You are on page 1of 430
EX Pc eA T CON oN a CONSCIOUSNESS THEHARDPROBLEM JONATHAN SHEAR Explaining Consciousness — The ‘Hard Problem’ Edited by Jonathan Shear ABradford Book ‘The MIT Press Cambridge, Massachusetts London, England Copyright © 1995-7 by the Journal of Consciousness Studies Imprint Academic, PO Box 1, Thorverton EXS SYX, UK. All rights reserved. No part of this book may be reproduced in any form by any electronic or mechanical means (including photocopying, recording, or information storage and retrieval) without permission in writing from the copyright holder. ‘This book was set in Times New Roman by Imprint Academic. Printed and bound in the United States of America. Library of Congress Cataloging-in-Publication-Data Explaining consciousness: the “hard problem" / edited by Jonathan Shear. pcm. “ABradtford book.” Includes bibliographical references. ISBN-13 978-0-262-19388-7 (he, : alk. paper) — 978-0-262-69221-2 (pbk. : alk. paper) ISBN-10 0-262-19388-4 (he. : alk, paper) — 0-262-69221-X (pbk. : alk. paper) 1. Consciousness. 1. Shear, J. Gonathan) B808.9.E85 1997 126—de21 97-8302 cP 1098765 Contents Introduction Jonathan Shear The Hard Problem Facing Up to the Problem of Consciousness David J. Chalmers Deflationary Perspectives Facing Backwards on the Problem of Consciousness Daniel C. Dennett ‘The Hornswoggle Problem Patricia Smith Churchland Function and Phenomenology: Closing the Explanatory Gap Thomas W. Clark ‘The Why of Consciousness: A Non-Issue for Materialists Valerie Gray Hardcastle There Is No Hard Problem of Consciousness Kieron O'Hara and Tom Scutt Should We Expect to Feel as if We Understand Consciousness? Mark C. Price The Explanatory Gap Consciousness and Space Colin McGinn Giving Up on the Hard Problem of Consciousness Eugene O. Mills 33 37 61 7 vi CONTENTS There Are No Easy Problems of Consciousness EJ. Lowe ‘The Easy Problems Ain't So Easy David Hodgson Facing Ourselves: Incorrigibility and the Mind-Body Problem Richard Warner ‘The Hardness of the Hard Problem William S. Robinson Physics The Nonlocality of Mind CJS. Clarke Conscious Events as Orchestrated Space-Time Selections Stuart R. Hameroff and Roger Penrose The Hard Problem: A Quantum Approach Henry P Stapp Physics, Machines and the Hard Problem Douglas J. Bilodeau Neuroscience and Cognitive Science Why Neuroscience May Be Able to Explain Consciousness Francis Crick and Christof Koch Understanding Subjectivity: Global Workspace Theory and the Resurrection of the Observing Self Bernard J. Baars ‘The Elements of Consciousness and Their Neurodynamical Correlates Bruce MacLennan Rethinking Nature Consciousness, Information and Panpsychism Wiltiam Seager Rethinking Nature: A Hard Problem within the Hard Problem Gregg H. Rosenberg Solutions to the Hard Problem of Consciousness Benjamin Libet Turning ‘the Hard Problem’ Upside Down and Sideways Piet Hut and Roger N. Shepard 117 125 133, 149 237 241 249 269 287 301 305 CONTENTS First-Person Perspectives The Relation of Consciousness to the Material World Max Velmans Neurophenomenology: A Methodological Remedy for the Hard Problem Francisco J. Varela ‘The Hard Problem: Closing the Empirical Gap Jonathan Shear Response Moving Forward on the Problem of Consciousness David J. Chalmers 325 337 359 379 Jonathan Shear Introduction In the view of modem science, the universe is fundamentally physical and existed and evolved for billions of years without any consciousness present in it a all. Consciousness, on this now commonplace view, is a very recent addition, a byproduct of the complex chemical processes that gave rise to sophisticated biological organisms. The question thus arises of how consciousness ever came to be.' For it is clear that science describes the world purely in terms of spatial distributions of dynamic potentials to occupy and influence regions of space, without any reference whatsoever to the colours, sounds, feelings, anticipations, meanings, and thoughts with which our consciousness is so wonderfully filled. Moreover, the experiential qualities of our subjective awareness are so different in kind from the colourless, soundless, non-conscious spatio-temporal struc- tures described by the physical sciences that it seems impossible to understand how such experiential qualities could ever be produced by the interactions of purely physical systems. This radical difference between the qualia (experiential qualities) of our sub- jective awareness and the qualia-free nature of non-conscious physical systems thus produces what David Chalmers calls, in the keynote article of this volume, the “hard problem’ of explaining how any physical system, no matter how complex and well-organized, (could) give rise to experience at all ... Why is it that all this processing does not go on ‘in the dark’, without any subjective quality? .. . This is the phenomenon that makes consciousness a real mystery. The underlying premise of the ‘hard problem’, that the universe can exist and unfold independently of the existence of, and interaction with, consciousness of any sort, while appearing commonsensical in today’s scientific world, was of course not always taken for granted. The first paragraph of Genesis, for example, proclaims that it was God’s ‘spirit’ that enlivened the formless ‘depths’ to unfold into the created universe, and this notion is echoed in the subsequent theological doctrines that insist God’s active partici- pation is required for the continued existence of the universe. Outside the realm of faith, the pre-Socratics, Anaxagoras and Heraclitus, were famous for their claims that mind is the foundation of the activities of the natural world and ‘steers all things through all things’, Plato argued that the very intelligibility of the universe presupposes the existence ofa pure transcendental intelligence underlying the structural and dynamical order of the universe. And even the empirically-minded Aristotle argued that purpose permeates the inanimate as well as the animate universe, and that a transcendental intelligence, the Unmoved Mover, underlies all the activity in nature. Although there were also of course 1 ‘This is to be distinguished from the question of why consciousness arose, that is, what evolution- ary purpose it serves. rf . J, SHEAR materialists such as Democritus, the Greek founders of Western thought, operating in a pre-scientific context, were not much occupied with the ‘hard problem’ of accounting for consciousness in a fundamentally physical universe that seems so difficult to us today. Indeed, as Plato’s dialogues repeatedly show, the existence of mental entities unexplain- able in physical terms were likely simply to be taken as evidence for the independent existence of mind, rather than as perplexing anomalies in a materialistic world. With the rise of the materialistic-mechanical viewpoint characteristic of modern sci- ence, all of this changed, of course. But this viewpoint gained dominance only after a protracted struggle with the Aristotelian and neo-Platonic descendents of the ancient Greek views, as the notion of what we today call ‘science’ evolved during the sixteenth century. Each of these worldviews had a very different basic paradigm for understanding nature, The Aristotelian view, which had dominated during the middle ages, relied on the notion of organism as its basic explanatory image. The neo-Platonic view took mind, emphasizing mathematics and creative activity, as basic. The materialist-mechanistic view relied on the notion of machines unfolding automatically and unconsciously. The materialist-mechanistic view won out, of course. Before this happened, however, empiri- cal science was forced to internalize a major component of the neo-Platonic perspective, newly reintroduced to European thought in the late fifteenth century. From our own vantage point, nothing could seem more natural to the empirical scientific approach than its extensive reliance on mathematics for formulating the laws of nature and evaluating their experimental status. But, surprising as it may seem today, fifteenth and sixteenth century scientific investigators at first often resisted the mathematizing approach empha- sized by the neo-Platonists. For numbers, algebra and mathematical formulas are obvi- ously not physical entities; they are purely mental and not objectively observable. Consequently early observation-oriented investigators tended to reject their use as an escape from the real world, more appropriate to neo-Platonic mysticism and metaphysics than to hard-boiled empirical science. But the success of mathematics, based on its usefulness both in the formulation of theory and in the articulation of precisely testable predictions, was so great that it became seamlessly integrated into the evolving, ulti mately mechanistic scientific paradigm, and the early objections to its inclusion in empirical science on the grounds of its mentalistic nature became largely forgotten. Indeed, by the mid-seventeenth century we find mathematics even came to be seen as defining the objective physical world in contrast to the generally non-mathematizable subjective world of consciousness, as formalized, for example, in Descartes’ influential mind-body distinction, ‘Thus the modern scientific worldview, regarding the universe as material in nature and unfolding like a machine according to precise, mathematically articulatable laws gained ascendency and became the context of much of our Western intellectual discourse. Nevertheless, the largely forgotten early puzzle about the role of the mathematics in the physical sciences remained and has continued to draw the attention of influential physicists.? For mathematics appears to be paradigmatically non-physical. Its objects are 2 For an account of the interactions between the three approaches in the early development of modem science, see Hugh Kearney, Science and Change: 1500-1700 (New York: McGraw-Hill Book Company, 1971). 3 See, for example, Eugene Wigner, “The unreasonable effectiveness of mathematics in the natural Sciences’, in Symmetries and Reflections (Bloomington: Indiana University Press, 1967). INTRODUCTION 3 characteristically universal in nature, in sharp contrast to the complete particularity of all physical entities and empirical observations; itis discovered within the realm of thought, not in the physical world; and the truth of its theorems are almost always evaluated completely mentally, rather than by objective scientific investigation.* Thus as physics becomes progressively more mathematized, its objects often seem puzzlingly to become more mental than physical. This is not, of course, the only problem with the standard materialist worldview that has come to the attention of philosophers of science. For it is obvious that insofar as it is observation-dependent, science is also to a degree dependent on the subjectivity of the observing scientific investigators. Consequently, philosophers of science came up with the notions of intersubjectivity and intersubjective corroboration to recognize the role of subjectivity in empirical science, and over the last half century this terminology has found widespread currency in scientific as well as philosophical discourse. It has also become apparent that our observations in general, and our scientific observations in particular, are characteristically (and even, it is often argued, always) theory impregnated. This has often been taken (especially by writers associated with ‘postmodern’ thought) to imply that the notion of ‘objectivity independent of subjectivity’ is simply incoherent. To the extent that this observation is correct, the notion of a purely physical object or universe, free from any admixture of mind, also becomes problematic. ‘Advances in physics, the paradigmatically materialist discipline, have sometimes seemed to reintroduce the notion of consciousness as well. The most extreme example is that of quantum mechanics, where the role of the ‘observer’ paradoxically seemed to become essential not only for knowing the discrete phenomena of nature, but even for their very existence as discrete phenomena. For on many well-known interpretations of ‘quantum mechanics it has seemed that it is the act of observation itself which somehow causes the ‘collapse’ of the wave function (describing the probabilities of a particle manifesting in any particular location throughout the universe) and the consequent localized existence of the particle in question. These interpretations are of course radi- cally counter-intuitive, calling the materialist paradigm directly into question. The prob- ems involved here are clearly illustrated by Schridinger’s famous ‘cat in the box’ paradox, Schrédinger invites us to consider a physical system consisting of a cat in a box; until someone observes it, the equations of quantum mechanics imply that this system can be in a variety of states — including those in which the cat is dead and others in which it is alive — but that it is not in fact in any one of these states until it is actually observed. Consequently, it is incorrect to think that the cat is either alive or dead until an observation that collapses the quantum probabilities into a particular state (in which the cat is either alive or dead) takes place. Thus for the time (a day or a week, etc.) during which the cat remains unobserved it is neither dead nor alive, despite the fact that once the system is relevantly observed one will be able to say definitively that it had been one or the other during this time. Thus, problematically, the state (including even the location) of objects in the universe appears to be significantly undetermined until a conscious act of observation takes place. And even after this act takes place it remains * Computer assisted evaluations of proofs too lengthy to be calculated by hand are arguably a partial exception to this general rule. 4 J, SHEAR paradoxical, for one is compelled to hold that even though the cat was neither dead nor alive prior to being observed, after the observation takes place one can be sure that exactly one of these two alternatives had been the case, namely, that it had been alive (if it was alive on the subsequent observation) or that it had not. Clearly this logical dilemma, too, poses a deep problem for our commonsensical materialistic-mechanistic conception of objective reality. It also demonstrates significant inadequacies in our notion of consciousness. For an obvious rejoinder to the ‘cat in the box’ paradox is that the cat should ‘know’ if it were alive, For the cat, when conscious, would often be observing its environment, and the specificity of its perceptions would imply that the wave functions had already collapsed prior to our own (human) observations. But could a cat's consciousness cause the collapse of the wave function? Would its observations be relevant to the formalisms of quantum mechanics? Would they provide the relevant ‘information’ or ‘knowledge’? If not, why not? And if a cat’s awareness could do this, what about that of an amoeba? For amoebas clearly display information-driven behavior. Or is an amoeba not conscious at all, or not relevantly information-responsive? Again, if not, why not? Are there different types of consciousness, and only some of these relevant to quantum-mechanical collapse? If so, what are they, what are their distinctive differences, and how do these differences relate to the dynamics of quantum mechanical processes? In short, the ‘cat in the box’ paradox, in addition to highlighting the puzzle of observation-driven quantum mechani- cal ‘collapse’ and the paradox of states simultaneously existing-and-not-existing, also suggests that such paradigmatically mental phenomena as ‘consciousness’, ‘knowledge’ and ‘information’ need to be analysed with much finer discrimination before their roles in — and overall relation to — physical systems can be properly understood. Thus, after three centuries of dominance, the standard materialist paradigm has been forced by advances in quantum mechanics to wrestle once again with what appears to be a major anomaly: the nature and role of consciousness in physical systems. The emer- gence of the ‘hard problem’ of consciousness into the forefront of philosophical discourse can be understood as part of this reexamination process. How, in a basically material universe, are we to understand even the bare existence of consciousness? How could it ever have emerged? We should remind ourselves, however, that alternatives to the materialist perspective have of course been developed and defended throughout the history of modern Western philosophy. Descartes’ dualism split the universe into two distinct, independent types of ‘substance’, comprising, respectively, the mathematically-describable objective physi- cal-spatial realm studied by science,’ and the subjective spiritual realm of mind or consciousness, properly the domain of religion.® Spinoza split the universe into the same two realms, but argued monistically that they were both attributes of a further, underlying substance, Leibnitz developed a ‘panpsychic’ theory of pre-spatio-temporal monads in which every thing, no matter how merely physical and nonconscious it might seem, 5 Descartes’ work in this area was, of course, a major factor in the development of the now standard materialistic-mechanistic paradigm. ® Indeed, the very fact that the ‘hard problem’ seems so hard today is often held to be the legacy of Descartes’ dualism, and the problem itself is often formulated in terms of the difficulty of reconnect- ing causally and ontologically what Descartes so forcefully divided. INTRODUCTION 5 actually had at least some faint degree of sentience and perception, And Hegel developed an idealistic theory of the universe evolving dialectically out of ‘Spirit’ or consciousness itself. From the perspective of each of these theories, of course, the ‘hard problem’, if articulatable at all, would certainly not be the central philosophical anomaly it is today. For since none of them take matter as basic, none need be constrained to attempt to account for consciousness as arising from matter. But, for a variety of reasons — not the least being their emphasis on metaphysical entities quite outside the realm of scientific enquiry and corroboration — these views have all dropped by the wayside in our modern scientific world, and the materialist view has come to be so dominant that it is generally regarded as simple common sense. In this context the hard problem of explaining consciousness has come to be recognized as a very hard problem indeed. The materialist view, however, has often resolved what had seemed to be insurmount- able obstacles. The physical sciences have repeatedly been successful in incorporating realms and processes once thought to be necessarily outside their scope. Newton's demonstration that the same gravitational and mechanical laws apply to both celestial and terrestrial phenomena, and contemporary explanations of biological processes in terms of DNA, RNA and molecular chemistry are notable examples. Such past successes have Jed many influential defenders of the standard paradigm to insist that, whatever the conceptual difficulties, consciousness and its phenomenal content will ultimately turn out to be explicable as and reducible to purely physical phenomena, even if we cannot now foresee how this will happen. However, to others the gap between the qualia-free material world of the physical sciences and the qualia-filled contents of phenomenal awareness at present remains so great that this insistence appears to be little more than an act of faith.” ‘The enduring difficulty of the ‘hard problem’ has accordingly prompted a number of otherwise hard-headed, scientifically-oriented thinkers to go outside the standard context of discourse in the search for unconventional solutions. The articles of the present volume, which originally appeared in a multi-part special ‘hard problem’ issue of the Journal of Consciousness Studies (SCS),® reflect this diversity. Indeed, the editors of JCS were surprised at how many of the papers submitted to the special issue developed theories containing distinctly non-standard suggestions, ranging from discussions of the foundations of quantum-mechanics to panexperientialist theses in which all of nature, no matter how seemingly non-sentient, actually has a phenomenal-experiential component. There were, of course, also quite a few papers that defended the standard paradigm and attempted to explain the ‘hard problem’ away as being not such a hard problem after all. The present book consists of a keynote paper by David Chalmers, ‘Facing Up to the Problem of Consciousness’, followed by twenty-six papers, twenty-four of which were written in response to the keynote, and Chalmers’ reply to the responses. The responses have been grouped into six categories. The categories are, to be sure, somewhat arbitrary, for a number of the papers could easily fit in more than one. Nevertheless, some jon seemed to be in order, especially to help readers who were not yet conver- sant with the diversity of serious approaches to the topic. 7 For a useful overview see Given Giizeldere ‘Consciousness: what it is, how to study it, what to learn from its history’, JCS, 2 (1), 1995, pp. 30-51; and ‘Problems of consciousnes perspective ‘on contemporary issues, current debates’, JCS, 2 (2), 1995, pp. 112-44, 8 ‘The one exception is the short commentary by Crick and Koch from Scientific American. 6 J. SHEAR ‘The keynote paper begins by defining the ‘hard problem’ as the problem of accounting for the existence of phenomenal experience, and discusses why this problem is so hard. Chalmers differentiates this problem from the relatively ‘easy’ problems of under- standing the structural and functional properties of cognitive processes, arguing that these are at least in principle amenable to conventional scientific theories and methodologies, while the former problem does not appear to be. He then suggests that to resolve the ‘hard problem’ we need to postulate experience as a fundamental feature of nature alongside the more conventional features such as space-time, mass, and charge. On this view, a final theory of consciousness will consist of a set of fundamental ‘psychophysical’ laws accompanying the fundamental laws of physics. Chalmers then offers a preliminary outline of such a theory, centred on a ‘double-aspect’ (phenomenological and physical) view of information. ‘The papers in ‘Deflationary Perspectives’ argue that the problem of phenomenal consciousness is nowhere near as hard as it might seem; they argue, for example, that it is a pseudo-problem, that materialist reduction will dissipate the problem, and that it is simply premature, in our present ignorance of how consciousness and the brain work, to talk about any uniquely ‘hard’ problem. Those in “The Explanatory Gap’ argue, on the other hand, that the problem is very hard; several analyze further aspects of the problem that Chalmers didn’t contemplate, and some take the position that the problem will necessarily remain unresolvable. The papers of the next four sections, however, argue that it ought to be resolvable, and offer a variety of possible resolutions. Those in ‘Physics’ base their suggestions on a reexamination of relationships between conscious- ness and the objective world in the light of contemporary physics, especially quantum mechanics. The papers in ‘Neuroscience and Cognitive Science’ contain suggestions, ranging from conservative to highly speculative, about how consciousness and the nervous system may be related. Those in ‘Rethinking Nature’ explore implications of the radical hypothesis that consciousness ought to be understood as a pervasive feature of the ‘objective universe. Finally, the papers in ‘First-Person Perspectives’ argue that resolving the hard problem will require development of a new science of the subjective phenomena of consciousness, complementing our existing studies of the objective material universe. ‘The articles in this volume are thus quite diverse, and contain a variety of non-standard components, as well as those from within the usual scientific paradigm. This sort of creative diversity is of course what should be expected as we wrestle with what has come to be recognized as a serious challenge for standard materialism, namely the existence of consciousness itself. Acknowledgments 1 would like to thank a number of people for their help in preparing this book: Keith Sutherland, publisher of Journal of Consciousness Studies who had the original idea for the book, Anthony Freeman, JCS managing editor; David Chalmers, for his keynote and response articles and his ongoing encouragement; Jean Burns, another editor of JCS, for her help with the scientific submissions; Gené Mills of Virginia Commonwealth Univer- sity, for his feedback on many of the early papers; the other referees, too numerous to cite here, for their careful analyses; my sweet wife Patricia, for more than I can express; and, of course, all the contributing authors. The Hard Problem David J. Chalmers Facing Up to the Problem of Consciousness 1: Introduction* Consciousness poses the most baffling problems in the science of the mind. There is nothing that we know more intimately than conscious experience, but there is nothing that is harder to explain. All sorts of mental phenomena have yielded to scientific investigation in recent years, but consciousness has stubbornly resisted. Many have tried to explain it, but the explanations always seem to fall short of the target. Some have been led to suppose that the problem is intractable, and that no good explanation can be given. To make progress on the problem of consciousness, we have to confront it directly. In this paper, I first isolate the truly hard part of the problem, separating it from more tractable parts and giving an account of why it is so difficult to explain. I critique some recent work that uses reductive methods to address consciousness, and argue that these methods inevitably fail to come to grips with the hardest part of the problem. Once this failure is recognized, the door to further progress is opened. In the second half of the paper, I argue that if we move to a new kind of nonreductive explanation, a naturalistic account of consciousness can be given. I put forward my own candidate for such an account: a nonreductive theory based on principles of structural coherence and organiza- tional invariance and a double-aspect view of information. Il: The Easy Problems and the Hard Problem ‘There is not just one problem of consciousness, ‘Consciousness’ is an ambiguous term, referring to many different phenomena. Each of these phenomena needs to be explained, but some are easier to explain than others. At the star, it is useful to divide the associated problems of consciousness into ‘hard’ and ‘easy’ problems. The easy problems of consciousness are those that seem directly susceptible to the standard methods of cogni- tive science, whereby a phenomenon is explained in terms of computational or neural mechanisms. The hard problems are those that seem to resist those methods. The easy problems of consciousness include those of explaining the following phenomena: * This paper was originally published in the Journal of Consciousness Studies, 2, No.3 (1995), pp. 200-19. The arguments are presented in much greater depth in my book The Conscious Mind (Chalmers, 1996). Thanks to Francis Crick, Peggy DesAutels, Matthew Elton, Liane Gabora, Christof Koch, Paul Rhodes, Gregg Rosenberg, and Sharon Wahl for helpful comments. 