conception, egocentric speech is a phenomenon of the transition from interpsychic to intrapsychic functioning, i.e., from the social, collective activity of the child to his more individualised activity - a pattern of development common to all the higher psychological functions. Speech for oneself originates through differentiation from speech for others. Since the main course of the child’s development is one of gradual individualisation, this tendency is reflected in the function and structure of his speech.

The function of egocentric speech is similar to that of inner speech: It does not merely accompany the child’s activity; it serves mental orientation, conscious understanding; it helps in overcoming difficulties; it is speech for oneself, intimately and usefully connected with the child’s thinking. Its fate is very different from that described by Piaget. Egocentric speech develops along a rising not a declining, curve; it goes through an evolution, not an involution. In the end, it becomes inner speech.

Our hypothesis has several advantages over Piaget’s: It explains the function and development of egocentric speech and, in particular, its sudden increase when the child’s face’s difficulties that demand consciousness and reflection – a fact uncovered by our experiments and which Piaget’s theory cannot explain. But the greatest advantage of our theory is that it supplies a satisfying answer to a paradoxical situation described by Piaget himself. To Piaget, the quantitative drop in egocentric speech as the child grows older means the withering of that form of speech. If that were so, its structural peculiarities might also be expected to decline; it is hard to believe that the process would affect only its quantity, and not its inner structure. The child’s thought becomes infinitely less egocentric between the ages of three and seven. If the characteristics of egocentric speech that make it incomprehensible to others are indeed rooted in egocentrism, they should become less apparent as that form of speech becomes less frequent; Egocentric speech should approach social speech and become ever more intelligible. Yet what are the facts? Is the talk of a three-year-old harder to follow than that of a seven-year-old? Our investigation established that the traits of egocentric speech that makes for inscrutability are at their lowest point at three and at their peak at seven. They develop in a reverse direction to the frequency of egocentric speech. While the latter keeps declining and reaches the point of zero at school age, the structural characteristics become more pronounced.

This throws a new light on the quantitative decrease in egocentric speech, which is the cornerstone of Piaget’s thesis.

What does this decrease mean? The structural peculiarities of speech for oneself and its differentiation from external speech increase with age. What is it that diminishes? Only one of its aspects verbalizes. Does this mean that egocentric speech as a whole is dying out? We believe that it does not, for how then could we explain the growth of the functional and structural traits of egocentric speech? On the other hand, their growth is perfectly compatible with the decrease of vocalisation - indeed, clarifies its meaning. Its rapid dwindling and the equally rapid growth of the other characteristics are contradictory in appearance only.

To explain this, let us start from an undeniable, experimentally established fact. The structural and functional qualities of egocentric speech become more marked as the child develops. At three, the difference between egocentric and social speech matches that to zero; At seven, we have speech that in structure and function is totally unlike social speech. A differentiation of the two speech functions has taken place. This is a fact - and facts are notoriously hard to refute.

Once we accept this, everything else falls into place. If the developing structural and functional peculiarities of egocentric speech progressively isolate it from external speech, then its vocal aspect must fade away. This is exactly what happens between three and seven years. With the progressive isolation of speech for oneself, its vocalisation becomes unnecessary and meaningless and, because of its growing structural peculiarities, also impossible. Speech for oneself cannot find expression in external speech. The more independent and autonomous egocentric speech becomes, the poorer it grows in its external manifestations. In the end it separates itself entirely from speech for others, ceases to be vocalised, and thus appears to die out.

But this is only an illusion. To interpret the sinking coefficient of egocentric speech as a sign that this kind of speech is dying out is like saying that the child stops counting when he ceases to use his fingers and starts adding in his head. In reality, behind the symptoms of dissolution lies a progressive development, the birth of a new speech form.

The decreasing vocalisation of egocentric speech denotes a developing abstraction from sound, the child’s new faculty to ‘think words’ instead of pronouncing them. This is the positive meaning of the sinking coefficient of egocentric speech. The downward curve indicates development toward inner speech.

We can see that all the known facts about the functional, structural, and genetic characteristics of self-indulgent or egocentric speech points to one thing: It develops in the direction of inner speech. Its developmental history can be understood only as a gradual unfolding of the traits of inner speech.

We believe that this corroborates our hypothesis about the origin and nature of egocentric speech. To turn our hypothesis into a certainty, we must devise an experiment capable of showing which of the two interpretations is correct. What are the data for this critical experiment?

Let us restate the theories between which we must decide as for Piaget believes, that egocentric speech stems from the insufficient socialisation of speech and that its only development is decrease and eventual death. Its culmination lies in the past. Inner speech is something new brought in from the outside along with socialisation. We demonstrated that in egocentric speech stems from the insufficient individualisation of primary social speech. Its culmination lies in the future. It develops into inner speech.

To obtain evidence for one or the other view, we must place the child alternately in experimental situations encouraging social speech and in situations discouraging it, and see how these changes affect egocentric speech. We consider this an experimentum crucis for the following reasons.

If the child’s egocentric talk results from the egocentrism of his thinking and its insufficient socialisation, then any weakening of the social elements in the experimental setup, any factor contributing to the child’s isolation from the group, must lead to a sudden increase in egocentric speech. But if the latter results from an insufficient differentiation of speech for oneself from speech for others, then the same changes must cause it to decrease.

We took as the starting point of our experiment three of Piaget’s own observations: (1) Egocentric speech occurs only in the presence of other children engaged in the same activity, and not when the child is alone; i.e., it is a collective monologue. (2) The child is under the illusion that his egocentric talk, directed to nobody, is understood by those who surround him. (3) Egocentric speech has the character of external speech: It is audible or whispered. These are certainly not chance peculiarities. From the child’s own point of view, egocentric speech is not yet separated from social speech. It occurs under the subjective and objective conditions of social speech and may be considered a correlate of the insufficient isolation of the child’s individual consciousness from the social whole.

In our first series of experiments, we tried to destroy the illusion of being understood. After measuring the child’s coefficient of egocentric speech in a situation similar to that of Piaget’s experiments, we put him into a new situation: Either with deaf-mute children or with children speaking a foreign language. In all other respects the setup remained the same. The coefficient of egocentric speech dropped to zero in the majority of cases, and in the rest to one-eighth of the previous figure, on the average. This proves that the illusion of being understood is not a mere epiphenomenon of egocentric speech but is functionally connected with it. Our results must seem paradoxical from the point of view of Piaget’s theory: The weaker the child’s contact is with the group – amounting to less of the social situation forces’ him to adjust his thoughts to others and to use social speech – that there is more as freely should be the egocentrism of his thinking and speech manifest itself. But from the point of view of our hypothesis, the meaning of these findings is clear: Egocentric speech, springing from the lack of differentiation of speech for oneself from speech for others, disappears when the feeling of being understood, essential for social speech, is absent.

In the second series of experiments, the variable factor was the possibility of some collective monologue. Having measured the child’s coefficient of egocentric speech in a situation permitting collective monologue, we put him into a situation excluding it - in a group of children who were strangers to him, or by his being of self, at which point, a separate table in a corner of the room, for which he worked entirely alone, even the experimenter leaving the room. The results of this series agreed with the first results. The exclusion of the group monologue caused a drop in the coefficient of egocentric speech, though not such a striking one as in the first case - seldom to zero and, on the average, to one-sixth of the original figure. The different methods of precluding a collective characterize monologues that were not equally effective in reducing the coefficient of egocentric speech. The trend, however, was obvious in all the variations of the experiment. The exclusion of the collective factor, instead of giving full freedom to egocentric speech, depressed it. Our hypothesis was once more confirmed.

In the third series of experiments, the variable factor was the vocal quality of egocentric speech. Just outside the laboratory where the experiment was in progress, an orchestra played so loudly, or so much noise was made, that it drowned out not only the voices of others but the child’s own; in a variant of the experiment, the child was expressly forbidden to talk loudly and allowed to talk only in whispers. Once again the coefficient of egocentric speech went down, the relation to the original unit being the different methods were not equally effective, but the basic trend was invariably present.

The purpose of all three series of experiments was to eliminate those characteristics of egocentric speech that bring it close to social speech. We found that this always led to the dwindling of egocentric speech. It is logical, then, to assume that egocentric speech is a form developing out of social speech and not yet separated from it in its manifestation, though already distinct in function and structure.

The disagreement between us and Piaget on this point will be made quite clear by the following example: I am sitting at my desk talking to a person who is behind me and whom I cannot see; he leaves the room without my noticing it, and I continue to talk, under the illusion that he listens and understands. Outwardly, I am talking with myself and for myself, but psychologically my speech is social. From the point of view of Piaget’s theory, the opposite happens in the case of the child: His egocentric talk is for and with himself; it only has the appearance of social speech, just as my speech gave the false impression of being egocentric. From our point of view, the whole situation is much more complicated than that: Subjectively, the child’s egocentric speech already has its own peculiar function - to that extent, it is independent from social speech; Yet its independence is not complete because it is not felt as inner speech and is not distinguished by the child from speech for others. Objectively, also, it is different from social speech but again not entirely, because it functions only within social situations. Both subjectively and objectively, egocentric speech represents a transition from speech for others to speech for oneself. It already has the function of inner speech but remains similar to social speech in its expression.