10 DJ. CHALMERS the ability to discriminate, categorize, and react to environmental stimuli; the integration of information by a cognitive system; the reportability of mental states; the ability of a system to access its own internal states; the focus of attention; the deliberate control of behaviour; the difference between wakefulness and sleep. All of these phenomena are associated with the notion of consciousness. For example, one sometimes says that a mental state is conscious when it is verbally reportable, or when it is internally accessible. Sometimes a system is said to be conscious of some information when it has the ability to react on the basis of that information, or, more strongly, when it attends to that information, or when it can integrate that information and exploit it in the sophisticated control of behaviour. We sometimes say that an action is conscious precisely when it is deliberate. Often, we say that an organism is conscious as. another way of saying that it is awake. ‘There is no real issue about whether these phenomena can be explained scientifically. All of them are straightforwardly vulnerable to explanation in terms of computational or neural mechanisms. To explain access and reportability, for example, we need only specify the mechanism by which information about internal states is retrieved and made available for verbal report. To explain the integration of information, we need only exhibit mechanisms by which information is brought together and exploited by later processes. For an account of sleep and wakefulness, an appropriate neurophysiological account of the processes responsible for organisms’ contrasting behaviour in those states will suffice. In each case, an appropriate cognitive or neurophysiological model can clearly do the explanatory work. If these phenomena were all there was to consciousness, then consciousness would not be much of a problem. Although we do not yet have anything close to a complete explanation of these phenomena, we have a clear idea of how we might go about explaining them. This is why I call these problems the easy problems. Of course, ‘easy’ is a relative term. Getting the details right will probably take a century or two of difficult empirical work. Still, there is every reason to believe that the methods of cognitive science and neuroscience will succeed. The really hard problem of consciousness is the problem of experience. When we think and perceive, there is a whir of information-processing, but there is also a subjective aspect. As Nagel (1974) has put it, there is something itis like to be a conscious organism. This subjective aspect is experience. When we see, for example, we experience visual sensations: the felt quality of redness, the experience of dark and light, the quality of depth in a visual field, Other experiences go along with perception in different modalities: the sound of a clarinet, the smell of mothballs. Then there are bodily sensations, from pains to orgasms; mental images that are conjured up internally; the felt quality of emotion, and the experience of a stream of conscious thought. What unites all of these states is that there is something itis like to be in them, All of them are states of experience. It is undeniable that some organisms are subjects of experience. But the question of how it is that these systems are subjects of experience is perplexing. Why is it that when FACING UP TO THE PROBLEM OF CONSCIOUSNESS it our cognitive systems engage in visual and auditory information-processing, we have visual or auditory experience: the quality of deep blue, the sensation of middle C? How ‘can we explain why there is something it is like to entertain a mental image, or to experience an emotion? It is widely agreed that experience arises from a physical basis, but we have no good explanation of why and how it so arises. Why should physical processing give rise to a rich inner life at all? It seems objectively unreasonable that it should, and yet it does. If any problem qualifies as the problem of consciousness, it is this one. In this central sense of ‘consciousness’, an organism is conscious if there is something it is like to be that organism, and a mental state is conscious if there is something it is like to be in that state. Sometimes terms such as ‘phenomenal consciousness’ and ‘qualia’ are also used here, but I find it more natural to speak of ‘conscious experience’ or simply ‘experience’. Another useful way to avoid confusion (used by e.g. Newell 1990, Chalmers 1996) is to reserve the term ‘consciousness’ for the phenomena of experience, using the less loaded term ‘awareness’ for the more straightforward phenomena described earlier. If such a convention were widely adopted, communication would be much easier. As things stand, those who talk about ‘consciousness’ are frequently talking past each other. ‘The ambiguity of the term ‘consciousness’ is often exploited by both philosophers and scientists writing on the subject. It is common to see a paper on consciousness begin with an invocation of the mystery of consciousness, noting the strange intangibility and ineffability of subjectivity, and worrying that so far we have no theory of the phenome- non. Here, the topic is clearly the hard problem — the problem of experience. In the second half of the paper, the tone becomes more optimistic, and the author's own theory of consciousness is outlined. Upon examination, this theory tums out to be a theory of one of the more straightforward phenomena — of reportability, of introspective access, or whatever. At the close, the author declares that consciousness has turned out to be tractable after all, but the reader is left feeling like the victim of a bait-and-switch. The hard problem remains untouched. III: Functional Explanation Why are the easy problems easy, and why is the hard problem hard? The easy problems are easy precisely because they concern the explanation of cognitive abilities and ‘functions, To explain a cognitive function, we need only specify a mechanism that can perform the function. The methods of cognitive science are well-suited for this sort of explanation, and so are well-suited to the easy problems of consciousness. By contrast, the hard problem is hard precisely because it is not a problem about the performance of functions. The problem persists even when the performance of all the relevant functions is explained.! To explain reportability, for instance, is just to explain how a system could perform the function of producing reports on internal states. To explain internal access, we need to explain how a system could be appropriately affected by its internal states and use information about those states in directing later processes, To explain integration and | Here ‘function’ is not used in the narrow teleological sense of something that a system is designed to do, but in the broader sense of any causal role in the production of behaviour that a system might perform. 12 DJ, CHALMERS control, we need to explain how a system’s central processes can bring information contents together and use them in the facilitation of various behaviours, These are all problems about the explanation of functions. How do we explain the performance of a function? By specifying a mechanism that performs the function. Here, neurophysiological and cognitive modelling are perfect for the task. If we want a detailed low-level explanation, we can specify the neural mecha- nism that is responsible for the function. If we want a more abstract explanation, we can specify a mechanism in computational terms. Bither way, a full and satisfying explana- tion will result. Once we have specified the neural or computational mechanism that performs the function of verbal report, for example, the bulk of our work in explaining reportability is over. Ina way, the point is trivial. It is a conceptual fact about these phenomena that their explanation only involves the explanation of various functions, as the phenomena are ‘functionally definable. All it means for reportability to be instantiated in a system is that the system has the capacity for verbal reports of internal information. All it means for a system to be awake is for it to be appropriately receptive to information from the environment and for it to be able to use this information in directing behaviour in an appropriate way. To see that this sort of thing is a conceptual fact, note that someone who says ‘you have explained the performance of the verbal report function, but you have not explained reportability’ is making a trivial conceptual mistake about reportability. All it could possibly take to explain reportability is an explanation of how the relevant function is performed; the same goes for the other phenomena in question. ‘Throughout the higher-level sciences, reductive explanation works in just this way. To explain the gene, for instance, we needed to specify the mechanism that stores and transmits hereditary information from one generation to the next. It turns out that DNA performs this function; once we explain how the function is performed, we have ex- plained the gene. To explain life, we ultimately need to explain how a system can reproduce, adapt to its environment, metabolize, and so on. All of these are questions about the performance of functions, and so are well-suited to reductive explanation. The same holds for most problems in cognitive science. To explain learning, we need to explain the way in which a system’s behavioural capacities are modified in light of environmental information, and the way in which new information can be brought to bear in adapting a system’s actions to its environment. If we show how a neural or computa- tional mechanism does the job, we have explained learning. We can say the same for other cognitive phenomena, such as perception, memory, and language, Sometimes the rele- vant functions need to be characterized quite subtly, but itis clear that insofar as cognitive science explains these phenomena at all, it does so by explaining the performance of functions. ‘When it comes to conscious experience, this sort of explanation fails. What makes the hard problem hard and almost unique is that it goes beyond problems about the perform- ance of functions. To see this, note that even when we have explained the performance of all the cognitive and behavioural functions in the vicinity of experience — perceptual discrimination, categorization, internal access, verbal report — there may still remain a further unanswered question: Why is the performance of these functions accompanied by experience? A simple explanation of the functions leaves this question open. FACING UP TO THE PROBLEM OF CONSCIOUSNESS 13 There is no analogous further question in the explanation of genes, or of life, or of learning. If someone says ‘I can see that you have explained how DNA stores and transmits hereditary information from one generation to the next, but you have not explained how it is a gene,” then they are making a conceptual mistake. All it means to be a gene is to be an entity that performs the relevant storage and transmission function. But if someone says ‘I can see that you have explained how information is discriminated, integrated, and reported, but you have not explained how it is experienced,’ they are not making a conceptual mistake. This is a nontrivial further question. This further question is the key question in the problem of consciousness. Why doesn’t all this information-processing go on ‘in the dark’, free of any inner feel? Why is it that when electromagnetic waveforms impinge on a retina and are discriminated and catego- rized by a visual system, this discrimination and categorization is experienced as a sensation of vivid red? We know that conscious experience does arise when these functions are performed, but the very fact that it arises is the central mystery. There is an explanatory gap (a term due to Levine 1983) between the functions and experience, and we need an explanatory bridge to cross it. A mere account of the functions stays on one side of the gap, so the materials for the bridge must be found elsewhere. This is not to say that experience has no function. Perhaps it will turn out to play an important cognitive role. But for any role it might play, there will be more to the explanation of experience than a simple explanation of the function. Perhaps it will even turn out that in the course of explaining a function, we will be led to the key insight that allows an explanation of experience. If this happens, though, the discovery will be an extra explanatory reward. There is no cognitive function such that we can say in advance that explanation of that function will automatically explain experience. To explain experience, we need a new approach. The usual explanatory methods of cognitive science and neuroscience do not suffice. These methods have been developed precisely to explain the performance of cognitive functions, and they do a good job of it. But as these methods stand, they are only equipped to explain the performance of functions. When it comes to the hard problem, the standard approach has nothing to say. IV: Some Case-Studies In the last few years, a number of works have addressed the problems of consciousness within the framework of cognitive science and neuroscience. This might suggest that the analysis above is faulty, but in fact a close examination of the relevant work only lends the analysis further support. When we investigate just which aspects of consciousness these studies are aimed at, and which aspects they end up explaining, we find that the ultimate target of explanation is always one of the easy problems. I will illustrate this with two representative examples. ‘The first is the ‘neurobiological theory of consciousness’ outlined by Francis Crick and Christof Koch (1990; see also Crick 1994). This theory centers on certain 3575 hertz neural oscillations in the cerebral cortex; Crick and Koch hypothesize that these oscilla- tions are the basis of consciousness. This is partly because the oscillations seem to be correlated with awareness in a number of different modalities — within the visual and olfactory systems, for example — and also because they suggest a mechanism by which 14 DJ. CHALMERS the binding of information contents might be achieved, Binding is the process whereby separately represented pieces of information about a single entity are brought together to be used by later processing, as when information about the colour and shape of a perceived object is integrated from separate visual pathways, Following others (e.g. Eckhomn et al. 1988), Crick and Koch hypothesize that binding may be achieved by the synchronized oscillations of neuronal groups representing the relevant contents. When two pieces of information are to be bound together, the relevant neural groups will oscillate with the same frequency and phase, The details of how this binding might be achieved are still poorly understood, but suppose that they can be worked out. What might the resulting theory explain? Clearly it might explain the binding of information contents, and perhaps it might yield a more general account of the integration of information in the brain. Crick and Koch also suggest that these oscillations activate the mechanisms of working memory, so that there may be an account of this and perhaps other forms of memory in the distance. The theory might eventually lead to a general account of how perceived information is bound and stored in memory, for use by later processing. Such a theory would be valuable, but it would tell us nothing about why the relevant contents are experienced. Crick and Koch suggest that these oscillations are the neural correlates of experience. This claim is arguable — does not binding also take place in the processing of unconscious information? — but even if it is accepted, the explanatory question remains: Why do the oscillations give rise to experience? The only basis for an explanatory connection is the role they play in binding and storage, but the question of why binding and storage should themselves be accompanied by experience is never addressed. If we do not know why binding and storage should give rise to experience, telling a story about the oscillations cannot help us. Conversely, if we knew why binding and storage gave rise to experience, the neurophysiological details would be just the icing on the cake. Crick and Koch’s theory gains its purchase by assuming a connection between binding and experience, and so can do nothing to explain that link. I do not think that Crick and Koch are ultimately claiming to address the hard problem, although some have interpreted them otherwise. A published interview with Koch gives a clear statement of the limitations on the theory’s ambitions. Well, let's first forget about the really difficult aspects, like subjective feelings, for they may not have a scientific solution. The subjective state of play, of pain, of pleasure, of seeing blue, of smelling a rose — there seems to be a huge jump between the materialistic level, of explaining molecules and neurons, and the subjective level. Let’s focus on things that are easier to study — like visual awareness. You're now talking to me, but you're not looking at me, you're looking at the cappuccino, and so you are aware of it. You can say, ‘It's a cup and there’s some liquid in it. If I give it to you, you'll move your arm and you'll take it — you'll respond in a meaningful manner. That's what I call awareness. (‘What is Consciousness?’, Discover, November 1992, p. 96.) ‘The second example is an approach at the level of cognitive psychology. This is Bernard Baars’ global workspace theory of consciousness, presented in his book A Cognitive Theory of Consciousness (1988). According to this theory, the contents of consciousness FACING UP TO THE PROBLEM OF CONSCIOUSNESS 15 are contained in a global workspace, a central processor used to mediate communication between a host of specialized nonconscious processors. When these specialized proces- sors need to broadcast information to the rest of the system, they do so by sending this information to the workspace, which acts as a kind of communal blackboard for the rest of the system, accessible to all the other processors. Baars uses this model to address many aspects of human cognition, and to explain a number of contrasts between conscious and unconscious cognitive functioning. Ulti- mately, however, itis a theory of cognitive accessibility, explaining how it is that certain information contents are widely accessible within a system, as well as a theory of informational integration and reportability. The theory shows promise as a theory of awareness, the functional correlate of conscious experience, but an explanation of expe- rience itself is not on offer. ‘One might suppose that according to this theory, the contents of experience are precisely the contents of the workspace, But even if this is so, nothing internal to the theory explains why the information within the global workspace is experienced. The best the theory can do is to say that the information is experienced because it is globally accessible. But now the question arises in a different form: why should global accessibil- ity give rise to conscious experience? As always, this bridging question is unanswered. ‘Almost all work taking a cognitive or neuroscientific approach to consciousness in recent years could be subjected to a similar critique. The ‘Neural Darwinism’ model of Edelman (1989), for instance, addresses questions about perceptual awareness and the self-concept, but says nothing about why there should also be experience, The ‘multiple drafts’ model of Dennett (1991) is largely directed at explaining the reportability of certain mental contents. The ‘intermediate level’ theory of Jackendoff (1987) provides an account of some computational processes that underlie consciousness, but Jackendoff stresses that the question of how these ‘project’ into conscious experience remains mysterious. Researchers using these methods are often inexplicit about their attitudes to the problem of conscious experience, although sometimes they take a clear stand. Even among those who are clear about it, attitudes differ widely. In placing this sort of work with respect to the problem of experience, a number of different strategies are available. It would be useful if these strategic choices were more often made explicit. ‘The first strategy is simply to explain something else. Some researchers are explicit that the problem of experience is too difficult for now, and perhaps even outside the domain of science altogether. These researchers instead choose to address one of the more tractable problems such as reportability or the self-concept. Although I have called these problems the ‘easy’ problems, they are among the most interesting unsolved problems in cognitive science, so this work is certainly worthwhile. The worst that can be said of this choice is that in the context of research on consciousness itis relatively unambitious, and the work can sometimes be misinterpreted. ‘The second choice is to take a harder line and deny the phenomenon. (Variations on this approach are taken by Allport 1988; Dennett 1991; Wilkes 1988.) According to this, Tine, once we have explained the functions such as accessibility, reportability, and the like, there is no further phenomenon called ‘experience’ to explain. Some explicitly deny the phenomenon, holding for example that what is not externally verifiable cannot be real. 16 D.J. CHALMERS Others achieve the same effect by allowing that experience exists, but only if we equate ‘experience’ with something like the capacity to discriminate and report. These ap- proaches lead to a simpler theory, but are ultimately unsatisfactory. Experience is the most central and manifest aspect of our mental lives, and indeed is perhaps the key explanandum in the science of the mind. Because of this status as an explanandum, experience cannot be discarded like the vital spirit when a new theory comes along. Rather, itis the central fact that any theory of consciousness must explain. A theory that denies the phenomenon ‘solves’ the problem by ducking the question. Ina third option, some researchers claim to be explaining experience in the full sense, ‘These researchers (unlike those above) wish to take experience very seriously; they lay out their functional model or theory, and claim that it explains the full subjective quality of experience (e.g. Flohr 1992; Humphrey 1992). The relevant step in the explanation is usually passed over quickly, however, and usually ends up looking something like magic. After some details about information processing are given, experience suddenly enters the picture, but it is left obscure how these processes should suddenly give rise to experience. Perhaps it is simply taken for granted that it does, but then we have an incomplete explanation and a version of the fifth strategy below. ‘A fourth, more promising approach appeals to these methods to explain the structure of experience. For example, itis arguable that an account of the discriminations made by the visual system can account for the structural relations between different colour experiences, as well as for the geometric structure of the visual field (see e.g. Clark 1992; Hardin 1992). In general, certain facts about structures found in processing will corre- spond to and arguably explain facts about the structure of experience, This strategy is plausible but limited, At best, it takes the existence of experience for granted and accounts for some facts about its structure, providing a sort of nonreductive explanation of the structural aspects of experience (I will say more on this later). This is useful for many purposes, but it tells us nothing about why there should be experience in the first place. A fifth and reasonable strategy is to isolate the substrate of experience. After all, almost everyone allows that experience arises one way or another from brain processes, and it makes sense to identify the sort of process from which it arises, Crick and Koch put their work forward as isolating the neural correlate of consciousness, for example, and Edel- man (1989) and Jackendoff (1987) make related claims. Justification of these claims requires a careful theoretical analysis, especially as experience is not directly observable in experimental contexts, but when applied judiciously this strategy can shed indirect light on the problem of experience. Nevertheless, the strategy is clearly incomplete. For a satisfactory theory, we need to know more than which processes give rise to experience; we need an account of why and how. A full theory of consciousness must build an explanatory bridge. V: The Extra Ingredient ‘We have seen that there are systematic reasons why the usual methods of cognitive science and neuroscience fail to account for conscious experience. These are simply the wrong sort of methods: nothing that they give to us can yield an explanation. To account FACING UP TO THE PROBLEM OF CONSCIOUSNESS 17 for conscious experience, we need an extra ingredient in the explanation. This makes for a challenge to those who are serious about the hard problem of consciousness: What is your extra ingredient, and why should that account for conscious experience? ‘There is no shortage of extra ingredients to be had. Some propose an injection of chaos and nonlinear dynamics. Some think that the key lies in nonalgorithmic processing. Some appeal to future discoveries in neurophysiology. Some suppose that the key to the mystery will lie at the level of quantum mechanics. It is easy to see why all these suggestions are put forward. None of the old methods work, so the solution must lie with something new. Unfortunately, these suggestions all suffer from the same old problems. Nonalgorithmic processing, for example, is put forward by Penrose (1989; 1994) because of the role it might play in the process of conscious mathematical insight. The arguments about mathematics are controversial, but even if they succeed and an account of nonalgorithmic processing in the human brain is given, it will still only be an account of the functions involved in mathematical reasoning and the like. For a nonalgorithmic process as much as an algorithmic process, the question is left unanswered: why should this process give rise to experience? In answering this question, there is no special role for nonalgorithmic processing. The same goes for nonlinear and chaotic dynamics. These might provide a novel account of the dynamics of cognitive functioning, quite different from that given by standard methods in cognitive science. But from dynamics, one only gets more dynamics. ‘The question about experience here is as mysterious as ever. The point is even clearer for new discoveries in neurophysiology. These new discoveries may help us make significant progress in understanding brain function, but for any neural process we isolate, the same question will always arise. It is difficult to imagine what a proponent of new neurophysi- ology expects to happen, over and above the explanation of further cognitive functions. Its not as if we will suddenly discover a phenomenal glow inside a neuron! Perhaps the most popular ‘extra ingredient’ of all is quantum mechanics (e.g. Hameroff 1994), The attractiveness of quantum theories of consciousness may stem from a Law of Minimization of Mystery: consciousness is mysterious and quantum mechanics is mys- terious, so maybe the two mysteries have a common source, Nevertheless, quantum theories of consciousness suffer from the same difficulties as neural or computational theories, Quantum phenomena have some remarkable functional properties, such as nondeterminism and nonlocality. It is natural to speculate that these properties may play some role in the explanation of cognitive functions, such as random choice and the integration of information, and this hypothesis cannot be ruled out a priori. But when it comes to the explanation of experience, quantum processes are in the same boat as any other. The question of why these processes should give rise to experience is entirely unanswered.” 