The investigation of egocentric speech has paved the way to the understanding of inner speech, while our experiments convinced us that inner speech must be regarded, not as speech minus sound, but as an entirely separate speech function. Its main distinguishing trait is its peculiar syntax. Compared with external speech, inner speech appears disconnected and incomplete.

This is not a new observation. All the students of inner speech, even those who approached it from the behaviouristic standpoint, noted this trait. The method of genetic analysis permits us to go beyond a mere description of it. We applied this method and found that as egocentric speech transforms by its showing tendencies toward an altogether specific form of abbreviation: Namely, omitting the subject of a sentence and all words connected with it, while preserving the predicate. This tendency toward predication appears in all our experiments with such regularity that we must assume it to be the basic syntactic form of inner speech.

It may help us to understand this tendency if we recall certain situations in which external speech shows a similar structure. Pure predication occurs in external speech in two cases: Either as an answer or when the subject of the sentence is known beforehand to all concerned. The answer to ‘Would you like a cup of tea?’ is never ‘No, I do not want a cup of tea ‘ but a simple ‘No.?’ Obviously, such a sentence is possible only because its subject is tacitly understood by both parties. To ‘Has your brother read this book?’ No one ever replies, ‘Yes, my brother has read this book.’ The answer is a short ‘Yes,’ or ‘Yes, he has.’ Now let us imagine that several people are waiting for a bus. No one will say, on seeing the bus approach, ‘The bus for which we are waiting is coming.’ The sentence is likely to be an abbreviated ‘Coming,’ or some such expression, because the subject is plain from the situation. Exceptionally hold to a frequent shortened sentence causing confusion. The listener may relate the sentence to a subject foremost in his own mind, not the one meant by the speaker. If the thoughts of two people coincide, perfect understanding can be achieved through the use of mere predicates, but if they are thinking about different things they are bound to misunderstand each other.

Having examined abbreviation in external speech, we can now return enriched to the same phenomenon in inner speech, where it is not an exception but the rule. It will be instructive to compare abbreviation in oral, inner, and written speech. Communication in writing relies on the formal meanings of words and requires a much greater number of words than oral speech to convey the same idea. It is addressed to an absent person who rarely has in mind the same subject as the writer. Therefore, it must be fully deployed; Syntactic differentiation is at a maximum, and expressions are used that would seem unnatural in conversation. Griboedov’s ‘He talks like writing’ refers to the droll effect of elaborate constructions in daily speech.

The multifunctional nature of language, which has recently attracted the close attention of linguists, had already been pointed out by Humboldt in relation to poetry and prose – two forms very different in function and in the means they use. Poetry, according to Humboldt, is inseparable from music, while prose depends entirely on language and is dominated by thought. Consequently, each has its own diction, grammar, and syntax. This is a conception of primary importance, although neither Humboldt nor those who encourage in developing his thought fully realised its implications. They distinguished only between poetry and prose, and within the latter between the exchange of ideas and ordinary conversation, i.e., the mere exchange of news or conventional chatter. There are other important functional distinctions in speech. One of them is the distinction between dialogue and monologue, as if written through the avenue of inner speech representation whereby it seems profoundly definitely strung by the monologue; The totalities of expression are uttered of some oral fashion as their linguistic manner as to be inferred by the spoken exchange that might be correlated by speech, in that in most cases, are contained through dialogue.

Dialogue always presupposes that in accordance with the collaborator’s formality that holds within the forming of knowledge, which it is maintained by its subject and is likely to be approved by an abbreviated speech and, under certain conditions, purely predicative sentences. It also presupposes that each person can see his partners, their facial expressions and gestures, and hear the tone of their voices. We have already discussed abbreviation and will consider here only its auditory aspect, using a classical example from Dostoevski’s, The Diary of a Writer, to show how much intonation helps the subtly differentiated understanding of a word’s meaning.

Dostoevski relates a conversation of drunks that entirely consisted of one unprintable word: ‘One Sunday night I happened to walk for some fifteen paces next to a group of six drunken young labourers, and I suddenly realised that all thoughts, feelings and even a whole chain of reasoning could be expressed by that one noun, which is moreover extremely short. One young fellow said it harshly and forcefully, to express his utter contempt for whatever it was they had all been talking about. Another answered with the same noun but in a quite different tone and sense - doubting that the negative attitude of the first one was warranted. A third suddenly became incensed against the first and roughly intruded on the conversation, excitedly shouting the same noun, this time as a curse and obscenity. Here the second fellow interfered again, angry with the third, the aggressor, and restraining him, in the sense of ‘Presently, as to implicate the now in question why to do you have to butt in, we were discussing things quietly and here you come and start swearing. And he told this whole thought in one word, the same venerable word, except that he also raised his hand and put it on the third fellow’s shoulder. All at once a fourth, the youngest of the group, who had kept silent till then, probably having suddenly found a solution to the original difficulty that had started the argument, raised his hand in a transport of joy and shouted . . . Eureka, do you think? I have it? No, not eureka and not I have it; he repeated the unprintable noun, one word, merely one word, but with ecstasy, in a shriek of delight - which was apparently too strong, because the sixth and the oldest, a glum-looking fellow, did not like it and cut the infantile joy of the other one short, addressing him in a sullen, exhortative bass and repeating . . . yes, still the same noun, forbidden in the presence of ladies but which this time clearly meant ‘What are you yelling yourself hoarse for? So, without uttering a single other word, they repeated that one beloved word is six times in a row, and only one after another, and understood one another completely.’ [The Diary of a Writer]

Inflection reveals the psychological context within which a word is to be understood. In Dostoevski’s story, it was contemptuous negation in one case, doubt in another, anger in the third. When the context is as clear as in this example, it really becomes possible to convey all thoughts, feelings, and even a whole chain of reasoning by one word.

In written speech, as tone of voice and knowledge of subject are excluded, we are obliged to use many more words, and to use them more exactly. Written speech is the most elaborate form of speech.

Some linguists consider dialogue the natural form of oral speech, the one in which language fully reveals its nature, and monologue to a greater degree for being artificial. Psychological investigation leaves no doubt that monologue is indeed the higher, more complicated form, and of later historical development. At present, however, we are interested in comparing them only in regard with the tendency toward abbreviation.

The speed of oral speech is unfavourable to a complicated process of formulation, but it does not leave time for deliberation and choice. Dialogue implies immediate unpremeditated utterance. It consists of replies, repartee; it is a chain of reactions. Monologue, by comparison, is a complex formation; the linguistic elaboration can be attended too leisurely and consciously.

In written speech, lacking situational and expressive supports, communication must be achieved only through words and their combinations; this requires the speech activity to take complicated forms - hence the use of first drafts. The evolution from the draft to the final copy reflects our mental process. Planning has an important part in written speech, even when we do not actually write out a draft. Usually we say to ourselves what we are going to write; This is also a draft, though in thought only. As we tried to show in the preceding chapter, this mental draft is inner speech. Since inner speech functions as a draft not only in written but also in oral speech, we will now compare both these forms with inner speech in respect to the tendency toward abbreviation and predication.

This tendency, never found in written speech and only some times in oral speech, arises in inner speech always. Predication is the natural form of inner speech, psychologically as it consists of predicates only. It is as much a law of inner speech to omit subjects as it is a law of written speech to contain both subjects and predicates.

The key to this experimentally established fact is the invariable, inevitable presence in inner speech of the factors that facilitate pure predication: We know what we are thinking about -, i.e., we always know the subject and the situation. Psychological contact between partners in a conversation may establish a mutual perception leading to the understanding of abbreviated speech. In inner speech, the ‘mutual’ perception is always there, in absolute form; Therefore, a practically wordless ‘communisation’ of even the most complicated thoughts is the rule. The predominance of predication is a product of development. In the beginning, egocentric speech is identical in structure with social speech, but in the process of its transformation into inner speech it gradually becomes less thorough and coherent as it becomes governed by the entire predicative syntax. Experiments show clearly how and why the new syntax takes hold. The child talks about the things he sees or hears or does at a given moment. As a result, he tends to leave out the subject and all words connected with it, condensing his speech frequently until only predicates are left. The more differentiated the specific function of egocentric speech becomes, the more pronounced are its syntactic peculiarities - simplification and predication. Hand in hand with this change goes decreasing vocalisation. When we converse with ourselves, we need even fewer words than Kitty and Levin did. Inner speech is speech almost without words.

With syntax and sound reduced to a minimum, meaning is more than ever in the forefront. Inner speech works with semantics, not phonetics. The specific semantic structure of inner speech also contributes to abbreviation. The syntax of meanings in inner speech is no less original than its grammatical syntax. Our investigation established three main semantic peculiarities of inner speech.