2 One special attraction of quantum theories is the fact that on some interpretations of quantum. mechanics, consciousness plays an active role in ‘collapsing’ the quantum wave function, Such interpretations are controversial, but in any case they offer no hope of explaining consciousness in terms of quantum processes. Rather, these theories assume the existence of consciousness, and use it in the explanation of quantum processes. At best, these theories tell us something about a physical role that consciousness may play. They tell us nothing about how it arises. 18 D.J. CHALMERS At the end of the day, the same criticism applies to any purely physical account of consciousness. For any physical process we specify there will be an unanswered question: Why should this process give rise to experience? Given any such process, it is conceptu- ally coherent that it could be instantiated in the absence of experience, It follows that no mere account of the physical process will tell us why experience arises. The emergence of experience goes beyond what can be derived from physical theory. Purely physical explanation is well-suited to the explanation of physical structures, explaining macroscopic structures in terms of detailed microstructural constituents; and it provides a satisfying explanation of the performance of functions, accounting for these functions in terms of the physical mechanisms that perform them. This is because a physical account can entail the facts about structures and functions: once the internal details of the physical account are given, the structural and functional properties fall out as an automatic consequence. But the structure and dynamics of physical processes yield only more structure and dynamics, so structures and functions are all we can expect these processes to explain. The facts about experience cannot be an automatic consequence of any physical account, as it is conceptually coherent that any given process could exist without experience. Experience may arise from the physical, but it is not entailed by the physical. The moral of all this is that you can’t explain conscious experience on the cheap. It is. a remarkable fact that reductive methods — methods that explain a high-level phenome- non wholly in terms of more basic physical processes — work well in so many domains. In a sense, one can explain most biological and cognitive phenomena on the cheap, in that these phenomena are seen as automatic consequences of more fundamental proc- esses. It would be wonderful if reductive methods could explain experience, too; [hoped for a long time that they might. Unfortunately, there are systematic reasons why these methods must fail. Reductive methods are successful in most domains because what needs explaining in those domains are structures and functions, and these are the kind of thing that a physical account can entail. When it comes to a problem over and above the explanation of structures and functions, these methods are impotent, This might seem reminiscent of the vitalist claim that no physical account could explain life, but the cases are disanalogous. What drove vitalist scepticism was doubt about whether physical mechanisms could perform the many remarkable functions associated with life, such as complex adaptive behaviour and reproduction. The concep- tual claim that explanation of functions is what is needed was implicitly accepted, but lacking detailed knowledge of biochemical mechanisms, vitalists doubted whether any physical process could do the job and put forward the hypothesis of the vital spirit as an alternative explanation. Once it turned out that physical processes could perform the relevant functions, vitalist doubts melted away. With experience, on the other hand, physical explanation of the functions is not in question. The key is instead the conceptual point that the explanation of functions does not suffice for the explanation of experience. This basic conceptual point is not something that further neuroscientific investigation will affect. In a similar way, experience is disanalogous to the elan vital. The vital spirit was put forward as an explanatory posit, in order to explain the relevant functions, and could therefore be discarded when those functions were explained without it. Experience is not an FACING UP TO THE PROBLEM OF CONSCIOUSNESS 19 explanatory posit but an explanandum in its own right, and so is not a candidate for this sort of elimination. tis tempting to note that all sorts of puzzling phenomena have eventually turned out to be explainable in physical terms. But each of these were problems about the observable behaviour of physical objects, coming down to problems in the explanation of structures and functions, Because of this, these phenomena have always been the kind of thing that a physical account might explain, even if at some points there have been good reasons to suspect that no such explanation would be forthcoming. The tempting induction from these cases fails in the case of consciousness, which is not a problem about physical structures and functions. The problem of consciousness is puzzling in an entirely differ- ent way. An analysis of the problem shows us that conscious experience is just not the kind of thing that a wholly reductive account could succeed in explaining. VI: Nonreductive Explanation At this point some are tempted to give up, holding that we will never have a theory of conscious experience. McGinn (1989), for example, argues that the problem is too hard for our limited minds; we are ‘cognitively closed’ with respect to the phenomenon. Others have argued that conscious experience lies outside the domain of scientific theory altogether. I think this pessimism is premature. This is not the place to give up; itis the place where things get interesting, When simple methods of explanation are ruled out, we need to investigate the alternatives. Given that reductive explanation fails, nonreductive explana- tion is the natural choice. Although a remarkable number of phenomena have turned out to be explicable wholly in terms of entities simpler than themselves, this is not universal. In physics, it occasion- ally happens that an entity has to be taken as fundamental. Fundamental entities are not explained in terms of anything simpler. Instead, one takes them as basic, and gives a theory of how they relate to everything else in the world, For example, in the nineteenth century it turned out that electromagnetic processes could not be explained in terms of the wholly mechanical processes that previous physical theories appealed to, so Maxwell and others introduced electromagnetic charge and electromagnetic forces as new funda- ‘mental components of a physical theory. To explain electromagnetism, the ontology of physics had to be expanded. New basic properties and basic laws were needed to give a satisfactory account of the phenomena. Other features that physical theory takes as fundamental include mass and space-time. No attempt is made to explain these features in terms of anything simpler. But this does not rule out the possibility of a theory of mass or of space-time. There is an intricate theory of how these features interrelate, and of the basic laws they enter into. These basic principles are used to explain many familiar phenomena concerning mass, space, and time at a higher level. T suggest that a theory of consciousness should take experience as fundamental. We know that a theory of consciousness requires the addition of something fundamental to our ontology, as everything in physical theory is compatible with the absence of con- sciousness. We might add some entirely new nonphysical feature, from which experience 20 DJ. CHALMERS can be derived, but it is hard to see what such a feature would be like, More likely, we will take experience itself as a fundamental feature of the world, alongside mass, charge, and space-time. If we take experience as fundamental, then we can go about the business of constructing a theory of experience. ‘Where there is a fundamental property, there are fundamental laws. A nonreductive theory of experience will add new principles to the furniture of the basic laws of nature. These basic principles will ultimately carry the explanatory burden in a theory of consciousness. Just as we explain familiar high-level phenomena involving mass in terms of more basic principles involving mass and other entities, we might explain familiar phenomena involving experience in terms of more basic principles involving experience and other entities. In particular, a nonreductive theory of experience will specify basic principles telling us how experience depends on physical features of the world. These psychophysical principles will not interfere with physical laws, as it seems that physical laws already form a closed system. Rather, they will be a supplement to a physical theory. A physical theory gives a theory of physical processes, and a psychophysical theory tells us how those processes give rise to experience. We know that experience depends on physical processes, but we also know that this dependence cannot be derived from physical laws alone. The new basic principles postulated by a nonreductive theory give us the extra ingredient that we need to build an explanatory bridge. Of course, by taking experience as fundamental, ther sense in which this approach does not tell us why there is experience in the first place. But this is the same for any fundamental theory. Nothing in physics tells us why there is matter in the first place, but we do not count this against theories of matter. Certain features of the world need to be taken as fundamental by any scientific theory. A theory of matter can still explain all sorts of facts about matter, by showing how they are consequences of the basic laws. The same goes for a theory of experience. This position qualifies as a variety of dualism, as it postulates basic properties over and above the properties invoked by physics. But itis an innocent version of dualism, entirely compatible with the scientific view of the world, Nothing in this approach contradicts anything in physical theory; we simply need to add further bridging principles to explain how experience arises from physical processes. There is nothing particularly spiritual or mystical about this theory — its overall shape is like that of a physical theory, with a few fundamental entities connected by fundamental laws. It expands the ontology slightly, to ‘be sure, but Maxwell did the same thing. Indeed, the overall structure of this position is entirely naturalistic, allowing that ultimately the universe comes down to a network of basic entities obeying simple laws, and allowing that there may ultimately be a theory of consciousness cast in terms of such laws. If the position is to have a name, a good choice might be naturalistic dualism. If this view is right, then in some ways a theory of consciousness will have more in common with a theory in physics than a theory in biology. Biological theories involve no principles that are fundamental in this way, so biological theory has a certain complexity and messiness to it; but theories in physics, insofar as they deal with fundamental principles, aspire to simplicity and elegance. The fundamental laws of nature are part of the basic furniture of the world, and physical theories are telling us that this basic FACING UP TO THE PROBLEM OF CONSCIOUSNESS 21 furniture is remarkably simple. If a theory of consciousness also involves fundamental principles, then we should expect the same. The principles of simplicity, elegance, and even beauty that drive physicists’ search for a fundamental theory will also apply to a theory of consciousness. VII: Toward of a Theory of Consciousness It is not too soon to begin work on a theory. We are already in a position to understand some key facts about the relationship between physical processes and experience, and about the regularities that connect them. Once reductive explanation is set aside, we can lay those facts on the table so that they can play their proper role as the initial pieces in a nonreductive theory of consciousness, and as constraints on the basic laws that constitute an ultimate theory. ‘There is an obvious problem that plagues the development of a theory of conscious- ness, and that is the paucity of objective data. Conscious experience is not directly observable in an experimental context, so we cannot generate data about the relationship between physical processes and experience at will. Nevertheless, we all have access to a rich source of data in our own case. Many important regularities between experience and processing can be inferred from considerations about one’s own experience. There are also good indirect sources of data from observable cases, as when one relies on the verbal report of a subject as an indication of experience. These methods have their limitations, but we have more than enough data to get a theory off the ground. Philosophical analysis is also useful in getting value for money out of the data we have. This sort of analysis can yield a number of principles relating consciousness and cogni- tion, thereby strongly constraining the shape of an ultimate theory. The method of thought-experimentation can also yield significant rewards, as we will see. Finally, the fact that we are searching for a fundamental theory means that we can appeal to such nonempirical constraints as simplicity, homogeneity, and the like in developing a theory. ‘We must seek to systematize the information we have, to extend it as far as possible by careful analysis, and then make the inference to the simplest possible theory that explains the data while remaining a plausible candidate to be part of the fundamental furniture of the world. Such theories will always retain an element of speculation that is not present in other scientific theories, because of the impossibility of conclusive intersubjective experimen- tal tests, Still, we can certainly construct theories that are compatible with the data that 3 Some philosophers argue that even though there is a conceptual gap between physical processes and experience, there need be no metaphysical gap, so that experience might in a certain sense stillbe physical (e.g. Hill 1991; Levine 1983; Loar 1990). Usually this line of argument is supported byan ‘appeal to the notion of @ posteriori necessity (Kripke 1980). I think that this position rests on a misunderstanding of a posteriori necessity, however, or else requires an entirely new sort of necessity that we have no reason to believe in; see Chalmers 1996 (also Jackson 1994; Lewis 1994) for details. In any case, this position still concedes an explanatory gap between physical processes ‘and experience. For example, the principles connecting the physical and the experiential will not be derivable from the laws of physics, so such principles must be taken as explanatorily fundamental. So even on this sort of view, the explanatory structure of a theory of consciousness will be much as Thave described, 22 DJ. CHALMERS we have, and evaluate them in comparison to each other. Even in the absence of intersubjective observation, there are numerous criteria available for the evaluation of such theories: simplicity, internal coherence, coherence with theories in other domains, the ability to reproduce the properties of experience that are familiar from our own case, and even an overall fit with the dictates of common sense, Perhaps there will be significant indeterminacies remaining even when all these constraints are applied, but we can at least develop plausible candidates. Only when candidate theories have been developed will we be able to evaluate them. ‘A nonreductive theory of consciousness will consist of a number of psychophysical principles, principles connecting the properties of physical processes to the properties of experience. We can think of these principles as encapsulating the way in which experi- ence arises from the physical. Ultimately, these principles should tell us what sort of physical systems will have associated experiences, and for the systems that do, they should tell us what sort of physical properties are relevant to the emergence of experience, and just what sort of experience we should expect any given physical system to yield. This is a tall order, but there is no reason why we should not get started. In what follows, I present my own candidates for the psychophysical principles that might go into a theory of consciousness. The first two of these are nonbasic principles — systematic connections between processing and experience at a relatively high level. These principles can play a significant role in developing and constraining a theory of consciousness, but they are not cast at a sufficiently fundamental level to qualify as truly basic laws. The final principle is a candidate for a basic principle that might form the comerstone of a fundamental theory of consciousness. This principle is particularly speculative, but it is the kind of speculation that is required if we are ever to have a satisfying theory of consciousness. I can present these principles only briefly here; I argue for them at much greater length in Chalmers 1996. I. The principle of structural coherence This is a principle of coherence between the structure of consciousness and the structure of awareness. Recall that ‘awareness’ was used earlier to refer to the various functional phenomena that are associated with consciousness. I am now using it to refer to a somewhat more specific process in the cognitive underpinnings of experience. In particu- lar, the contents of awareness are to be understood as those information contents that are accessible to central systems, and brought to bear in a widespread way in the control of behaviour. Briefly put, we can think of awareness as direct availability for global control. To a first approximation, the contents of awareness are the contents that are directly accessible and potentially reportable, at least in a language-using system. Awareness is a purely functional notion, but it is nevertheless intimately linked to conscious experience. In familiar cases, wherever we find consciousness, we find aware- ness, Wherever there is conscious experience, there is some corresponding information in the cognitive system that is available in the control of behaviour, and available for verbal report. Conversely, it seems that whenever information is available for report and for global control, there is a corresponding conscious experience. Thus, there is a direct correspondence between consciousness and awareness. FACING UP TO THE PROBLEM OF CONSCIOUSNESS 23 ‘The correspondence can be taken further. It is a central fact about experience that it has a complex structure. The visual field has a complex geometry, for instance, There are also relations of similarity and difference between experiences, and relations in such things as relative intensity. Every subject’s experience can be at least partly characterized and decomposed in terms of these structural properties: similarity and difference relations, perceived location, relative intensity, geometric structure, and so on. It is also a central fact that to each of these structural features, there is a corresponding feature in the information-processing structure of awareness. Take colour sensations as an example. For every distinction between colour experi- ences, there is a corresponding distinction in processing. The different phenomenal colours that we experience form # complex three-dimensional space, varying in hue, saturation, and intensity. The properties of this space can be recovered from information- processing considerations: examination of the visual systems shows that waveforms of light are discriminated and analysed along three different axes, and it is this three-dimen- sional information that is relevant to later processing. The three-dimensional structure of phenomenal colour space therefore corresponds directly to the three dimensional struc- ture of visual awareness. This is precisely what we would expect. After all, every colour distinction corresponds to some reportable information, and therefore to a distinction that is represented in the structure of processing. In a more straightforward way, the geometric structure of the visual field is directly reflected in a structure that can be recovered from visual processing. Every geometric relation corresponds to something that can be reported and is therefore cognitively represented. If we were given only the story about information-processing in an agent’s visual and cognitive system, we could not directly observe that agent's visual experi- ences, but we could nevertheless infer those experiences’ structural properties. In general, any information that is consciously experienced will also be cognitively represented. The fine-grained structure of the visual field will correspond to some fine-grained structure in visual processing. The same goes for experiences in other modalities, and even for nonsensory experiences. Internal mental images have geometric properties that are represented in processing. Even emotions have structural properties, such as relative intensity, that correspond directly to a structural property of processing; where there is greater intensity, we find a greater effect on later processes. In general, precisely because the structural properties of experience are accessible and reportable, those properties will be directly represented in the structure of awareness. It is this isomorphism between the structures of consciousness and awareness that constitutes the principle of structural coherence. This principle reflects the central fact that even though cognitive processes do not conceptually entail facts about conscious experience, consciousness and cognition do not float free of one another but cohere in an intimate way. This principle has its limits. It allows us to recover structural properties of experience from information-processing properties, but not all properties of experience are structural properties, There are properties of experience, such as the intrinsic nature of a sensation of red, that cannot be fully captured in a structural description, The very intelligibility of inverted spectrum scenarios, where experiences of red and green are inverted but all structural properties remain the same, show that structural properties constrain experi- 24 D.J. CHALMERS ence without exhausting it. Nevertheless, the very fact that we feel compelled to leave structural properties unaltered when we