The first and basic one is the preponderance of the sense of a word over its meaning, and a distinction we accredit to Paulhan. The sense of a word, according to him, is the sum of all the psychological events aroused in our consciousness by the word. It is a dynamic, fluid, complex whole, which has several zones of unequal stability. Means is only one of the zones of sense, are the most stable and precise area. A word acquires its sense from the context in which it appears; in different contexts, it changes its sense. Meaning remains stable throughout the changes of sense. The dictionary meaning of a word is no more than a stone in the edifice of sense, no more than a potentiality that finds diversified realisation in speech.

The last words of the previously mentioned fable by Krylov, ‘The Dragonfly and the Ant,’ is a good illustration of the difference between sense and meaning. The words ‘Go and dances’ comprise of a definite and constant meaning, but in the context of the fable they acquire a much broader intellectual and affective sense. They mean both to ‘Enjoy yourself’ and ‘Perish.’ This enrichment of words by the sense they gain from the context is the fundamental law of the dynamics of word meanings. A frame in the circumstance of having to a context, it means both are more and fewer than the same word in isolation: More, because it acquires new content; less, because its meaning is limited and narrowed by the context. The sense of a word, says Paulhan, is a complex, mobile, protean phenomenon; it changes in different minds and situations and is almost unlimited. A word derives its sense from the sentence, which in turn gets its sense from the paragraph, the paragraph from the book, the book from all the works of the author.

Paulhan rendered a further service to psychology by analysing the relation between word and sense and showing that they are much more independent of each other than word and meaning. It has long been known that words can change their sense. Recently it was pointed out that sense can change words or, better, that ideas often change their names. Just as the sense of a word is connected with the whole word, and not with its single sounds, the sense of a sentence is connected with the whole sentence, and not with its individual words. Therefore, a word may sometimes be replaced by another without any change in sense. Words and sense are relatively independent of each other.

In inner speech, the predominance of sense over meaning, of the sentences over communicative words as their formalities and of context over sentences that are the rule.

This leads us to the other semantic peculiarities of inner speech. Both concern word combination. One of them is rather like agglutination, and a way of combining words fairly frequents in some languages and comparatively rare in others. German often forms one noun out of several words or phrases. In some primitive languages, such adhesion of words is a general rule. When several words are merged into one word, the new word not only expresses a rather complex idea but designates all the separate elements contained in that idea. Because the stress is always on the main root or idea, such languages are easy to understand. The egocentric speech of the child displays some analogous phenomena. As egocentric speech approaches inner speech, the child uses agglutination frequently as a way of forming compound words to express complex ideas.

The third basic semantic peculiarity of inner speech is the way in which senses of words combine and unite - a process governed by different laws from those governing combinations of meanings. When we observed this singular way of uniting words in egocentric speech, we called it ‘influx of sense.’ The senses of different words flow into one another - literally ‘influence’ one and another - so that the earlier ones are contained in, and modify, the later ones. Thus, a word that keeps recurring in a book or a poem sometimes absorbs all the variety of sense contained in it and becomes, in a way, equivalent to the work itself. The title of literary works expresses its content and completes its sense to a much greater degree than does the name of a painting or of a piece of music. Titles like Don Quixote, Hamlet, and Anna Karenina illustrate this very clearly - the whole sense of its operative word is contained in one name. Another excellent example is Gogol’s Dead Souls. Originally, the title referred to dead serfs whose names had not yet been removed from the official lists and who could still be bought and sold as if they were alive. It is in this sense that the words are used throughout the book, which is built up around this traffic in the dead. But through their intimate relationship with which the work as a whole, as these two words acquire the diversity of new and changing significance, an infinitely broader sense. When we reach the end of the book, ‘Dead Souls’ means to us not so much the defunct serfs as all the characters in the story, who are alive physically but dead spiritually.

In inner speech, the phenomenon reaches its peak. A single word is so saturated with sense that many words would be required to explain it in external speech. No wonder about why egocentric speech is incomprehensible to others. Watson says that inner speech would be incomprehensible even if it could be recorded. Its opaqueness is further increased by a related phenomenon that, incidentally, Tolstoy noted in external speech: In Childhood, Adolescence, and Youth, he describes how between people in close psychological contact words acquire special meanings understood only by the initiated. In inner speech, the same kind of idiom develops – the kind that is difficult to translate into the language of external speech.

With this we will conclude our survey of the peculiarities of inner speech, which we first observed in our investigation of egocentric speech. In looking for comparisons in external speech, we found that the latter already contain, potentially at least, the traits typical of inner speech; Predication, decreases the vocalisation, and preponderance of sense over meaning, agglutinations, etc., appear under certain conditions also in external speech. This, we believe, is the best confirmation of our hypothesis that inner speech originates through the differentiation of egocentric speech from the child’s primary social speech.

All our observations indicate that inner speech is an autonomous speech function. We can confidently regard it as a distinct plane of verbal thought. It is evident that the transition from inner to external speech is not a simple translation from one language into another. It cannot be achieved by merely vocalising silent speech. It is a complex, dynamic process involving the transformation of the predicative, idiomatic structure of inner speech into syntactically articulated speech intelligible to others.

We can now return to the definition of inner speech that we proposed before presenting our analysis. Inner speech is not the interior aspect of external speech, but it is a function in itself. It remains speech, i.e., thought connected with words. But while in external speech thought is embodied in words, in inner speech words die as they bring forth thought. Inner speech is to a large extent thinking in pure meanings. It is a dynamic, shifting, unstable thing, fluttering between word and thought, as two more or less sensible stables that are more or less firmly delineated components of verbal thought. Its true nature and place can be understood only after examining the next plane of verbal thought the one still more inward than inner speech.

That plane is thought itself. As we have said, every thought creates a connection, fulfils a function, solves a problem. The flow of thought is not accompanied by a simultaneous unfolding of speech. The two processes are not identical, and there is no rigid correspondence between the units of thought and speech. This is especially obvious when a thought process miscarries - when, as Dostoevski put it, a thought ‘will not enter words.’ Thought has its own structure, and the transition from it to speech is no easy matter. The theatre faced the problem of the thought behind the words before psychology did. In teaching his system of acting, Stanislavsky required the actors to uncover the ‘subtext’ of their lines in a play. In Griboedov’s comedy Woe from Wit, the hero, Chatsky, says to the hero, who maintains that she has never stopped thinking of him, ‘Thrice blessed who believes. Believing warms the heart.’ Stanislavsky interpreted this as ‘Let us stop this mutter’; However, to stop, it could just as well be interpreted as ‘I do not believe you. You say it to comfort me,’ or as ‘Don’t you see how you torment me? I wish I could believe you. That would be bliss.’ Every sentence that we say in real life has some kind of subtext, a thought hidden behind it. In the examples we gave earlier of the lack of coincidence between grammatical and psychological subject and predicate, we did not pursue our analysis to the end. Just as one sentence may express different thoughts, one thought may be expressed in different sentences. For instance, ‘The clock fell,’ in answer to the question ‘Why did the clock stop?’ Could mean? ‘It is not my fault that the clock is out of order; it fell.’ The same thought, for determining the self justification, could take the form of ‘It is not my habit to touch other people’s things. I was just dusting here,’ or a number of others.

Though, unlike speech, does not consist of separate units. When I wish to communicate the thought that today I saw a barefoot boy in a blue shirt running down the street, I do not see every item separately: the boy, the shirt, its blue colour, his running, the absence of shoes. I conceive of all this in one thought, but I put it into separate words. A speaker often takes several minutes to disclose one thought. In his mind the whole thought is present at once, but in speech it has to be developed successively. A thought may be compared with a cloud shedding a shower of words. Precisely because thought does not have its automatic counterpart in words, the transition from thought to word leads through meaning. In our speech, there is always the hidden thought, the subtext. Because a direct transition from thought to word is impossible, there have always been laments about the inexpressibility of thought: ‘How shall the heart express itself? How shall another understand?’

Direct communication between minds is impossible, not only physically but psychologically. Communication can be achieved only in a roundabout way. Thought must pass first through meanings and then through words.

We come now to the last step in our analysis of verbal thought. Though to be itself is too engendered by motivation, i.e., by our desires and needs, our interests and emotions. Behind every thought there is an affective-volitional tendency, which holds the answer to the last ‘why’ in the analysis of thinking. A true and full understanding of another’s thought is possible only when we understand its affective-volitional basis. We will illustrate this by an example already used: The interpretation of parts in a play. Stanislavsky, in his instructions to actors, listed the motives behind the words of their parts.

To understand another’s speech, it is not sufficient to understand his words, but we must understand his thought. But even that is not enough - we must also know its motivation. No psychological analysis of an utterance is complete until that plane is reached.