imagine experiences inverted between function- ally identical systems shows how central the principle of structural coherence is to our conception of our mental lives. It is not a logically necessary principle, as after all we can imagine all the information processing occurring without any experience at all, but it is nevertheless a strong and familiar constraint on the psychophysical connection, ‘The principle of structural coherence allows for a very useful kind of indirect explana- tion of experience in terms of physical processes. For example, we can use facts about neural processing of visual information to indirectly explain the structure of colour space. ‘The facts about neural processing can entail and explain the structure of awareness; if we take the coherence principle for granted, the structure of experience will also be ex- plained. Empirical investigation might even lead us to better understand the structure of awareness within animals, shedding indirect light on Nagel’s vexing question of what it is like to be a bat. This principle provides a natural interpretation of much existing work on the explanation of consciousness (e.g. Clark 1992, Hardin 1992 on colours; Akins 1993 on bats), although it is often appealed to inexplicitly. It is so familiar that itis taken for granted by almost everybody, and is a central plank in the cognitive explanation of consciousness. The coherence between consciousness and awareness also allows a natural interpreta- tion of work in neuroscience directed at isolating the substrate (or the neural correlate) of consciousness. Various specific hypotheses have been put forward. For example, Crick and Koch (1990) suggest that 40-hertz oscillations may be the neural correlate of consciousness, whereas Libet (1993) suggests that temporally-extended neural activity is central. If we accept the principle of coherence, the most direct physical correlate of consciousness is awareness: the process whereby information is made directly available for global control. The different specific hypotheses can be interpreted as empirical suggestions about how awareness might be achieved. For example, Crick and Koch suggest that 40-Hz oscillations are the gateway by which information is integrated into working memory and thereby made available to later processes. Similarly, itis natural to suppose that Libet’s temporally extended activity is relevant precisely because only that sort of activity achieves global availability. The same applies to other suggested corre- lates such as the ‘global workspace’ of Baars (1988), the ‘high-quality representations’ of Farah (1994), and the ‘selector inputs to action systems’ of Shallice (1972). All these can be seen as hypotheses about the mechanisms of awareness: the mechanisms that perform the function of making information directly available for global control. Given the coherence between consciousness and awareness, it follows that a mecha- nism of awareness will itself be a correlate of conscious experience. The question of just which mechanisms in the brain govern global availability is an empirical one; perhaps there are many such mechanisms. But if we accept the coherence principle, we have reason to believe that the processes that explain awareness will at the same time be part of the basis of consciousness. FACING UP TO THE PROBLEM OF CONSCIOUSNESS 25 2. The principle of organizational invariance This principle states that any two systems with the same fine-grained functional organi- zation will have qualitatively identical experiences. If the causal patterns of neural organization were duplicated in silicon, for example, with a silicon chip for every neuron and the same patterns of interaction, then the same experiences would arise, According to this principle, what matters for the emergence of experience is not the specific physical makeup of a system, but the abstract pattern of causal interaction between its components. This principle is controversial, of course, Some (e.g. Searle 1980) have thought that consciousness is tied to a specific biology, so that a silicon isomorph of a human need not be conscious. I believe that the principle can be given significant support by the analysis of thought-experiments, however. Very briefly: suppose (for the purposes of a reductio ad absurdum) that the principle is false, and that there could be two functionally isomorphic systems with different experiences, Perhaps only one of the systems is conscious, or perhaps both are conscious but they have different experiences. For the purposes of illustration, let us say that one system is made of neurons and the other of silicon, and that one experiences red where the other experiences blue. The two systems have the same organization, so we can imagine gradually transforming one into the other, perhaps replacing neurons one at a time by silicon chips with the same local function. We thus gain a spectrum of inter- mediate cases, each with the same organization, but with slightly different physical makeup and slightly different experiences. Along this spectrum, there must be two systems A and B between which we replace less than one tenth of the system, but whose experiences differ. These two systems are physically identical, except that a small neural circuit in A has been replaced by a silicon circuit in B. The key step in the thought-experiment is to take the relevant neural circuit in A, and install alongside it a causally isomorphic silicon circuit, with a switch between the two. ‘What happens when we flip the switch? By hypothesis, the system's conscious experi- ences will change; from red to blue, say, for the purposes of illustration. This follows from the fact that the system after the change is essentially a version of B, whereas before the change it is just A. But given the assumptions, there is no way for the system to notice the changes! Its causal organization stays constant, so that all of its functional states and behavioural dispositions stay fixed. As far as the system is concerned, nothing unusual has happened. There is no room for the thought, ‘Hmm! Something strange just happened!" In general, the structure of any such thought must be reflected in processing, but the structure of processing remains constant here. If there were to be such a thought it must float entirely free of the system and would be utterly impotent to affect later processing. (If it affected later processing, the systems would be functionally distinct, contrary to hypothesis.) We might even flip the switch a number of times, so that experiences of red and blue dance back and forth before the system's ‘inner eye’. According to hypothesis, the system can never notice these ‘dancing qualia’. This I take to be a reductio of the original assumption, It is a central fact about experience, very familiar from our own case, that whenever experiences change signifi- cantly and we are paying attention, we can notice the change; if this were not to be the 26 DJ. CHALMERS case, we would be led to the sceptical possibility that our experiences are dancing before cour eyes all the time. This hypothesis has the same status as the possibility that the world was created five minutes ago: perhaps it is logically coherent, but it is not plausible. Given the extremely plausible assumption that changes in experience correspond to changes in processing, we are led to the conclusion that the original hypothesis is impossible, and that any two functionally isomorphic systems must have the same sort of experiences. To put it in technical terms, the philosophical hypotheses of ‘absent qualia’ and ‘inverted qualia’, while logically possible, are empirically and nomologically impos- sible* ‘There is more to be said here, but this gives the basic flavour. Once again, this thought experiment draws on familiar facts about the coherence between consciousness and cognitive processing to yield a strong conclusion about the relation between physical structure and experience. If the argument goes through, we know that the only physical properties directly relevant to the emergence of experience are organizational properties. This acts as a further strong constraint on a theory of consciousness. 3. The double-aspect theory of information The two preceding principles have been nonbasic principles. They involve high-level notions such as ‘awareness’ and ‘organization’, and therefore lie at the wrong level to constitute the fundamental laws in a theory of consciousness. Nevertheless, they act as strong constraints. What is further needed are basic principles that fit these constraints and that might ultimately explain them. ‘The basic principle that I suggest centrally involves the notion of information. I understand information in more or less the sense of Shannon (1948). Where there is information, there are information states embedded in an information space. An informa- tion space has a basic structure of difference relations between its elements, charac- terizing the ways in which different elements in a space are similar or different, possibly in complex ways. An information space is an abstract object, but following Shannon we can sec information as physically embodied when there is a space of distinct physical states, the differences between which can be transmitted down some causal pathway. The states that are transmitted can be seen as themselves constituting an information space. To borrow a phrase from Bateson (1972), physical information is a difference that makes a difference. The double-aspect principle stems from the observation that there is a direct isomor- phism between certain physically embodied information spaces and certain phenomenal (or experiential) information spaces. From the same sort of observations that went into the principle of structural coherence, we can note that the differences between phenome- nal states have a structure that corresponds directly to the differences embedded in physical processes; in particular, to those differences that make a difference down certain causal pathways implicated in global availability and control. That is, we can find the * Some may worry that a silicon isomorph of a neural system might be impossible for technical reasons, That question is open. The invariance principle says only that if an isomorph is possible, then it will have the same sort of conscious experience. FACING UP TO THE PROBLEM OF CONSCIOUSNESS 27 same abstract information space embedded in physical processing and in conscious experience. This leads to a natural hypothesis: that information (or at least some information) has two basic aspects, a physical aspect and a phenomenal aspect. This has the status of a basic principle that might underlie and explain the emergence of experience from the physical. Experience arises by virtue of its status as one aspect of information, when the other aspect is found embodied in physical processing. ‘This principle is lent support by a number of considerations, which I can only outline briefly here. First, consideration of the sort of physical changes that correspond to changes in conscious experience suggests that such changes are always relevant by virtue of their role in constituting informational changes — differences within an abstract space of states that are divided up precisely according to their causal differences along certain ‘causal pathways. Second, if the principle of organizational invariance is to hold, then we need to find some fundamental organizational property for experience to be linked to, and information is an organizational property par excellence. Third, this principle offers some hope of explaining the principle of structural coherence in terms of the structure present within information spaces. Fourth, analysis of the cognitive explanation of our judgments and claims about conscious experience — judgments that are functionally explainable but nevertheless deeply tied to experience itself — suggests that explanation centrally involves the information states embedded in cognitive processing. It follows that a theory based on information allows a deep coherence between the explanation of experience and the explanation of our judgments and claims about it. Wheeler (1990) has suggested that information is fundamental to the physics of the universe, According to this ‘it from bit’ doctrine, the laws of physics can be cast in terms of information, postulating different states that give rise to different effects without actually saying what those states are. Itis only their position in an information space that counts. If so, then information is a natural candidate to also play a role in a fundamental theory of consciousness. We are led to a conception of the world on which information is truly fundamental, and on which it has two basic aspects, corresponding to the physical and the phenomenal features of the world. Of course, the double-aspect principle is extremely speculative and is also underdeter- mined, leaving a number of key questions unanswered. An obvious question is whether all information has a phenomenal aspect. One possibility is that we need a further constraint on the fundamental theory, indicating just what sort of information has a phenomenal aspect. The other possibility is that there is no such constraint. If not, then experience is much more widespread than we might have believed, as information is everywhere. This is counterintuitive at first, but on reflection I think the position gains a certain plausibility and elegance, Where there is simple information processing, there is simple experience, and where there is complex information processing, there is complex experience, A mouse has a simpler information-processing structure than a human, and has correspondingly simpler experience; perhaps a thermostat, a maximally simple information processing structure, might have maximally simple experience? Indeed, if experience is truly a fundamental property, it would be surprising for it to arise only every now and then; most fundamental properties are more evenly spread. In any case, this is 28 DJ. CHALMERS very much an open question, but I believe that the position is not as implausible as it is often thought to be. Once a fundamental link between information and experience is on the table, the door is opened to some grander metaphysical speculation concerning the nature of the world. For example, it is often noted that physics characterizes its basic entities only extrinsi- cally, in terms of their relations to other entities, which are themselves characterized extrinsically, and so on. The intrinsic nature of physical entities is left aside. Some argue that no such intrinsic properties exist, but then one is left with a world that is pure causal flux (a pure flow of information) with no properties for the causation to relate. If one allows that intrinsic properties exist, a natural speculation given the above is that the intrinsic properties of the physical — the properties that causation ultimately relates — are themselves phenomenal properties. We might say that phenomenal properties are the internal aspect of information. This could answer a concer about the causal relevance of experience — a natural worry, given a picture on which the physical domain is causally closed, and on which experience is supplementary to the physical. The informational view allows us to understand how experience might have a subtle kind of causal relevance in virtue of its status as the intrinsic aspect of the physical. This metaphysical speculation is probably best ignored for the purposes of developing a scientific theory, but in addressing some philosophical issues it is quite suggestive. VII; Conclusion The theory I have presented is speculative, but it is a candidate theory. I suspect that the principles of structural coherence and organizational invariance will be planks in any satisfactory theory of consciousness; the status of the double-aspect theory of information is much less certain. Indeed, right now it is more of an idea than a theory. To have any hope of eventual explanatory success, it will have to be specified more fully and fleshed out into a more powerful form. Still, reflection on just what is plausible and implausible about it, on where it works and where it fails, can only lead to a better theory. Most existing theories of consciousness either deny the phenomenon, explain some- thing else, or elevate the problem to an eternal mystery. I hope to have shown that it is possible to make progress on the problem even while taking it seriously, To make further progress, we will need further investigation, more refined theories, and more careful analysis. The hard problem is a hard problem, but there is no reason to believe that it will remain permanently unsolved. FACING UP TO THE PROBLEM OF CONSCIOUSNESS 29 Further Reading ‘The problems of consciousness have been widely discussed in the recent philosophi- cal literature. For some conceptual clarification of the various problems of con- sciousness, see Block 1995, Nelkin 1993 and Tye 1995. Those who have stressed the difficulties of explaining experience in physical terms include Hodgson 1988, Jackson 1982, Levine 1983, Lockwood 1989, McGinn 1989, Nagel 1974, Seager 1991, Searle 1992, Strawson 1994 and Velmans 1991, among others. Those who take a reductive approach include Churchland 1995, Clark 1992, Dennett 1991, Dretske 1995, Kirk 1994, Rosenthal 1996 and Tye 1995. There have not been many attempts to build detailed nonreductive theories in the literature, but see Hodgson 1988 and Lockwood 1989 for some thoughts in that direction. Two excellent collections of recent articles on consciousness are Block, Flanagan and Gtizeldere 1996 and Metzinger 1995. References Akins, K. (1993), ‘What is it like to be boring and myopic?’ in Dennett and his Critics, ed. B. Dahlbom (Oxford: Blackwell). Allport, A. (1988), ‘What concept of consciousness?" in ( eds.) Consciousness in Contemporary Science, ed. A. Marcel and E. Bisiach (Oxford: Oxford University Press). Baars, B.J. (1988), A Cognitive Theory of Consciousness (Cambridge: Cambridge University Press). Bateson, G. (1972), Steps to an Ecology of Mind (Chandler Publishing). Block, N. (1995), ‘On a confusion about the function of consciousness’, Behavioral and Brain ‘Sciences, in press. Block, N, Flanagan, O. and Gitzeldere, G. (eds. 1996), The Nature of Consciousness: Philosophical and Scientific Debates (Cambridge, MA: MIT Press). Chalmers, DJ. (1996), The Conscious Mind (New York: Oxford University Press). Churchland, P.M. (1995), The Engine of Reason, The Seat of the Soul: A Philosophical Journey into the Brain (Cambridge, MA: MIT Press). Clark, A. (1992), Sensory Qualities (Oxford: Oxford University Press). Crick, F, and Koch, C. (1990), ‘Toward a neurobiological theory of consciousness’, Seminars in the Neurosciences, 2, pp. 263-75. Crick, F. (1994), The Astonishing Hypothesis: The Scientific Search for the Soul (New York: Seribners). Dennett, D.C. (1991), Consciousness Explained (Boston: Little, Brown). Dretske, Fl (1995), Naturalizing the Mind (Cambridge, MA: MIT Press). Edelman, G. (1989), The Remembered Present: A Biological Theory of Consciousness (New York: Basic Books). Farah, M.J. (1994), ‘Visual perception and visual awareness after brain damage: a tutorial ‘overview’, in Consciousness and Unconscious Information Processing: Attention and Perform- ance 15, ed, C. Umilta and M. Moscovitch (Cambridge, MA: MIT Press). Flohr, H. (1992), ‘Qualia and brain processes’, in Emergence or Reduction?: Prospects for Nonreductive Physicalism, ed. A. Beckerman, H. Flohr, and J. Kim (Berlin: De Gruyter). Hameroff, S.R. (1994), ‘Quantum coherence in microtubules: a neural basis for emergent consciousness?", Journal of Consciousness Studies, 1, pp. 91-118. Hardin, C.L. (1992), ‘Physiology, phenomenology, and Spinoza’s true colors’, in Emergence or Reduction?: Prospects for Nonreductive Physicalism, ed. A. Beckerman, H, Flohr, and J. Kim (Berlin: De Gruyter). Hill, C.S. (1991), Sensations: A Defense of Type Materialism (Cambridge: Cambridge University Press). Hodgson, D, (1988), The Mind Matters: Consciousness and Choice in a Quantum World (Oxford: Oxford University Press). Humphrey, N. (1992), A History of the Mind (New York: Simon and Schuster). Jackendoff, R. (1987), Consciousness and the Computational Mind (Cambridge, MA: MIT Press). 30 DJ. CHALMERS Jackson, F. (1982), ‘Epiphenomenal qualia’, Philosophical Quarterly, 32, pp. 127-36. Jackson, F. (1994), ‘Finding the mind in the natural world’, in Philosophy and the Cognitive Sciences, ed, R. Casati, B. Smith, and S. White (Vienna: Holder-Pichler-Tempsky). Kirk, R. (1994), Raw Feeling: A Philosophical Account of the Essence of Consciousness (Oxford: ‘Oxford University Press). Kripke, S. (1980), Naming and Necessity (Cambridge, MA: Harvard University Press). Levine, J. (1983), ‘Materialism and qualia: the explanatory gap’, Pacific Philosophical Quarterly, 64, pp. 354-61, Lewis, D. (1994), ‘Reduction of mind’, in A Companion to the Philosophy of Mind, ed. S. Guttenplan (Oxford: Blackwell). Libet, B. (1993), “The neural time factor in conscious and unconscious events’, in Experimental and Theoretical Studies of Consciousness (Ciba Foundation Symposium 174), ed. G.R. Block and J. Marsh (Chichester: John Wiley and Sons). Loar, B. (1990), ‘Phenomenal states’, Philosophical Perspectives, 4, pp. 81-108. Lockwood, M. (1989), Mind, Brai, and the Quantum (Oxford: Blackwell). McGinn, C. (1989), ‘Can we solve the mind-body problem?", Mind, 98, pp. 349-66. Metzinger, T. (ed. 1995), Conscious Experience (Exeter: Imprint Academic). Nagel, T. (1974), ‘What is it like to be a bat?', Philosophical Review, 4, pp. 435-50. Nelkin, N. (1993), ‘What is consciousness?, Philosophy of Science, 60, pp. 419-34. Newell, A. (1990), Unified Theories of Cognition (Cambridge, MA: Harvard University Press). Penrose, R. (1989), The Emperor's New Mind (Oxford: Oxford University Press). Penrose, R. (1994), Shadows of the Mind (Oxford: Oxford University Press). Rosenthal, D.M. (1996), ‘A theory of consciousness’, in The Nature of Consciousness, ed. N. Block, O. Flanagan, and G. Gilzeldere (Cambridge, MA: MIT Press). Seager, W.E. (1991), Metaphysics of Consciousness (London: Routledge). Searle, J.R. (1980), ‘Minds, brains and programs’, Behavioral and Brain Sciences, 3, pp. 417-57. Searle, I.R. (1992), The Rediscovery of the Mind (Cambridge, MA: MIT Press). Shallice, T. (1972), ‘Dual functions of consciousness’, Psychological Review, 79, pp. 383-93. Shannon, CE. (1948), ‘A mathematical theory of communication’, Bell Systems Technical Journal, 21, pp. 379-423. ‘Strawson, G. (1994), Mental Reality (Cambridge, MA: MIT Press). ‘Tye, M. (1995), Ten Problems of Consciousness (Cambridge, MA: MIT Press). Velmans, M. (1991), ‘Is human information-processing conscious?’ Behavioral and Brain Sciences, 14, pp. 651-69. Wheeler, J.A. (1990), ‘Information, physics, quantum: the search for links’, in Complexity, Entropy, and the Physics of Information, ed. W. Zurek (Redwood City, CA: Addison-Wesley), Wilkes, K.V. (1988), ‘—, Yishi, Duh, Um and consciousness’, in Consciousness in Contemporary Science, ed. A. Marcel and E. Bisiach (Oxford: Oxford University Press). Deflationary Perspectives Daniel C. Dennett Facing Backwards on the Problem of Consciousness ‘The strategy of divide and conquer is usually an excellent one, but it all depends on how you do the carving. Chalmers’ attempt to sort the ‘easy’ problems of consciousness from the ‘really hard’ problem is not, I think, a useful contribution to research, but a major misdirector of attention, an illusion-generator. How could this be? Let me describe two somewhat similar strategic proposals, and compare them to Chalmers’ recommendation. 1. The hard question for vitalism Imagine some vitalist who says to the molecular biologists: ‘The easy problems of life include those of explaining the following phenomena: reproduction, development, growth, metabolism, self-repair, immunological self- defence . .. These are not all that easy, of course, and it may take another century or 0 to work out the fine points, but they are easy compared to the really hard problem: life itself, We can imagine something that was capable of reproduction, development, growth, metabolism, self-repair and immunological self-defence, but that wasn’t, you know, alive. The residual mystery of life would be untouched by solutions to all the easy problems. In fact, when I read your accounts of life, I am left feeling like the victim of a bait-and-switch. This imaginary vitalist just doesn't see how the solution to all the easy problems amounts to a solution to the imagined hard problem. Somehow this vitalist has got under the impression that being alive is something over and above all these subsidiary compo- nent phenomena. I don’t know what we can do about such a person beyond just patiently saying: your exercise in imagination has misfired; you can’t imagine what you say you can, and just saying you can doesn’t cut any ice. (Dennett, 1991, p. 281-2.) 2. The hard question for Crock Francis Crick (1994) gives us an example of what happens when you adopt Chalmers” distinction, when he says, at the close of his book on consciousness. ‘I have said almost nothing about qualia — the redness of red — except to brush it to one side and hope for This paper was originally published in the Journal of Consciousness Studies, 3, No.1 (1996), pp. 4-6. 