In the end, the verbal thought appeared as a complex, dynamic entity, and the relation of thought and word within it as a movement through a series of planes. Our analysis followed the process from the outermost to the innermost plane. In reality, the development of verbal thought takes the opposite course: From the motive that engenders a thought to the shaping of the thought, first in inner speech, then in meanings of words, and finally in words. It would be a mistake, however, to imagine that this is the only road from thought to word. The development may stop at any point in its complicated course; An infinite variety of movements back and forth, of ways still unknown to us, is possible. A study of these manifold variations lies beyond the scope of our present task.

Here we have wished to study the inner workings of thought and speech, hidden from direct observation. Meaning and the whole inward aspects of language, the position of which its turning toward the person, is not toward the outer world, have been so far an almost unknown territory. No matter how they were interpreted, the relations between thought and word were always considered constant, established forever. Our investigation has shown that they are, on the contrary, delicate, changeable relations between processes, which arise during the development of verbal thought. We did not intend to, and could not, exhaust the subject of verbal thought. We tried only to give a general conception of the infinite complexity of this dynamic structure - a conception starting from experimentally documented facts.

To association psychology, thought and its inscription of words was united by external bonds, similar to the bonds between two nonsense syllables. Gestalt psychology introduced the concept of structural bonds but, like the older theory, did not account for the specific relations between thought and word. All the other theories grouped themselves around two poles - either the behaviourist concept of thought as speech minus sound or the idealistic view, held by the Wuerzburg school and Bergson, that thought could be ‘pure,’ unrelated to language, and that it was distorted by words. Tjutchev’s ‘A thought once uttered is a lie’ could well serve as an epigraph for the latter group. Whether inclining toward pure naturalism or extreme idealism, all these theories have one trait in common - their antihistorical bias. They study thought and speech without any reference to their developmental history.

A historical theory of inner speech can deal with this immense and complex problem. The relation between thought and word is a living process; Thought is born through words. A word devoid of thought is a dead thing, and a thought unembodied in words remains a shadow. The connection between them, however, is not a preformed and constant one. It emerges in the course of development, and it evolves. To the Biblical ‘In the beginning was the Word,’ Goethe makes Faust reply, ‘In the beginning was the deed.’ The intent here is to detract from the value of the word, but we can accept this version if we emphasise it differently: In the beginning was the deed. The word was not the beginning, and action was there first; it is the end of development, crowning the deed.

We cannot, without mentioning the perspectives that our investigation opens. We studied the inward aspects of speech, which were as unknown to science as the other side of the moon. We showed that a generalised reflection of reality is the basic characteristic of words. This aspect of the word brings us to the threshold of a wider and deeper subject - the general problem of consciousness. Though and language, for which reflect reality in a way different from that of perception, that which is the key to the nature of human consciousness. Words play a central part not only in the development of thought but in the historical growth of consciousness as a whole. A word is a microcosm of human consciousness

The hermetic tradition has long been concerned with the relationship between the inner world of our consciousness and the outer world of nature, between the microcosm and the macrocosm, below and the above, the material and the spiritual, the centric and the peripheral. The hermetic world view held by such as Robert Fludd, having conceived by some great chain of being linking our inner spark of consciousness with all the facets of the Great World. There were grands to see the platonic metaphysical clockwork, as it were, through which our inner world was linked by means of a hierarchy of beings and planes to the highest unity of the Divine.

This view though comforting is philosophically unsound, and the developments in thought since the early 17th century have made such a hermetic world view seems as untenable and still philosophically naive. It is impossible to try to argue the case for such an hermetic metaphysic with anyone who has had philosophical training, for they will quickly and mercilessly reveal deep philosophical contradictions in this world view.

So do we now have to abandon such a beautiful and spiritual world view and adopt the prevailing reductionist materialist conception of the world that has become accepted in the intellectual tradition of the West?

I am not so sure. There still remains the problem of our consciousness and its relationship to our material form - the Mind / Brain problem. Behavioural psychologists such as Skinner tried to reduce this to one level - the material brain - by viewing the mental or consciousness events from the outside for being merely, stimulus-response loops. This simplistic view works well for basic reflex actions - ‘I itch therefore I scratch’ - but dissolves into absurdity when applied to any real act of the creative intellect or artistic imagination. Skinners’ determinism collapses when confronted with trying to explain the creative source of our consciousness revealing itself in an artist at work or a mathematician discovering through his thinking a new property of an abstract mathematical system. The psychologists' attempts to reduce the mind/brain problem to a merely material one of neurophysiology obviously failed. The idea that consciousness is merely a secretion or manifestation of a complex net of electrical impulses working within the mass of cells in our brain, is now discredited. The advocates of this view are strongly motivated by a desire to reduce the world to one level, to get rid of the necessity for ‘consciousness,’ ‘mind’ or ‘spirit’ as a real facet of the world.

This materialistic determinism in which everything in the world (including the phenomenon of consciousness) can be reduced to simple interactions on a physical/chemical level, belongs really to the nineteenth century scientific landscape. Nineteenth century science was founded upon a ‘Newtonian Absolute Physics’ which provided a description of the world as an interplay of forces obeying immutable laws and following a predetermined pattern. This is the ‘billiard ball’ view of the world - one in which, provided we are given the initial state of the system (the layout of the balls on the table, and the exact trajectory, momentum and other parameters of the cue ball, etc.) then theoretically the exact layout after each interaction can be precisely calculated to absolute precision. All could be reduced to the determinate interplay of matter obeying the immutable laws of physics. The concept of the ‘spiritual’ was unnecessary, even ‘mind’ was dispensable, and ‘God’ of course had no place in this scheme of things.

This comfortably solid ‘Newtonian’ world view of the materialists has however been entirely undermined by the new physics of the twentieth century, and in particular through Quantum Theory. Physicists investigating the properties of sub-atomic matter, found that the deterministic Newtonian absolutism broke down at the foundation level of matter. An element of probability had to be introduced into the physicists' calculations, and each sub-atomic event was itself inherently unpredictably - one could only ascribe a probability to the outcome. The simple billiard ball model collapsed at the sub-atomic level. For if the billiard table was intended as a picture of a small region of space on the atomic scale and each ball was to be a particle (an electron, proton, or neutron, etc.), then physicists came to realise that this model could not represent reality on that level. For in Quantum theory one could not define the position and momentum of a particle both at the same moment. As soon as we establish the parameters of motion of a body, its position is uncertain and can only be described mathematically as a wave of probability. Our billiard table dissolved into a fluid ever-moving undulating surface, with each ball at one moment focussed to a point then at another dissolving and spreading itself out over an area of the space of the table. Trying to play billiards at this sub-atomic level was rather difficult.

In the Quantum picture of the world, each individual event cannot be determined exactly, but has to be described by a wave of probability. There is a kind of polarity between the position and energy of any particle in which they cannot be simultaneously determined. This was not a failing of experimental method but a property of the kinds of mathematical structures that physicists have to use to describe this realm of the world. The famous equation of Quantum theory embodying Heisenberg's Uncertainty Principle is: Planck's constant = (uncertainty in energy) x (uncertainty in position)

Thus if we try to fix the position of the particle (i.e., reduce the uncertainty in its position to a small factor) then as a consequence of this equation the uncertainty in the energy must increase to balance this, and therefore we cannot find a value for the energy of the particle simultaneous with fixing its position. Planck's constant being very small means that these infractions as based of the factors only become dominant on the extremely small scale, which are within the realm of the atom.

So we see that the Quantum picture of reality has at its foundation a non-deterministic view of the fundamental building Forms of matter. Of course, when dealing with large masses of particles these quantum indeterminacies effectively cancel each other out, and physicists can determine and predict the state of large systems. Obviously planets, suns, galaxies being composed of large numbers of particles do not exhibit any uncertainty in their position and energies, for when we look at such large aggregates as some of its totality, the total quantum uncertainty is a systems reduction as placed by zero, and in respect to their large scale properties can effectively be treated as deterministic systems.

Thus on the large scale we can effectively apply a deterministic physics, but when we wish to look in detail at the properties of the sub-atomic realm, lying at the root and foundation of our world, we must enter a domain of quantum uncertainties and find the neat ordered picture dissolving into a sea of ever flowing forces that we cannot tie down or set into fixed patterns.

Some people when faced with this picture of reality find comfort in dismissing the quantum world as having little to do with the ‘real world’ of appearances. We do not live within the sub-atomic level after all. However, it does spill out into our outer world. Most of the various electronic devices of the past decades rely on the quantum tunnelling effect in transistors and silicon chips. The revolution in quantum physics has begun to influence the life sciences, and biologists and botanists are beginning to come up against quantum events as the basis of living systems, in the structure of complex molecules in the living tissues and membranes of cells for example. When we look at the blue of the sky, we are looking at a phenomenon only recently understood through quantum theory.

Although the Quantum picture of reality might seem strange indeed, I believe the picture it presents of the foundations of the material world, the ever flowing sea of forces metamorphosing and interacting through the medium of ‘virtual’ or quantum messenger particles, has certain parallels with nature of our consciousness.

No comments:

Post a Comment