34 D.C. DENNETT the best.” (p. 256.) But consider what would be wrong with the following claim made by an imaginary neuroscientist (Crock) substituting ‘perception’ for ‘qualia’ in the quotation from Crick: ‘I have said almost nothing about perception — the actual analysis and comprehension of the visual input — except to brush it to one side and hope for the best.” Today we can all recognize that whatever came before Crock’s declaration would be forlorn, because not so many years ago this was a mistake that brain scientists actually made: they succumbed all too often to the temptation to treat vision as if it were television — as if it were simply a matter of getting ‘the picture’ from the eyes to the screen somewhere in the middle where it could be handsomely reproduced so that the phenom- ena of appreciation and analysis could then get underway. Today we realize that the analysis — the whatever you want to call it that composes, in the end, all the visual understanding — begins right away, on the retina; if you postpone consideration of it, you misdescribe how vision works, Crock has made a mistake: he has created an artifactual ‘hard’ problem of perception, not noticing that it evaporates when the piece- meal work on the easy problems is completed. Is it similarly a mistake for Crick, following Chalmers, to think that he can make progress on the easy questions of consciousness without in the process answering the hard question? I think so (Dennett, 1991). I make the parallel claim about the purported “subjective qualities’ or ‘qualia’ of experience: if you don’t begin breaking them down into their (functional) components from the outset, and distributing them throughout your model, you create a monster — an imaginary dazzle in the eye of a Cartesian homunculus (Dennett, 1995). Chalmers has not yet fallen in either of these traps — not quite. He understands that he must show how his strategic proposal differs from these, which he recognizes as doomed. He attempts this by claiming that consciousness is strikingly unlike life, and unlike the features of perception misconstrued by Crock: when it comes to consciousness, the hard problem is ‘almost unique’ in that it “goes beyond problems about the performance of functions.’ Almost unique? He gives us no other phenomena with this special feature, but in any case, what he says in support of this claim simply repeats the claim in different words (p.12): To see this, note that when we have explained the performance of all the cognitive and behavioural functions in the vicinity of experience . . . there may still remain a further unanswered question: Why is the performance of these functions accompa- nied by experience? A simple explanation of the functions leaves this question open. Our vitalist can surely ask the same dreary question: Why is the performance of these functions accompanied by life? Chalmers says that this would be a conceptual mistake on the part of the vitalist, and I agree, but he needs to defend his claim that his counterpart is not a conceptual mistake as well. When he confronts the vitalist parallel! head-on, he simply declares that whereas vitalist scepticism was driven by doubts about whether physical mechanisms could ‘perform the many remarkable functions associated with life’, it is otherwise with his scepticism: With experience, on the other hand, physical explanation of the functions is not in question. The key is instead the conceptual point that the explanation of functions does not suffice for the explanation of experience. (p. 18.) FACING BACKWARDS ON THE PROBLEM OF CONSCIOUSNESS 35 I submit that he is flatly mistaken in this claim, Whether people realize it or not, it is precisely the ‘remarkable functions associated with’ consciousness that drive them to wonder about how consciousness could possible reside in a brain. In fact, if you carefully dissociate all these remarkable functions from consciousness — in your own, first-person case — there is nothing left for you to wonder about. What impresses me about my own consciousness, as I know it so intimately, is my delight in some features and dismay over others, my distraction and concentration, my unnamable sinking feelings of foreboding and my blithe disregard of some perceptual details, my obsessions and oversights, my ability to conjure up fantasies, my inability to hold more than a few items in consciousness at a time, my ability to be moved to tears by a vivid recollection of the death of a loved one, my inability to catch myself in the act of framing the words I sometimes say to myself, and so forth. These are all ‘merely’ the ‘performance of functions’ or the manifestation of various complex dispositions to perform functions. In the course of making an introspective catalogue of evidence, I wouldn't know what I was thinking about if I couldn't identify them for myself by these functional differentia, Subtract them away, and nothing is left beyond a weird conviction (in some people) that there is some ineffable residue of ‘qualitative content’ bereft of all powers to move us, delight us, annoy us, remind us of anything. Chalmers recommends a parallel with physics, but it backfires. He suggests that a theory of consciousness should ‘take experience itself as a fundamental feature of the world, alongside mass, charge, and space-time.’ As he correctly notes, ‘No attempt is made [by physicists] to explain these features in terms of anything simpler,” but they do cite the independent evidence that has driven them to introduce these fundamental categories. Chalmers needs a similar argument in support of his proposal, but when we ask what data are driving him to introduce this concept, the answer is disappointing: It is a belief in a fundamental phenomenon of ‘experience’. The introduction of the concept does not do any explanatory work. The evidential argument is circular. (Roberts, 1995, fin 8.) We can see this by comparing Chalmers’ proposal with yet one more imaginary non-starter: cutism, the proposal that since some things are just plain cute, and other things aren’t cute at all — you can just see it, however hard it is to describe or explain — we had better postulate cuteness as a fundamental property of physics alongside mass, charge and space-time. (Cuteness is not a functional property, of course; I can imagine somebody who wasn't actually cute at all but who nevertheless functioned exactly as if ‘cute — trust me.) Cutism is in even worse shape than vitalism. Nobody would have taken vitalism seriously for a minute if the vitalists hadn’t had a set of independently describ- able phenomena — of reproduction, metabolism, self-repair and the like — that their postulated fundamental life-element was hoped to account for. Once these phenomena were otherwise accounted for, vitalism fell flat, but at least it had a project. Until Chalmers gives us an independent ground for contemplating the drastic move of adding ‘experience’ to mass, charge, and space-time, his proposal is one that can be put on the back burner. 36 D.C. DENNETT References Chalmers, David (1995), ‘Facing up to the problem of consciousness’, Journal of Consciousness Studies, 2 (3), pp. 200-19 (reprinted in this volume). Ceick, Francs (1994), The Astonishing Hyposhesie: The Stemifc Search forthe Sou! (New Yor cribners). Dennett, Danie! (1991), Consciousness Explained (Boston, MA: Little, Brown & Co.). Dennett, Daniel (1995), ‘Our vegetative soul’ — review of Damasio, Descartes’ Error, in Times Literary Supplement, August 25, 1995, pp. 3-4. Roberts, John (1995), “Qualia and animal consciousness’, Center for Cognitive Studies working Paper, 1995-#1, Tufts University (unpublished). Patricia Smith Churchland The Hornswoggle Problem I: Introduction Conceptualizing a problem so we can ask the right questions and design revealing experiments is crucial to discovering a satisfactory solution to the problem. Asking where animal spirits are concocted, for example, turns out not to be the right question to ask about the heart. When Harvey asked instead, ‘How much blood does the heart pump in an hour?", he conceptualized the problem of heart function very differently. The recon- ceptualization was pivotal in coming to understand that the heart is really a pump for circulating blood; there are no animal spirits to concoct. My strategy here,' therefore, is to take the label, ‘The Hard Problem’ in a constructive spirit — as an attempt to provide a useful conceptualization concerning the very nature of consciousness that could help steer us in the direction of a solution. My remarks will focus mainly on whether in fact anything positive is to be gained from the ‘hard problem’ characterization, or whether that conceptualization is counterproductive. I cannot hope to do full justice to the task in short compass, especially as this characterization of the problem of consciousness has a rather large literature surrounding it. The watershed articulation of this view of the problem is Thomas Nagel’ classic paper “What is it like to be a bat?" (1974) In his opening remarks, Nagel comes straight to the point: ‘Consciousness is what makes the mind-body problem really intractable.’ Deline- ating a contrast between the problem of consciousness and all other mind-body problems, ‘Nagel asserts: ‘While an account of the physical basis of mind must explain many things, this (conscious experience] appears to be the most difficult.’ Following Nagel’s lead, many other philosophers, including Frank Jackson, Saul Kripke, Colin McGinn, John Searle, and most recently, David Chalmers, have extended and developed Nagel’s basic idea that consciousness is not tractable neuroscientifically. Although I agree that consciousness is, certainly, a difficult problem, difficulty per se does not distinguish it from oodles of other neuroscientific problems. Such as how the brains of homeotherms keep a constant internal temperature despite varying external conditions. Such as the brain basis for schizophrenia and autism. Such as why we dream and sleep. Supposedly, something sets consciousness apart from all other macro-function | This paper was originally published in the Journal of Consciousness Studies, 3, No.5/6 (1996), pp. 402-8. It is based on a talk presented by the author at the ‘Tuscon II’ conference, “Toward a Science of Consciousness’ held at Tucson, Arizona, in April 1996. Many thanks are owed to the organizers of the meeting, and thanks also to Paul Churchland, David Rosenthal, Rodolfo Llinés, Michael Stack, Dan Dennett, Ilya Farber and Joe Ramsay for advice and ideas. 38 P.S. CHURCHLAND brain riddles such that it stands alone as The Hard Problem. As I have tried to probe precisely what that is, I find my reservations multiplying. I: Carving Up the Problem Space The-Hard-Problem label invites us to adopt a principled empirical division between consciousness (the hard problem) and problems on the ‘easy’ (or perhaps hard but not Hard?) side of the ledger. The latter presumably encompass problems such as the nature of short-term memory, long-term memory, autobiographical memory, the nature of representation, the nature of sensori-motor integration, top-down effects in perception — not to mention such capacities as attention, depth perception, intelligent eye movement, skill acquisition, planning, decision-making, and so forth. On the other side of the ledger, all on its own, stands consciousness — a uniquely hard problem. My lead-off reservation arises from this question: what is the rationale for drawing the division exactly there? Dividing off consciousness from all of the so-called ‘easy prob- ems’ listed above implies that we could understand all those phenomena and still not know what it was for... what? The ‘qualia-light’ to go on? ? Is that an insightful conceptualization? What exactly is the evidence that we could explain all the ‘easy’ phenomena and still not understand the neural mechanisms for consciousness? (Call this the ‘left-out’ hypothesis.) That someone can imagine the possibility is not evidence for the real possibility. It is only evidence that somebody or other believes it to be a possibility. That, on its own, is not especially interesting, Imaginary evidence, needless to say, is not as interesting as real evidence, and what needs to be produced is some real evidence The left-out hypothesis — that consciousness would still be a mystery, even if we could explain all the easy problems — is dubious on another count: it begs the question against those theories that are exploring the possibility that functions such as attention and short-term memory are crucial elements in the consciousness (see especially Crick, 1994; P.M. Churchland, 1995). The rationale sustaining this approach stems from observations such as that awake persons can be unaware of stimuli to which they are not paying attention, but can become aware those stimuli when attention shifts. There is a vast psychological literature, and a nontrivial neuroscientific literature, on this topic, Some of it powerfully suggests that attention and awareness are pretty closely connected. The approach might of course be wrong, for it is an empirical conjecture. But if it is wrong, itis wrong because of the facts, not because of an arm-chair definition, The trouble with 2 As lacked time in my talk at Tucson to address the ‘Mary’ problem, a problem first formulated by Frank Jackson in 1982, let me make several brief remarks about it here. In sum, Jackson's idea ‘was that there could exist someone, call her Mary, who knew everything there was to know about how the brain works but still did not know what it was to see the colour green (suppose she lacked “green cones’ , to put it crudely). This possibility Jackson took to show that qualia are therefore not explainable by science, The main problem with the argument is that to experience green qualia, certain wiring has to be in place in Mary’s brain, and certain patterns of activity have to obtain and since, by Jackson's own hypothesis, she does not have that wiring, then presumably the relevant activity patterns in visual cortex are not caused and she does not experience green. Who would expect her visual cortex — V4, say — would be set ahumming just by virtue of her propositional linguistic) knowledge about activity patterns in V4? Not me, anyhow. She can have propositional knowledge via other channels, of course, including the knowledge of what her own brain lacks vis 4 vis green qualia. Nothing whatever follows about whether science can or cannot explain qualia. THE HORNSWOGGLE PROBLEM 39 the ‘hard problem’ characterization is that on the strength of a proprietary definition, it rejects them as wrong. I do find that unappealing, since the nature of consciousness is an empirical problem, not problem that can be untangled by semantic gerrymandering. ‘What drives the left-out hypothesis? Essentially, a thought-experiment, which roughly goes as follows: we can conceive of a person, like us in all the aforementioned easy-to- explain capacities (attention, short term memory, etc.), but lacking qualia. This person would be exactly like us, save that he would be a Zombie — an anaqualiac, one might say, Since the scenario is conceivable, itis possible, and since itis possible, then whatever consciousness is, it is explanatorily independent of those activities. I take this argument to be a demonstration of the feebleness of thought-experiments. Saying something is possible does not thereby guarantee it is a possibility, so how do we know the anaqualiac idea is really possible? To insist that it must be is simply to beg the question at issue. As Francis Crick has observed, it might be like saying that one can imagine a possible world where gases do not get hot, even though their constituent molecules are moving at high velocity. As an argument against the empirical identifica- tion of temperature with mean molecular KE, the thermodynamic thought-experiment is feebleness itself. Is the problem on the ‘hard’ side of the ledger sufficiently well-defined to sustain the division as a fundamental empirical principle? Although it is easy enough to agree about the presence of qualia in certain prototypical cases, such as the pain felt after a brick has fallen on a bare foot, or the blueness of the sky on a sunny summer afternoon, things are less clear-cut once we move beyond the favoured prototypes. Some of our perceptual capacities are rather subtle, as, for example, positional sense is often claimed to be. Some philosophers, e.g. Elizabeth Anscombe, have actually opined that we can know the position of our limbs without any ‘limb-position’ qualia. As for me, I am inclined to say I do have qualitative experiences of where my limbs are — it feels different to have my fingers clenched than unclenched, even when they are not visible. The disagreement itself, however, betokens the lack of consensus once cases are at some remove from the central prototypes. Vestibular system qualia are yet another non-prototypical case. Is there something “vestibular-y’ it feels like to have my head moving? To know which way is up? Whatever the answer here, at least the answer is not glaringly obvious. Do eye movements have eye-movement qualia? Some maybe do, and some maybe do not. Are there ‘introspective qualia’, or is introspection just paying attention to perceptual qualia and talking to yourself? Ditto, plus or minus a bit, for self-awareness. Thoughts are also a bit problem- atic in the qualia department. Some of my thoughts seem to me to be a bit like talking to myself and hence like auditory imagery but some just come out of my mouth as I am talking to someone or affect decisions without ever surfacing as a bit of inner dialogue. None of this is to deny the pizzazz of qualia in the prototypical cases, Rather, the point is just that prototypical cases give us only a starting point for further investigation, and nothing like a full characterization of the class to which they belong. My suspicion with respect to The Hard Problem strategy is that it seems to take the class of conscious experiences to be much better defined than it is. The point is, if you 3 Something akin to this was argued by Saul Kripke in the 1970's, 40 P.S. CHURCHLAND are careful to restrict your focus to the prototypical cases, you can easily be hornswoggled into assuming the class is well-defined. As soon as you broaden your horizons, trouble- some questions about fuzzy boundaries, about the connections between attention, short term memory and awareness, are present in full, what-do-we-do-with-that glory. Are the easy problems known to be easier than The Hard Problem? Is the hard/easy division grounded in fact? To begin with, it is important to acknowledge that for none of the so-called ‘easy’ problems, do we have an understanding of their solution (see the partial list on p. 38). Itis just false that we have anything approximating a comprehensive theory of sensori-motor control or attention or short-term memory or long-term memory. Consider one example. A signature is recognizably the same whether signed with the domi- nant or non-dominant hand, with the foot, with the mouth or with the pen strapped to the shoulder. How is ‘my signature’ represented in the nervous system? How can completely different muscle sets be invoked to do the task, even when the skill was not acquired using those muscles? We do not understand the general nature of motor representation, Notice that it is not merely that we are lacking details, albeit important details. The fact is, we are lacking important conceptual/theoretical ideas about how the nervous system performs fundamental functions — such as time management, such as motor control, such as learning, such as information retrieval. We do not understand the role of back projections, or the degree to which processing is organized hierarchically. These are genuine puzzles, and it is unwise to ‘molehill’ them in order to ‘mountain’ up the problem of consciousness. Although quite a lot is known at the cellular level, the fact remains that how real neural networks work and how their output properties depend on cellular properties still abounds with nontrivial mysteries. Naturally I do not wish to minimize the progress that has been made in neuroscience, but it is prudent to have a cautious assessment of what we really do not yet understand. Carving the explanatory space of mind-brain phenomena along the hard and the easy line, as Chalmers proposes, poses the danger of inventing an explanatory chasm where there really exists just a broad field of ignorance. It reminds me of the division, deep to medieval physicists, between sublunary physics (motion of things below the level of the moon) and superlunary physics (motion of things above the level of the moon). The conviction was that sublunary physics was tractable, and it is essentially based on Aristotelian physics. Heavy things fall because they have gravity, and fall to their natural place, namely the earth, which is the centre of the universe. Things like smoke have levity, and consequently they rise, up being their natural place. Everything in the sublunary realm has a ‘natural place’, and that is the key to explaining the behaviour of sublunary objects. Superlunary events, by contrast, we can neither explain nor under- stand, but in any case, they have neither the gravity nor levity typical of sublunary things. This old division was not without merit, and it did entail that events such as planetary motion and meteors were considered unexplainable in terrestrial terms, but probably were divinely governed. Although I do not know that Chalmers’ easy/hard distinction will prove ultimately as misdirected as the sublunary/superlunary distinction, neither do I know it is any more sound, What I do suspect, however, is that it is much too early in the science of nervous systems to command much credence. ‘The danger inherent in embracing the distinction as a principled empirical distinction is that it provokes the intuition that only a real humdinger of a solution will suit The Hard THE HORNSWOGGLE PROBLEM 41 Problem. Thus the idea seems to go as follows: the answer, if it comes at all, is going to have to come from somewhere Really Deep — like quantum mechanics — or pethaps it requires a whole new physics. As the lone enigma, consciousness surely cannot be just a matter of a complex dynamical system doing its thing. Yes, there are emergent properties from nervous systems such as co-ordinated movement as when an owl catches a mouse, but consciousness (the hard problem) is an emergent property like unto no other. Consequently, it will require a very deep, very radical solution. That much is evident sheerly from the hardness of The Hard Problem. Tconfess I cannot actually see that. I do not know anything like enough to see how to solve either the problem of sensori-motor control or the problem of consciousness. I certainly cannot see enough to know that one problem will, and the other will not, require a Humdinger solution. Ill: Using Ignorance as a Premise In general, what substantive conclusions can be drawn when science has not advanced very far on a problem? Not much. One of the basic skills we teach our philosophy students is how to recognize and diagnose the range of nonformal fallacies that can undermine an ostensibly appealing argument: what it is to beg the question, what a non sequitur is, and soon. A prominent item in the fallacy roster is argumentum ad ignorantiam — argument from ignorance. The canonical version of this fallacy uses ignorance as the key premise from which a substantive conclusion is drawn. The canonical version looks like this: We really do not understand much about a phenomenon P. (Science is largely ignorant about the nature of P.) Therefore: we do know that: (1) P can never be explained, or (2) Nothing science could ever discover would deepen our understanding of P, or (3) P can never be explained in terms of properties of kind S. In its canonical version, the argument is obviously a fallacy: none of the tendered conclusions follow, not even a litle bit. Surrounded with rhetorical flourish, much brow furrowing and hand-wringing, however, versions of this argument can homswoggle the unwary. From the fact that we do not know something, nothing very interesting follows — we just don’t know. Nevertheless, the temptation to suspect that our ignorance is telling us something positive, something deep, something metaphysical or even radical, is ever- present, Pethaps we like to put our ignorance in a positive light, supposing that but for the Profundity of the phenomenon, we would have knowledge. But there are many reasons for not knowing, and the specialness of the phenomenon is, quite regularly, not the real reason. I am currently ignorant of what caused an unusual rapping noise in the woods last night. Can I conclude it must be something special, something unimaginable, something . . . alien . . . other-worldly? Evidently not. For all I can tell now, it might merely have been a raccoon gnawing on the compost bin. Lack of evidence for something is just that: lack of evidence. It is not positive evidence for something else, let alone 42 P.S. CHURCHLAND something of a humdingerish sort. That conclusion is not very glamorous perhaps, but when ignorance is a premise, that is about all you can grind out of it. Now if neuroscience had progressed as far on the problems of brain function as molecular biology has progressed on transmission of hereditary traits, then of course we would be in a different position. But it has not, The only thing you can conclude from the fact that attention is mysterious, or sensori-motor integration is mysterious, or that consciousness is mysterious, is that we do not understand the mechanisms. Moreover, the mysteriousness of a problem is not a fact about the problem, it is not a metaphysical feature of the universe — it is an epistemological fact about us. It is about where we are in current science, it is about what we can and cannot understand, it is about what, given the rest of our understanding, we can and cannot imagine. It is not a property of the problem itself. It is sometimes assumed that there can be a valid transition from ‘we cannot now explain’ to ‘we can never explain’ , so long as we have the help of a subsidiary premise, namely, ‘I cannot imagine how we could ever explain . . .’ But it does nor help, and this transition remains a straight-up application of argument from ignorance. Adding ‘I cannot imagine explaining P’ merely adds a psychological fact about the speaker, from which again, nothing significant follows about the nature of the phenomenon in question. Whether we can or cannot imagine a phenomenon being explained in a certain way is a psychological fact about us, not an objective fact about the nature of the phenomenon itself, To repeat: it is an epistemological fact about what, given our current knowledge, we can and cannot understand, It is not a metaphysical fact about the nature of the reality of the universe, Typical of vitalists generally, my high school biology teacher argued for vitalism thus: I cannot imagine how you could get living things out of dead molecules. Out of bits of proteins, fats, sugars — how could life itself emerge? He thought it was obvious from the sheer mysteriousness of the matter that it could have no solution in biology or chemistry. He assumed he could tell that it would require a Humdinger solution. Typical of lone survivors, a passenger of a crashed plane will say: I cannot imagine how I alone could have survived the crash, when all other passengers died instantly. Therefore God must have plucked me from the jaws of death. Given that neuroscience is still very much in its early stages, it is actually not a very interesting fact that someone or other cannot imagine a certain kind of explanation of some brain phenomenon. Aristotle could not imagine how a complex organism could come from a fertilized egg. That of course was a fact about Aristotle, not a fact about embryogenesis. Given the early days of science (500 BC), it is no surprise that he could not imagine what it took many scientists hundreds of years to discover. I cannot imagine how ravens can solve a multi-step problem in one trial, or how temporal integration is achieved, or how thermoregulation is managed. But this is a (not very interesting) psychological fact about me. One could, of course, use various rhetorical devices to make it seem like an interesting fact about me, perhaps by emphasizing that it is a really really hard problem; but if we are going to be sensible about this, it is clear that my inability to imagine how thermoregulation works is au fond, pretty boring. ‘The ‘I-cannot-imagine’ gambit suffers in another way. Being able to imagine an explanation for P is a highly open-ended and under-specified business. Given the poverty THE HORNSWOGGLE PROBLEM 43 of delimiting conditions of the operation, you can pretty much rig the conclusion to go whichever way your heart desires. Logically, however, that flexibility is the kiss of death. ‘Suppose someone claims that she can imagine the mechanisms for sensori-motor integration in the human brain but cannot imagine the mechanisms for consciousness. What exactly does this difference amount to? Can she imagine the former in detail? No, because the details are not known. What is it, precisely, that she can imagine? Suppose she answers that in a very general way she imagines that sensory neurons interact with interneurons that interact with motor neurons, and via these interactions, sensori-motor integration is achieved. Now if that is all ‘being able to imagine’ takes, one might as well say one can imagine the mechanisms underlying consciousness. Thus: ‘The interneurons do it.’ The point is this: if you want to contrast being able to imagine brain mechanisms for attention, short term memory, planning etc., with being unable to imagine mecha- nisms for consciousness, you have to do more that say you can imagine neurons doing one but cannot imagine neurons doing the other. Otherwise one simply begs the question. To fill out the point, consider several telling examples from the history of science. Before the turn of the twentieth century, people thought that the problem of the precession of the perihelion of Mercury was essentially trivial. It was annoying, but ultimately, it would sort itself out as more data came in. With the advantage of hindsight, we can see that assessing this as an easy problem was quite wrong — it took the Einsteinian revolution in physics to solve the problem of the precession of the perihelion of Mercury. By contrast, a really hard problem was thought to be the composition of the stars. How could a sample ever be obtained? With the advent of spectral analysis, that turned out to be a readily solvable problem, When heated, the elements turn out to have a kind of fingerprint, easily seen when light emitted from a source is passed through a prism. Consider now a biological example. Before 1953, many people believed, on rather good grounds actually, that in order to address the copying problem (transmission of traits from parents to offspring), you would first have to solve the problem of how proteins fold. The former was deemed a much harder problem than the latter, and many scientists believed it was foolhardy to attack the copying problem directly. As we all know now, the basic answer to the copying problem lay in the base-pairing of DNA, and it was solved first. Humbling it is to realize that the problem of protein folding (secondary and tertiary) is still not solved. That, given the lot we now know, does seem to be a hard problem. What is the point of these stories? They reinforce the message of the argument from ignorance: from the vantage point of ignorance, it is often very difficult to tell which problem is harder, which will fall first, what problem will turn out to be more tractable than some other, Consequently our judgments about relative difficulty or ultimate tracta- bility should be appropriately qualified and tentative. Guesswork has a useful place, of course, but let’s distinguish between blind guesswork and educated guesswork, and between guesswork and confirmed fact. The philosophical lesson I learned from my biology teacher is this: when not much is known about a topic, don’t take terribly seriously someone else’s heartfelt conviction about what problems are scientifically tractable. Learn the science, do the science, and see what happens. 44 P.S. CHURCHLAND References Chalmers, David (1995), ‘Facing up to the problem of consciousness’, Journal of Consciousness Studies, 2 (3), pp. 200-19 (reprinted in this volume). Churchland, Paul M. (1995), The Engine of Reason; The Seat of the Soul (Cambridge, MA: MIT Press). Crick, Francis (1994), The Astonishing Hypothesis (New York: Scribner and Sons). Jackson, Frank (f982), ‘Epiphenomenal qualia’, Philosophical Quarterly, 32, pp. 127-36. Nagel, Thomas (1974), ‘What is it like to be a bat?", Philosophical Review, 83, pp.435-50. Thomas W. Clark Function and Phenomenology: Closing the Explanatory Gap To truly explain consciousness, we must find a convincing place for it in the natural world and ultimately in the scientific description of that world as expressed in physical, biological, and information theory. The default assumption when undertaking an expla- nation of consciousness should be that there is nothing ontologically special about it, nothing which sets it apart from the rest of nature as we presently conceive it. In his keynote paper, ‘Facing Up to the Problem of Consciousness’, David Chalmers assumes much the opposite: subjectivity is something more than the naturally evolved neural processes which seem likely, given the available evidence, to instantiate it. Qualitative consciousness — qualia or phenomenal experience — is said to ‘arise from’ or ‘accom- pany’ or ‘emerge from’ these processes. This dualism of underlying process versus accompanying subjectivity creates the ‘explanatory gap’ that so worries Chalmers and other philosophers: why should subjectivity arise from some processes and not others? My claim is that the central mystery about qualitative consciousness supposedly in need of explanation is an artifact generated by this presupposition about its nature — that it is an ‘effect’ of an underlying process — and itis precisely this that we must question if we are to find the true place of consciousness in the world. My strategy will be first to show that Chalmers’ initial assumption about consciousness compromises widely held methodological canons of scientific theory construction that he himself avows. Next, I will suggest some reasons why this assumption has gained such currency, despite its serious shortcomings. In parts three and four, using the principles of Chalmers’ theory of consciousness as discussion points, I will sketch an alternative picture, that subjectivity is certain sorts of physical, functional, informationally rich, and behaviour controlling (cybernetic) processes. Interestingly, Chalmers does much to support this picture in the latter part of his paper. Only the assumption that qualitative experience somehow arises from such functional processes prevents him from reaching the conclusion that I will defend here, that experience is identical to them. It is the elucidation of this (proposed) identity by the empirical investigation of what precisely these processes do — in contrast to unconscious processes — that will eventually 46 T.W. CLARK constitute a robust explanation of consciousness. On this picture, there is no question needing an answer about why just these processes give rise to consciousness, since indeed it never ‘arises’ or ‘emerges’ at all. Consciousness is what we consist of as physically instantiated subjects, not something extra that our brains create, I The hard problem of consciousness, according to Chalmers, is that of experience. ‘It is widely agreed that experience arises from a physical basis, but we have no good explanation of why and how it so arises. Why should physical processing give rise to a rich inner life at all? It seems objectively unreasonable that it should, and yet it does’ (p. 11). Here, at the very outset, a specific conception of consciousness sets the stage for the rest of Chalmers’ investigation. On this widely held view, experience emerges out of and accompanies certain neural functions, but is assumed not to be identical to these functions, The pressing question thus becomes, as Chalmers puts it, ‘Why is the perform- ance of these functions accompanied by experience?’ (p. 12, original emphasis). If experience is taken to be something over and above neurally instantiated functions, something extra which accompanies them, the ‘central mystery’ of consciousness be- comes the ‘explanatory gap .. . between function and experience.’ (Levine 1993 takes more or less the same position, pp. 130-5.) This in turn leads Chalmers to suppose that “To account for experience, we need an extra ingredient in the explanation’ (p. 16-17, original emphasis). He despairs of finding an explanation of consciousness within existing scientific theory, since physicalist and functionalist accounts will forever omit this extra ingredient: For any physical process we specify there will be an unanswered question: Why should this process give rise to experience? Given any such process, it is conceptu- ally coherent that it could be instantiated in the absence of experience. It follows that no mere account of the physical process will tell us why experience arises. The emergence of experience goes beyond what can be derived from physical theory. (p. 18) Of course, it is conceptually coherent that experience might be absent in the presence of certain physical, functional processes, but only on the assumption that they may not be, and probably are not, identical. But this is precisely the fundamental issue that must not be prejudged. As David Cole has pointed out recently, When a critic [of functionalism] supposes that there could be, ‘ex hypothesi,’ a system that instantiates the functional architecture yet fails to have the experience, the critic is ‘ex hypothesis’ supposing that any functionalist theory is false. To simply suppose that the theory is false is question begging, and cannot be the basis of an adequate argument against functionalism. (1994, p. 297, original emphasis.) If we begin our investigation into the nature of consciousness with Chalmers’ picture in mind — the picture of experience ‘arising from’ (hence not identical to) functional neural processes — naturally we will be led to doubt that functional explanations can fully account for experience. But the burden is on Chalmers to justify his starting point and its implicit dualism, the dualism that leads him to suppose that there must be an ‘extra FUNCTION AND PHENOMENOLOGY 47 ingredient’ in any explanation of qualitative consciousness which goes beyond descrip- tions of functions and physical processes. As Francis Crick observed recently, ‘Whether there is something extra remains to be seen’ (Crick, 1994 p. 12). Why, if we are conducting a more or less scientifically motivated investigation of a phenomenon, should we begin by assuming that this phenomenon is probably not of a piece with the rest of nature as we currently conceive it? The natural starting point, on the contrary, is to assume parsimoniously that we need not add experience, as Chalmers recommends we do, to our ‘fundamental’ ontological categories. Generally, we shouldn't posit as funda- mental that which we are seeking to explain. Instead, we should start with what seems clearly a simpler and more straightforward hypothesis, namely that experience is identi- cal to certain neurally instantiated cybernetic functions. The way is then open to under- stand consciousness by empirical investigation of whatever functions are found to correlate with experience. Such an explanation would best embody the virtues of ‘sim- plicity, elegance, and even beauty’ that Chalmers cites as hallmarks of good theory. Chalmers says that it is ‘a conceptual point that the explanation of functions does not suffice for the explanation of experience’ (p. 18, original emphasis), but again this is a conceptual point only under the particular initial conception of experience that Chalmers adopts, a conception that in effect assumes a conclusion about the nature of consciousness that although widely accepted, can hardly serve as a methodologically sound starting point. After all, the basic explanatory motive in science and philosophy is to incorporate heretofore inexplicable phenomena into an existing theoretical framework, modifying the framework only as minimally necessary to effect the incorporation. This motive is defeated by assuming at the very start that consciousness is a phenomenon that transcends the explanatory reach of existing theory. ‘The debate about where to begin in explaining consciousness can thus be framed as a competition between powerful and widespread intuitions that there is something ‘extra’ about consciousness in need of explanation and commitments to standard scientific explanatory practice. I suggest that we are better off allied with the latter than the former, and that the history of successful science is on my side. It is not that the identity of phenomenal consciousness and cognitive function is obviously true, since if it were we'd all be functionalists. Upon further investigation it may turn out (although I doubt it ) that experience does not correlate with any particular set of functions, or that in some essential respect it floats free of any physical or functional property. I am only arguing that it is methodologically more circumspect to start off with a simpler hypothesis, one which does not posit a special nature or essence of subjectivity to be explained. For some, this may appear to prejudge the issue in the opposite direction, since it seems to deny the existence of the explanandum itself, at least in the form they are used to conceptualizing it. But my suggestion is only that we remain open about the nature of consciousness as the investigation gets underway, and that our starting hypothesis be shaped by methodological constraints, not by concepts or intuitions motivating the supposition, at the start of the investigation, that there likely exist fundamental properties or entities not already included in naturalistic theories. 48 T.W. CLARK IL What is so compelling about Chalmers’ picture of subjectivity that it tends to override methodological considerations? That is, what leads many philosophers, scientists, and humanists to strongly doubt, at the outset, that qualitative consciousness might simply be identical to certain types of functional organization and hence suppose that it is some sort of contingent effect or accompaniment? This is a different question from asking why they might merely doubt that such an identity holds true. The answer to the latter, of course, is simply that the identity of phenomenal experience (qualia) and cognitive function is not obvious, and should be subject to the same legitimate methodological scepticism accorded any other hypothesis, But the answer to the former lies in appreciating the effect of several distinct trends in our philosophical tradition, some general and some specific, which bias our intuitions about the nature of consciousness against the possibility of identity. First and perhaps foremost is the residual, and by now mostly subliminal, grip of Cartesianism on basic assumptions about mental life in general. This influence leads us to regard talk of the mental not just as a useful, predictive, ‘intentional stance’ (Dennett 1987) sort of discourse, but as talk that refers to an independent, or at least parallel realm which interacts with or accompanies physical processes. Pain, pleasure, perceptions, emotions, thoughts, beliefs, and other mental phenomena present themselves as a related set of states and events which for everyday predictive purposes seem virtually autono- mous from the physical and functional (or, to use Dennett’s typology, the physical and design stances). We need not ordinarily bring in physicalist or functionalist talk in order to get by in the interpersonal world, so it is unsurprising that the commonsense assump- tion about the autonomy — and hence the ontological separateness — of the mental still generates a philosophical bias despite the best efforts of many philosophers to uproot it. T am not suggesting that dualism (of either substance or properties) is a priori unten- able, although it has well-known difficulties, but only making the uncontroversial point that our philosophical tradition has been heavily influenced by Cartesian intuitions about the ontological divide between the mental and the physical, intuitions reinforced by ordinary mentalistic discourse. Thus when it comes to choosing a picture of conscious- ness to start our investigations (and we have to choose some picture, after all) we may well be biased in favour of one which puts consciousness in the autonomous realm of the mental, even if only in a rather subtle, sophisticated way. Chalmers’ statements to the effect that certain neural processes ‘give rise’ to qualia constitute exactly such a picture, one which splits the mental from the physical/functional and which takes consciousness to be some sort of unextended, non-spatial property that likely eludes currently available scientific explanations. The resulting ‘naturalistic dualism’ Chalmers defends is Cartesian at its core, and despite his claim that such a position is ‘entirely compatible with the scientific view of the world,’ dualisms have fared badly as science proceeds to unify our conception of humankind in nature. Another factor strengthening the intuition that qualia emerge from physical or func- tional processes is the supposition that there most likely exists a clear demarcation between those sorts of states likely to ‘carry’ phenomenal consciousness and those that don’t. Qualia could well be absent in systems that are as smart as we are, but that are FUNCTION AND PHENOMENOLOGY 49 differently organized or instantiated, or so it is often supposed. (See, for instance, Flanagan, 1992, pp. 129-52 for a recent defence of this thesis.) Block’s Chinese nation and Searle’s Chinese room thought experiments trade on the intuition (although it begs the question against functionalism) that bizarre cognitive systems, however functional in real time, couldn't possibly constitute subjects (Block, 1978; Searle, 1980). Once the wedge is driven between function and qualia (or, as in Searle’s paper, between function and intentionality) in the bizarre cases, the stage is then set to suppose that qualia are only a contingent accompaniment to cognition, even in systems functionally very similar to us. Likewise, it seems plausible to many that qualitative states are probably absent in systems that fall short of the functional complexity given to humans and (perhaps) the higher animals. Although the line of demarcation is obscure in both cases, the basic assumption is that if a system fails to be sufficiently like us in some respect — in its physical instantiation or functional design — then the chances are it’s not a subject; it doesn’t produce or give rise to qualitative states. Ifonly acertain class of cognitive or functional processes (those more or less like ours) are taken as plausible correlates of subjectivity, this reinforces the notion that something special about those processes produces subjectivity as an ‘effect,’ whether epipheno- menal or efficacious. Consciousness emerges only at a certain level of complexity, or only out of a certain type of functional design, or only as the product of a certain type of physical instantiation, Drawing a line in advance between systems which are thought to be likely candidates for consciousness and those that are not pushes us towards the picture of subjectivity as a contingent accompaniment to functional processes. There is, of course, a good deal of anthropocentrism in drawing the line where those convinced by absent qualia thought experiments have placed it. But this may appear a reasonable prejudice, given that it seems we can only be certain of the existence of qualitative states from the single example of human experience. Surely it would be irresponsible as a starting assumption to grant qualia to just any old cognitive system, say ahorsefly, a frog, or Hal of 200/. Perhaps. But there is a difference, I suggest, between this sort of methodological conservatism (which in general I support) and the uncritical acceptance of the ‘similarity to ourselves’ criterion as a benchmark for assigning con- sciousness to a system, To turn the issue around, why should we assume, from the one example of human subjectivity, that qualia are not present in dissimilar cognitive sys- tems, real or imaginary, that manage well in the world? It is initially at least as plausible, to paraphrase Paul Churchland, that any cognitive system operating at or near our level must contain a ‘computational-executive organization’ which would be a ‘home for qualitative states’ (Churchland, 1989, p. 38).' And more basically, why assume that consciousness ‘accompanies’ or is ‘produced by’ functional processes in the first place? In response to the first question, I propose that we remain agnostic about the possible subjectivity of non-human and less complex systems, otherwise the anthropocentric bias " Lycan makes somewhat the same point when considering systems that are functionally very much like us: ‘Is it really possible to imagine something’s sharing my entire many-leveled functional organization and still not being conscious in the way that I am?" (1987, p. 24 his emphasis). Of course one can imagine this, but the possibility of absent qualia in the face of functional near-equiva- lence need hardly be the default assumption. Despite intuitions to the contrary, there is no a priori conceptual barrier that separates experience from function. 50 T.W. CLARK may blind us to the true nature of consciousness. To the second, the focus of this paper, I propose likewise that we do our best to keep the issue open, not to assume at the outset the non-identity of function and qualia. Since the narrowing of candidates for qualitative subjectivity based on similarity to us suggests that (only a certain class of) cognitive functions generate qualia, it can be seen that staying agnostic about the subjectivity of other sorts of systems is a good way to keep this deeper issue open. Qualia may not be generated at all; they may simply be certain types of functional organization, the range of which may not be limited to what is familiarly human. A third factor motivating the ‘emergence’ picture of qualitative consciousness is that only by having its own, independent existence could it possibly play the important causal role in our lives that it seems to. Involved here are fairly deep and emotional issues of human autonomy and specialness, especially the fear that if consciousness is nothing over and above physically instantiated function, then we lose our privileged status as rational agents riding above the flux of brute causality. As persons, we tend to identify ourselves with our conscious capacities, and moreover tend to believe that we are in some sense in control of these capacities. Consciousness, conceived of as a product of a very restricted class of functional, cognitive processes (ours), generated by a very restricted class of physical objects (human brains), is what crucially distinguishes us from the rest of insensate, mechanistic nature. If it turns out that subjectivity and the sense of self is ‘merely’ function, then it becomes terrifyingly (for some) clear that no principled distinction may exist between us and a very clever robot, at least on the question of who has ‘inner states.’ To the extent that we want, unconsciously or consciously, to preserve our special status vis-a-vis the robot we may opt for a picture of subjectivity, including qualia, that preserves for it a central causal role and restricts it to creatures very much like us. This picture of consciousness is just what the doctor ordered to keep ‘creeping mechanism’ at bay. (Dennett 1990 makes this point, pp. 523-4.) But of course we must not let such fears prejudice our initial conception of conscious- ness or restrict our investigations. If it turns out (as I propose) that consciousness has no independent causal role over and above the functions which instantiate it, and if such functions could be realized in creatures (or artifacts) quite unlike us in some respects, then the implications of that for our personal and moral status must be dealt with as a separate issue. I suspect that this same worry also works to make Chalmers’ anti-reductionist approach to explaining consciousness an attractive alternative to standard scientific practice. After all, a reductive explanation of subjectivity, one which identifies it with a class of functional processes that are in principle realizable in a wide range of instantiations, obliterates having an inner life as the basis for a special human status, (Others remain, however.) There are good arguments against certain types of facile reductionism, of course, and I don’t mean to imply that language referring to belief, intention, feeling, and thought is in any sense eliminable. But as a scientific strategy for unifying knowledge, the reductionist impulse is hardly to be eschewed but rather to be encouraged, one would suppose, and reductionism does not diminish us when we become the objects of knowl- edge. To ‘reduce’ mental phenomena to functional processes via some plausibly evi- denced identification is, after all, not to eliminate them, but simply to redescribe them FUNCTION AND PHENOMENOLOGY 51 from a third person perspective. Why such a redescription might seem threatening is an interesting question for another time. Ml Having adduced some possible explanations (there are undoubtedly others) for why Chalmers* picture of consciousness has such currency, I wish now to defend an alterna- tive, one which I believe is a better candidate for our attention this early in the game. I will argue for the identity of function and consciousness — what I will call the functional identity hypothesis — and as the argument develops, further reasons for doubting that ‘qualia arise from or are produced by functional processes will come to light, (See Lloyd, 1992 for a recommendation to adopt such an identity hypothesis.) ‘As mentioned earlier, Chalmers has set out some of the preliminary argument for such a position in his paper, and he describes what in my view is the key characteristic of functions likely to instantiate consciousness: their role in representing informational content essential for the control of complex, adaptive behaviour.” The basic functional identity hypothesis is that qualitative experiences are what itis to be a set of multi-modal, discriminative, representational processes which deploy information for the control of behaviour. I use this extended form of identity expression (‘what it is to be’) quite deliberately in order to emphasize that we, as experiencing subjects, are instantiated by such processes. To make this point clear, note the similarity, and dissimilarity, of this expression to Nagel’s (1974) popular characterization of subjectivity as ‘what it is like to be’ a such-and-such. The difference between the two is not trivial. Nagel’s formulation sug- gests the notion of an inner life which the subject somehow witnesses, or has direct epistemic ‘access’ to, so that for instance it (and only it) could say what it was like to have a particular quale. Nagel’s central thesis about subjectivity is that this ‘first person perspective’ cannot be captured by science. By contrast, the formulation I wish to defend hinges crucially on the proposal that as subjects we are constituted by and identical to cognitive processes which themselves instantiate qualia, hence qualia are what it is for us to be these processes. Under this proposal we can't, finally, say what it’s like to have qualia since we don’t have a first person perspective on them: we don’t ‘have’ them at all, neither do they ‘appear’ to us, nor are we ‘directly acquainted’ with them. We, as subjects, exist as them. The ineffability of qualia, among their other properties, is thus a consequence of and explained by the functional identity hypothesis, and qualia generally are not beyond the reach of science. (Part IV will return to these points.) Ibelieve Chalmers is very much on the right track in what he calls his two ‘non-basic’ psychophysical principles, although his commitment to the emergence picture of con- sciousness prevents him from realizing their full explanatory potential. He calls these the principles of ‘structural coherence’ and ‘organizational invariance;” both concern the functional deployment of information in a cognitive system. Structural coherence amounts to the well-established, if not yet completely fleshed out, correspondence 2 Several other philosophers have taken a similar tack, although they still by and large assume some version of the emergence picture. See for instance Flanagan, 1992, pp. 129-52; Van Gulick, 1993 pp. 152-3, and Van Gulick 1980 for discussions of the role of information in conscious processes. 52 T.W. CLARK between variable features within a sense modality and the empirically associated covari- ant neural structure and processing. For instance, the structure of our phenomenal experience of colour corresponds to the structure of the neural state space of our colour processing system (as described by Paul Churchland, 1989, pp. 102-10). There is at least a rough isomorphism between the sort of variability we report in our experience and the organization of neural events associated with this variability. Chalmers theorizes (correctly, I think) that the raison d'etre of cognitive functional processes is to embody informational content which ‘is brought to bear in a widespread way in the control of behaviour.’ Consciousness, it seems, is usually involved with most, if not all, higher order cognition and behaviour, including memory, anticipation, speech, learning, planning, and complex motor activity. As Flanagan (1992) puts it ‘What is consciously accessible is primarily just what we have the most need to know about: conditions in the sensory environment, and past facts, and events’ (p. 134). Such knowledge, whether recognitional, propositional, or performative, requires some sort of representational system involving information about the world and the body. All in all, the connection between consciousness, information and behaviour, although not water- tight and admitting of exceptions, seems a fruitful line of investigation.? The functional identity hypothesis makes a strong claim about this connection: that subjectivity is constituted by those central representational processes which transform and enhance sensory information to the point where it normally dominates in the control of behaviour. (See Van Gulick, 1993, pp. 147-50 for some interesting ideas on what sorts of processes these might be.) Ifit is plausible that functional processes correlated with consciousness are primarily informational and behaviour controlling, and if the principle of structural coherence holds, (and so far it seems empirically the case that there exists isomorphism between experience and neural organization), it follows that the structure of qualitative experience mirrors the informational content essential to the control of complex behaviour. The burning question, however, is whether indeed qualia mirror the informational content of functionat processes as some sort of separate, parallel entities in consciousness, or whether qualia are this informational content, Chalmers says, in support of the first position, that “There are properties of experience, such as the intrinsic nature of a sensation . .. that cannot be captured in a structural description’ (p. 23). So although the structure of experience might well be isomorphic to functionally derived informational content, experience itself cannot simply be this content, since (he believes) it has additional, non-functional mental properties, such as an intrinsic phenomenal nature. Such mental properties are irreducible to physical or functional properties, which is to say we can’t explain them or redescribe them using only physical or functional predicates without leaving out something crucial. As Nagel (1986, p. 15) has expressed it, ‘The subjective features of conscious mental processes — as opposed to their physical causes 3 Blindsight experiments seem the obvious counterexamples, since the subject clearly has what Flanagan terms ‘informational sensitivity’ to objects in the blindfield without ‘experiential sensitiv- ity’ (@ reportable experience) of them (Flanagan, 1992, pp. 147-52). But the very limited sorts of behaviour possible with respect to blindfield objects actually highlights the centrality of conscious- ness in mediating most of our interaction with the world. These limitations, when compared with normal behaviour, suggest what the special informational functions that instantiate consciousness might be. See Van Gulick, 1993, pp. 147-50 for some speculations about candidate functions. FUNCTION AND PHENOMENOLOGY 53 and effects — cannot be captured by the purified form of thought suitable for dealing with the physical world that underlies the appearances’. The truth of this claim is, of course, the central bone of contention between physicalism and functionalism on one side, and substance and property dualism and dual aspect theories on the other. In introducing his second principle, that of ‘organizational invariance’, Chalmers says, “This principle states that any two systems with the same fine-grained functional organi- zation will have qualitatively identical experiences’ (p. 25, original emphasis). This seems to contradict his earlier claim that the intrinsic nature of sensory experience is beyond the reach of structural distinctions, that is, beyond what the organizational structure of neural processes can account for. Indeed, organizational invariance comes perilously close, it seems, to identifying quality with function (same function, same quality) except that Chalmers continues to insist on the ‘emergence of experience’ as some sort of entity separate from functional organization. Chalmers’ passing observation that the principle of organizational invariance is ‘con- troversial’ is a bit of an understatement, since it is less a principle than a major thesis, about the mind. Were it proved, then, as he puts it, ‘the only physical properties directly relevant to the emergence of experience [would be] organizational properties’ (p. 26, original emphasis). If one were to drop the notion of emergence then this would pretty much amount to the hypothesis I want to defend, Experience is up-and-running functional organization, not something emerging out of it; qualia just are what it is 10 be a subject instantiated by a set of informationally rich, discriminative, behaviour-controlling proc- esses, not a separate ontology of intrinsic natures. Note that in this purified form, such a hypothesis explains the principle of structural coherence. That is, the reason the structure of conscious experience so nicely mirrors the informational goings on of neural processes is because they are one and the same, under different descriptions, or very loosely speaking, from different perspectives, first and third person. (As the reader will have noted above, I question the literal accuracy of the expression ‘first person perspective” as related to one’s experience.) Once we discard the idea of emergence the full explanatory power of the functional identity hypothesis starts to become apparent. And working from the other direction (as I will in part IV below), appreciating the extent to which the identity thesis can account for various mental property ascriptions will make it easier to abandon the picture of emergence as explana- torily superfluous. Chalmers’ third, ‘basic’ and highly speculative principle sketches a deeper explanation for both structural coherence and organizational invariance, but in the end it seems less explanatory than ontologically inflationary. On this principle information ‘is truly funda- mental, and . . . has two basic aspects, corresponding to the physical and phenomenal features of the world” (p. 27). The parallel emergence of phenomenal and physical properties, then, is attributable to a yet deeper substratum of information. But, to return to an earlier point, it seems premature to hypothesize new fundamental entities (in this case universal ‘bits’) at the start of an investigation when simpler, less inflationary hypotheses might do as well. Positing an underlying basis of information with two aspects may explain the congruence of experience with informationally contentful neural organization, and may support the notion that similar, information-bearing functions might generate similar phenomenal qualities, but we are then left with the equally 54 T.W, CLARK difficult task of explaining why information should possess two aspects to begin with, It doesn’t help to ‘solve’ one explanatory problem by creating yet another at a more speculative level, at least not until more economical approaches have been thoroughly explored. It is also worth noting that with this proposal Chalmers has shifted from recommending that we accept experience as fundamental (p. 20) to speculating that information is fundamental (pp. 26~7). Either way, all the explanatory work still remains to be done, Iv Thave already suggested how the hypothesis that qualia are identical with higher level functional processes can explain structural coherence: the parallel structure of experience and neural organization is accounted for by supposing, parsimoniously, that experience is to exist as certain types of neural processes. Likewise, the principle of organizational invariance is simply a corollary of the identity hypothesis: two functionally equivalent systems will have the same sorts of qualitative experience because qualia are particular informational values within some sort of functional state space, however it may be realized. (Below I will amend this account of phenomenal quality by questioning the reality of the intrinsic, essential nature of qualia.) But these explications will likely seem beside the point for those whose primary intuition about consciousness is precisely that, as Chalmers insists, experience is something over and above functional processes. To undercut this intuition, and thus make functional identity more appealing as an initial explanatory hypothesis, it will be helpful to consider some of the properties normally attributed to qualia: their privacy, ineffability, and intrinsic nature. The privacy of qualia, the fact that no one but myself can feel the pain that I feel, is relatively easy to explain as a matter of identity. Flanagan describes it thus: ‘.. a particular realization (of an experience] will be an experience only for the agent who is causally connected to the realization in the right sort of way . . . (T]he biological integrity of the human body can account straightforwardly for the happy fact that we each have our own, and only our own experiences’ (p. 94, original emphasis). T would modify Flanagan's account somewhat: the reason no one else feels my pain, even though they might conceivably have a complete description of its neural instantia- tion, is that only I as a subject consist (partially) of that pain. There is no separate agent or subject apart from the sum total of my experience which could be causally connected to pain in the first place, so naturally no one else could stand in that relation to my experience either. That experience is private is thus explained not as a matter of direct epistemic access but as a matter of instantiation: subjectivity consists of a complex array of coordinated, information-bearing neural states, and as a subject I consist of those states, and no one else can so consist. That I can report being in pain is not, therefore, because I perceive pain (as I might perceive a tree or chair) by virtue of a special causal connection to it, but because I subjectively exist as pain — as stimulated C fibers and associated higher level processing — to some extent. (Under torture I might subjectively exist mostly as pain and the concomitant terror.) Why, under the assumption of functional identity, might qualia be ineffable? Very much the same logic applies. The subject, since it consists of an ongoing stream of FUNCTION AND PHENOMENOLOGY Ea neurally instantiated experience, is not in a position to witness or observe the basic elements of that experience. We cannot, as it were, step back from and describe a quale as we might an external object; thus we can do no more than name basic qualitative experiences (‘red’, ‘hot’, ‘sweet’, ete.) and compare and contrast them to one another. We can't describe the redness of red or painfulness of pain precisely because we can’t get a perspective on these qualities. (Again, it is a mistake to suppose that we have a first person point of view of our experience.) They are, so to speak, the counters in the game of perception and so cannot be made the object of play. External particulars, on the other hand, and complex internal states, are describable just because they are constituted by ensembles of qualitative elements. Chairs are (sometimes) brown, assume a given shape in my visual field, are hard to the touch, resonant when struck, etc. That the subject is identical to a set of representational processes can thus explain why the primitive components of that set, what we call qualia, are descriptively opaque and non-decompos- able, that is, ineffable: we can’t stand in an epistemic relation to those representational states which we consist of as knowers. This brings us to the last, and perhaps most central (and controversial) of properties attributed to qualia, the notion that they have an intrinsic phenomenal nature. It is here, many suppose, that functionalist theories must founder, since the essential redness of red or painfulness of pain seem to float free of any functional role. We can imagine red objects looking blue (or imagine they look blue to someone else) without supposing that our (or the other person’s) discriminative and behavioural capacities would be in any sense compromised. The essence of red, just because it is a non-decomposable, ineffable primitive, is thought not to be amenable to functional or structural explanation. As Van Gulick puts it: ‘No matter how much structural organization we can find in the phenome- nal realm and explain neurophysiologically, [the critic of functionalism] will insist that the distinct redness of phenomenal red will not have been captured or explained by our theory’ (1993, p. 145). The purported intrinsic phenomenal nature of qualia creates the explanatory gap functionalism is supposed not to be able to fill. ‘The functional identity hypothesis can, however, meet this burden by supporting a challenge to the intuition that there are determinate phenomenal facts about the intrinsic nature of qualia in need of explanation. The challenge, simply put, is to specify what precisely is being claimed to be the case when we point to a patch of red and say our experience is like ‘that.’ Since there is nothing further we can say about the (purported) phenomenal essence of red, there is no way we can link such an essence with a property or state of affairs which specifies or fixes its occurrence, other than to gesture at red objects. Hence, referring to such an essence plays no explanatory or descriptive role in talk about our experiences, and this may begin to cast doubt on whether such an essence exists, even though it may strongly seem that it does. The ineffability of qualia, it turns out, could be a clue to their not having a determinate intrinsic nature after all. ‘The intuition that experiences do indeed have some sort of qualitative essence supports, the possibility of inverted or ‘alien’ qualia, that the intrinsic character of an experience might differ across subjects. My red might be the same as your red, or perhaps not. A Martian’s ‘red’ might even be a radically different sort of experience. But since experi- ence is private, intersubjective comparisons of qualia are of course impossible. If we can never tell, finally, whether another subject’s red is the same or different than mine, this 56 T.W. CLARK might make us further doubt the validity of the notion of intrinsic phenomenal natures. Why, after all, should we take seriously a ‘fact’ — the ‘fact’ that my experience of red is possibly like or unlike yours — which is in principle impossible to ascertain? Both the ineffability and privacy of qualia, therefore, undermine the plausibility of first person phenomenological facts involving determinate qualitative essences which science cannot capture. ‘That we can reliably distinguish red from other colours isn’t explained by there being an essence of red, a particular quality of redness that, for instance, we might suppose other perceivers of the same red object might or might not be experiencing. The ability to pick out red depends simply on the fact that there exists a range of contrasting colours against. which red is distinguished as a relationally defined member of that range. This is borne out by the obvious point that every distinguishable bit of red we see is experienced against a background of non-red. If all the world were red, red would drop out as a discriminable property of experience. Imagine Mary, Jackson’s colour-deprived neuroscientist, grow- ing up in an environment where everything was red instead of (as originally conceived in the thought experiment) black and white (Jackson, 1982). Could she have a concept of ‘first person’ red? Not until, I suggest, other colours were introduced into her environ- ment against which red could be reliably distinguished. A monochromatic world of whatever colour is a phenomenally colourless world, not a world in which the single colour could ‘declare’ itself by an intrinsic phenomenal nature. A functionalist account of our sensory capacities makes plausible the relational, mutually defined nature of qualitative experience, hence supports the attack on the existence of intrinsic phenomenal essences. As the Churchlands have theorized, sensory state spaces, as realized in neural organization, instantiate a range of possible sensory-in- formational values within a given modality (vision, hearing, taste, etc.), values which correspond more or less to some range of stimulation from the external world and the body. What distinguishes different values of a given modality is variation along one or more of its component dimensions (three in the case of vision, four in the case of taste), a variation that defines each state of the modality in relation to its other states. These states gain their functional significance by virtue of mapping differences in stimulation originating from the external world or the body. Hence it is the difference between the neural instantiation of red and the neural instantiation of blue (among other colours) which defines them as distinct qualitative experiences, not that red is instantiated by a particular state that by itself defines it as red or that blue is instantiated by a particular state that by itself defines it as blue. If, as I hypothesize, we as subjects exist as such sensory modalities (along with other sorts of cognitive processes) the experience of a particular shade of red is just one in an array of possible states of the colour system that gets its qualitative value solely by having a particular place in the array. The ‘way’ red is, what it is ‘like,’ is simply to exist as a given colour state in contradistinction to other states, There is no need to posit an intrinsic nature of red as separate from its functional role as a placeholder in the neural processing that constitutes the experience of colour, and that, partially, constitutes us as subjects.* * If one wishes to retain the notion of intrinsic qualitative nature but deny that this defeats functionalism, Paul Churchland offers a way out: admit that such natures exist but insist that ‘such intrinsic natures are nevertheless not essential to the type-identity of a given mental state, and may FUNCTION AND PHENOMENOLOGY 57 ‘Conclusions ‘The foregoing analysis is meant to help undercut the intuition that there are phenomenal facts in need of explanation that cannot be captured by functional facts. The explanatory ‘gap is closed by showing both that the privacy and ineffability of qualia are indeed explained by the functional identity hypothesis, and that their much vaunted intrinsic qualitative nature does not exist, hence needs no explaining. This helps to support the view that qualia are not phenomenal entities that emerge from or arise out of functional processes, but are instead best conceived of as being those processes. This is not to ‘quine’ qualia out of existence, but simply to identify them with certain sorts of neurally instantiated representational states. Of course, the meaning of the terms ‘phenomenal consciousness” and ‘qualia’ is not (presently) identical to the meaning of the expression ‘neurally instantiated representational states’ and its functionalist and physicalist cous- ins, But talk about qualitative experience refers to the same thing as talk about particular neural processes, although precisely which processes we don’t yet know. Thave tried to emphasize the explanatory virtues of the functional identity hypothesis, hoping to win converts to functionalism by showing that it can indeed account for subjectivity. Yet there still might seem a residual, fatal question left unaddressed (the frog grinning up at us from the bottom of Austin’s beer mug): why should existing as functional processes of a particular kind be identical to qualitative experience? What is it about them that makes them phenomenally conscious, as opposed to unconscious? (This, it will be seen, is Chalmers’ original, central question, but posed about identity instead of emergence.) Well, if by ‘phenomenally conscious’ we insist on meaning “possessed of a first person, intrinsic, essential nature’ then the question stands unan- swered, since nothing in functionalism or science will ever show us how to get from physics, chemistry, biology, and cybernetics to consciousness thus defined. The causal and structural aspects of subjectivity might fall to functional explanation, but intrinsic natures will not, since after all, these are custom made to resist assimilation by science. Why? Because scientific explanation works, in part, by showing how the causal relations among elements of a system at one level can account for features at another level. The macroscopic properties of water, for instance, can be explained by the microstructure of water molecules, But since the property of having an intrinsic qualitative nature is defined (by its adherents) as independent of any causal role and as having no structure, no lower level explanation will ever reduce it. As Levine (1993, p. 134) puts it, ‘[T]o the extent that there is an element in our concept of qualitative character that is not captured by features of its causal role, to that extent it will escape the explanatory net of physicalist reduction.’ Thave tried to cast doubt on the conviction that our concept of qualia need include the property of having a factually determinate intrinsic nature independent of function or indeed vary from instance to instance of the same type of mental state’. (1988, p. 39) Of course some would object, for example, that Martian pain couldn't be radically different in its intrinsic nature from our pain and still count as pain. But, the reply comes back, if it played the same functional role why wouldn't we call it pain? My point to both sides, however, is that the impossibility of intersubjective qualia comparison renders this debate undecidable, hence irrelevant. See Shoemaker (1991) for a defence of intrinsic quality compatible with functionalism and Dennett (1990, pp: 538-44) for arguments against the existence of intrinsic properties of qualia. 58 T.W. CLARK physical instantiation, It is logically possible that qualia so exist, but there are no good empirical reasons, nor for that matter any good phenomenological reasons, for supposing they do. There are good reasons, on the other hand, for supposing that qualia are instead a matter of the contrast relations among functionally defined, neurally realized repre- sentational processes. If such is the case, then the conscious/unconscious distinction lies in whether or not such processes are instantiated. Discovering just which processes constitute phenomenal consciousness is an empirical matter that only neuroscience can answer, by correlating neural states with various conscious capacities. ‘There may or may not be a clear functional or physical line to be drawn between the conscious and the unconscious, but when we ask why a process is conscious we will no longer be wondering why qualitative essences accompany particular neural events. Rather, under the identity hypothesis, we will be asking why that process needs to involve the functions typically operational when we are conscious. The beating of my heart or the (mostly) silent operation of my digestive system obviously need not involve conscious processes, since neither requires the representation of a complex environment, or moti- vational economy, or set of future contingencies. Short-term memory, on the other hand, is a conscious process because it requires those representational capacities only normally available when we are conscious, Some sort of qualitative experience (whether of occurrent thought, feeling, or perception) is necessary as input to the short-term memory system, and such experience is likely to be empirically cashed out as a complex set of discriminative and integrative functions which manage information that serves to control moment-to-moment behaviour. The functional identity hypothesis thus explains why we have subjective experience: qualitative representations are what get the job done. The functional identity hypothesis proposes no new fundamental classes of entities or properties or features to be added to naturalistic theory, and so has the virtues of simplicity, ontological parsimony, and methodological conservatism. If it turns out that consciousness depends fundamentally on a physical process, e.g., a phase locked 40 hertz neural oscillation, or if, as I suspect, it becomes identified with a rather sophisticated array of informational processes which could be realized by many different physical systems, in neither case will anything radically new be discovered to play an explanatory role (although figuring out the actual workings of the massively recursive networks embodied by the brain may reveal vast new realms of control theory), Of course it is remotely possible that a heretofore unknown fundamental property or entity, whether of experience or information or quantum coherence, may eventually be shown to exist. But asserting its probable existence in advance of strong empirical support and theoretical necessity would be to start off on the wrong foot. If indeed qualia are identical to certain functional informational processes, then the hard problem of consciousness (what Chalmers more or less takes to be the easy problem) is the empirical one of discovering just what these processes are. Equally hard for many, perhaps, will be dropping the picture that consciousness is something ontologically special produced by these processes. Instead, we must get used to the idea that as much as our ‘inner’ lives seem a categorically different sort of thing than the external world (a world which includes our brains), they are, in fact, just more of that external, physical world which we as subjects happen to be. We must also get used to the idea that consciousness does not have a cognitive or functional role over and above the functions FUNCTION AND PHENOMENOLOGY 59 which constitute it, and that any system, artificial or natural, which instantiates such functions will be conscious. Inasense there is nothing special about consciousness since, if Iam right, nothing extra is produced by or emerges out of this set of functions. They just constitute a marvelously complex and adaptive representational system which keeps us out of trouble, more or less. Yet on the other hand, consciousness is indeed special in that these functions, as carried out in our day to day lives, have, as a historical fact, led most of us to suppose that an ontologically separate world of subjectivity exists. It may be that the sorts of higher level cognitive processes which are found to correlate with consciousness inevitably generate (among language users) a self/world model containing the strong intuition that the self and its experience cannot simply be the body, cannot simply be a bit of the world suitably organized. Explaining consciousness satisfactorily will consist in overcoming that intuition, and in placing experience fully within the natural, physical realm.> References Block, Ned (1978), “Troubles with functionalism’, Minnesota Studies in the Philosophy of Science, IX, ed. C. Wade Savage (Minneapolis: University of Minnesota Press). Reprinted in Rosenthal 1991, pp. 211-28. Chalmers, David (1995), “Facing up to the problem of consciousness’, Journal of Consciousness ‘Studies, 2 (3), pp. 200-19 (reprinted in this volume). Churchland, Paul M. (1988), Matter and Consciousness (Cambridge, MA and London: MIT Press). Churchland, Paul M. (1989), A Newracomputational Perspective (Cambridge, MA and London: MIT Press). Cole, David (1994), “Thought and qualia’, Minds and Machines, 4, pp. 283-302. Crick, Francis (1994), interview, Journal of Consciousness Studies, I, pp. 10-17. Davies, M. and Humphreys, G.W. (eds.1993), Consciousness (Oxford and Cambridge, MA: Blackwell). Dennett, Daniel (1987), The Intentional Stance (Cambridge, MA and London: MIT Press). Dennett, Daniel (1990), ‘Quining qualia’, in Consciousness and Contemporary Scienceed. A. ‘Marcel and E. Bisiach (Oxford: Oxford University Press). Reprinted in Lycan 1990 pp. 519-47. Flanagan, Owen (1992), Consciousness Reconsidered (Cambridge, MA and London: MIT Press). Jackson, Frank (1982), ‘Epiphenomenal qualia’, Philosophical Quarterly, 32, pp. 127-36. Levine, Joseph (1993), ‘On leaving out what it's like’, in Davies and Humphreys 1993, pp. 121-36. Lloyd, Daniel (1992), “Toward an identity theory of consciousness’, Behavioral and Brain Sciences, 15, pp. 215-16. Lycan, William G. (1987), Consciousness (Cambridge, MA and London: MIT Press). Lycan, William (ed. 1990), Mind and Cognition (Cambridge, MA: Blackwell). Nagel, Thomas (1974), “What is it like to be a bat?", The Philosophical Review, LXXXIIL, 4, pp. 43-50. Nagel, Thomas (1986), The View From Nowhere (New York and Oxford: Oxford University Press). Rosenthal, David M. (ed. 1991), The Nature of Mind (New York: Oxford). Searle, John (1980), ‘Minds, Brains, and Programs’, Behavioral and Brain Sciences, 3, pp. 417-57. Shoemaker, Sydney (1991), ‘Qualia and consciousness’, Mind, C, pp. 507-24. Van Gulick, Robert (1980), “Functionalism, information and content’, Nature and System, 2, Reprinted in Lycan 1990, pp.10-129. Van Gulick, Robert (1993), ‘Understanding the phenomenal mind: are we all just armadillos?”, in Davies and Humphreys 1993, pp. 137-54. > Tam grateful to Mary Ellen Myhr for a close reading of an earlier draft of this paper, and to the referees for helpful criticisms from which the present version has benefitted. The paper was originally published in the Journal of Consciousness Studies, 2, No.3 (1995), pp. 241-54. Valerie Gray Hardcastle The Why of Consciousness: A Non-lssue for Materialists In my (albeit limited) experience of these matters, I have discovered that there are two sorts of people engaged in the study of consciousness. There are those who are committed naturalists; they believe that consciousness is part of the physical world, just as kings and queens and sealing wax are. It is completely nonmysterious (though it is poorly under- stood). They have total and absolute faith that science as it is construed today will someday explain this as it has explained the other so-called mysteries of our age. Others are not as convinced, They might believe that consciousness is part of the natural world, but surely it is completely mysterious (and maybe not physical after all). Thus far, science has little to say about conscious experience because it has made absolutely no progress in explaining why we are conscious at all. Different sceptics draw different morals from their observation. Some conclude that a scientific theory of consciousness is well-nigh impossible; others believe that it is possible, but do not expect anything of value to be immediately forthcoming; still others remain confused and are not sure what to think. (Perhaps unfairly, I put David Chalmers in the last category, as he remarks, ‘Why should physical processing give rise to a rich inner life at all? It seems objectively unreasonable that it should, and yet it does.” (p. 11.) His intuition that consciousness is too bizarre to be real, yet still exists anyway illustrates the sentiments of the third category quite nicely. Further, as I discuss below, I think his tentative programme of redoing our basic scientific ontology reflects some basic confu- sions on his part.) Thave also noticed that these two camps have little to say to one another, for their differences are deep and deeply entrenched. I can’t say that I expect to change that fact here. I fall into the former camp. 1 am a committed materialist and believe absolutely and certainly that empirical investigation is the proper approach in explaining consciousness. also recognize that I have little convincing to say to those opposed to me. There are few useful conversations; there are even fewer converts. In this brief essay, I hope to make clearer where the points of division lay. In the first section, I highlight the disagreements between Chalmers and me, arguing that conscious- ness is not a brute fact about the world, In section Il, I point out the fundamental difference between the materialists and the sceptics, suggesting that this difference is not something that further discussion or argumentation can overcome. In the final section, I ‘This paper was originally published in the Journal of Consciousness Studies, 3, No.1 (1996), pp. 7-13. 62 V.G. HARDCASTLE outline one view of scientific explanation and conclude that the source of conflict really tums on a difference in the rules each side has adopted in playing the game. I In large part, these divergent reactions turn on antecedent views about what counts as explanatory. There are those who are sold on the programme of science. They believe that the way to explain something is to build a model of it that captures at least some of its etiologic history and some of its causal powers. Their approach to explaining conscious- ness is the same as mine: isolate the causal influences with respect to consciousness and model them (cf. Churchland, 1984; Flanagan, 1992; Hardcastle, 1995; Hardin, 1988). In contrast, others (e.g. Block, 1995; Chalmers, 1995; McGinn, 1991; Nagel, 1974; Searle, 1992) do not believe that science and its commitment to modelling causal interactions are necessarily the end-all and be-all of explanation. They believe that some things — many things — are explained in terms of physical causes, but qualia may not be. Isolating the causal relations associated with conscious phenomena would simply miss the boat, for there is no way that doing that ever captures the qualitative aspects of awareness. What the naturalists might do is illustrate when we are conscious, but that won't explain the why of consciousness. The naturalists would not have explained why it is neuronal oscillations (cf. Crick and Koch, 1990), or the activation of episodic memory (cf. Hardcastle, 1995), or an executive processor (cf. Baars, 1988), or whatever, should have a qualitative aspect, and until they do that, they cannot claim to have done anything particularly interesting with consciousness. To them, I have little to say in defence of naturalism, for I think nothing that I as an already committed naturalist could say would suffice, for we don't agree on the terms of the argument in the first place. Nevertheless, I shall try to say something, if for no other reason than to make the points of disagreement clearer so that informed buyers can chose all the more wisely. Let me sketch in particular the point of conflict between Chalmers and me. Let us assume a prior and fundamental commitment to materialism. I say that if we are materialists, then we have to believe that consciousness is something physical. Presum- ably it is something in the brain. If we believe this and we want to know what conscious- ness is exactly, then we need to isolate the components of the brain or of brain activity that are necessary and sufficient for consciousness. If I understand Chalmers’ taxonomy of research programmes correctly, then I am advocating following option five: ‘isolate the substrate of experience’. Indeed, it is my contention that pointing out the relevant brain activity conjoined with explaining the structure of experience (his option four) and some functional story about what being conscious buys us biologically (not one of Chalmers’ options) would be a complete theory of consciousness. Let us pretend though that I have only completed the first step in this programme and have isolated the substrate of experience. Call this component of the brain C. Chalmers would reply that though I might have been successful in isolating the causal etiology of consciousness, I have not explained why itis that C should be conscious. Why this? For that matter, why anything? Part of a good explanation, he maintains, is making the identity statement (or whatever) intelligible, plausible, reasonable. I have not done that. Hence, I have not explained the most basic, most puzzling, most difficult question of consciousness. I haven’t removed the curiousness of the connection between mind and body. I haven't closed the explanatory gap. THE WHY OF CONSCIOUSNESS: A NON-ISSUE FOR MATERIALISTS 63 How should I respond? He is, of course, exactly right: scientific theories of conscious- ness won't explain the weirdness of consciousness to those who find the identity weird. One possible move is to claim that consciousness just being C (or whatever theory you happen to believe) is just a brute fact about the world. That is just the way our universe works. At times, I am sure, it appears that this is what the naturalists are assuming, especially when they dismiss out of hand those overcome by the eeriness of conscious- ness. This, too, is what Chalmers wants to do with his dual aspect theory: phenomenal qualities are just part and parcel of information. No further explanation needed. However, this response is too facile. It is true that we accept brute facts about our universe. We believe in things like gravitational attraction and the electromagnetic forces without question. We waste little energy wondering why our universe contains gravity. It just does, and we reason from there. On the other hand, there are other facts about the world that we do not accept as brute. We feel perfectly comfortable expecting an answer to why water is wet. That is not a brute fact. We explain the liquidity of water by appeal to other facts about the world, the molecular structure of water and its concomitant microphysical properties, for example. And these facts are explained in tun by other facts, such as the quantum mechanical structure of the world. Now these might be brute facts, but so it goes. (At least this is one popular and rosy view of scientific unity. Ishan’t defend that here.) Notice two things. First, the facts we accept as brute are few and basic. Essentially, we accept the most fundamental elements and relations of the universe as given. The rest then depend upon these key ingredients in some fashion. Second, and following from the first observation, it seems highly unlikely that some relatively chauvinistic biological fact should ever be brute. For those facts turn on the more fundamental items in the universe. Hence, if one is to claim that consciousness being C is simply a brute fact about the universe, then one is prima facie operating with a perverse metaphysics. Chalmers tries to overcome the latter difficulty by denying that consciousness is biological. However, he has no reason to claim this except that it saves his theory. Considerations of structural coherence and organizational invariance aren’t telling be- ‘cause they are generally taken to support material identity. That is, if you find structural isomorphisms between our perceptions and twitches in the brain, then that is taken to be good reason to think that the mind is nothing more than activity in the brain. (What other sort of evidence could you use?) And if you hypothesize that the same ‘fine-grained’ functional organization supports the same phenomenal experiences, then you are advo- cating some sort of materialistic functional theory; otherwise the perceptions can diverge even though the functional organization remains the same (cf. Shoemaker, 1975; 1981; see also Lycan, 1987).! ‘The only consideration he brings to bear is the putative ‘elegance’ of a dual aspect theory. However, when we weigh a suggestion’s simplicity and elegance against coun- "1 find it strange (though not inconsistent) that in the first portion of the paper, Chalmers uses the putative imaginability of inverted qualia as an argument against what he calls ‘reductionism’ (though to me it is simply a good old fashioned identity theory), yet in discussing constraints on possible theories he argues against the possibility of inverted qualia in support of his proto-theory. He should recognize that if the fine-grainedness of his functional organization is fine-grained enough, then we would be discussing the functional organization of neurons (or action potentials, IPSPs, EPSPs, or what have you), which is all one needs to muster a claim for mind—brain identity. 64 V.G. HARDCASTLE tervailing data, the data have to win. We already know that not all information has a phenomenal edge to it, insofar as we know quite a bit of our information processing is cartied out unconsciously. Documenting subliminal effects, implicit priming, and re- pressed but effective memories are all cottage industries in psychology, and have been since Freud,? Chalmers is either going to have to deny some of the most robust psycho- logical results we have and claim that no information processing is occurring in those cases, or do a “bait-and-switch’ and claim that, contrary to introspective verbal reports, we are conscious of all of those things (we just don’t realize it). Neither option is plausible. Chalmers gives us no counter-examples to the mass of psychological evidence, and denying that first person viewpoints can tell us whether we are conscious denies exactly what Chalmers wants to defend. Hence, we are left with the prima facie plausible claim that for all cases of consciousness of which we are aware, consciousness is biological. In any event, I don’t want to make the claim consciousness is brute. So what do I say if [think that consciousness is a biological phenomenon? How do I make my identifica- tion of consciousness with some neural activity intelligible to those who find it mysteri- ous? My answer is that I don’t. The ‘solution’ to this vexing difficulty, such as itis, is all a matter of attitude. That is, the problem itself depends on the spirit in which we approach an examination of consciousness, pif Let us return to the example of water being wet. Consider the following exchange. A water-mysterian wonders why water has this peculiar property. She inquires and you give an explanation of the molecular composition of water and a brief story about the connection between micro-chemical properties and macro-phenomena. Ah, she says, I am a materialist, so I am convinced that you have properly correlated water with its underlying molecular composition. I also have no reason to doubt that your story about the macro-effects of chemical properties to be wrong. But I still am not satisfied, for you have left off in your explanation what I find most puzzling. Why is water HO? Why couldn't it be XYZ? Why couldn’t it have some other radically different chemical story behind it? I can imagine a possible world in which water has all the macro-properties that it has now, but is not composed of H,0. Of course, people like Kripke have a ready response to the water-mysterians. ‘Water = HO" is an identity statement. Hence, you can’t really imagine possible worlds in which water is not HO because you aren’t imagining water in those cases (or, you aren't imagining properly). As Chalmers would claim, it is a conceptual truth about water that itis HO. But, to the sceptical and unconvinced, to those who insist that they can imagine honest-to-goodness water not being HzO, what can one say? I think nothing. Water- mysterians are antecedently convinced of the mysteriousness of water and no amount of ? take it that these facts are well known, I summarize quite a bit of this research in Hardcastle (1995). Aside from Freud, other important players include Endel Tulving, George Mandler, Anthony Marcel and Daniel Schacter. 3 Note that claiming that consciousness is biological does not mean that we could not create consciousness artificially. Life is a biological phenomenon too, but that doesn’t rule out creating life in test-tubes. THE WHY OF CONSCIOUSNESS: A NON-ISSUE FOR MATERIALISTS 65 scientific data is going to change that perspective. Bither you already believe that science is going to give you a correct identity statement, or you don't and you think that there is always going to be something left over, the wateriness of water. I doubt there are any such mysterians, so perhaps this is a silly example. Let us now turn to life-mysterians. Consider the following exchange. A life-mysterian wonders why living things have the peculiar property of being alive. She inquires and you give a just-so story about the origin of replicating molecules in primordial soup and wave your hands in the direction of increasing complexity. Ah, she says, I am an evolutionist, so I am convinced that you have properly correlated the history of living things with their underlying molecular composition. I also have no reason to doubt that your story about increasing complexity to be wrong. But I still am not satisfied, for you have left off in your explanation what I find most puzzling, the aliveness of life. Why couldn’t that be a soul? Why couldn't it have some other radically different evolutionary story behind it, namely, one with God in it? I can imagine a possible world in which living things have all the macro-properties that they have now, but are not comprised of DNA or RNA. Of course, as Chalmers indicates, we too have a ready response to the life-mysterians, We presume that there is some sort of identity statement for biological life. (Of course, we don’t actually have one yet, but for those of us who are not life-mysterians, we feel certain that one is in the offing.) Hence, they can’t really imagine possible worlds in which life is not whatever we ultimately discover it to be because they aren’t imagining life in those cases (or, they aren’t imagining properly). But, that aside, what can we say to those who insist that they can imagine life as requiring an animator? I think nothing. Just getting on with the biological enterprise is perhaps appropriate. Life-mysterians are antecedently convinced of the mysteriousness of life and no amount of scientific data is going to change that perspective. Either you already believe that science is going to give you a correct identity statement, or you don’t and you think that there is always going to be something left over, the aliveness of living things. So what about Chalmers and other consciousness-mysterians? They are no different. They are antecedently convinced of the mysteriousness of consciousness and no amount of scientific data is going to change that perspective. Either you already believe that science is going to give you a correct identity statement, or you don’t and you think that there is always going to be something left over, the phenomenal aspects of conscious experience, ‘Experience . . . is not entailed by the physical.’ Chalmers wants to know: “Why is the performance of these [cognitive] functions accompanied by experience?’ (p. 12; emphasis mine). Though he does believe that ‘experience arises one way or another from brain processes,’ he thinks that itis a ‘conceptual point’ that consciousness is not identical to C. In some sense, of course, I have a ready response to the consciousness-mysterians. Like the water-mysterian and the life-mysterian, consciousness-mysterians need to alter their concepts. To put it bluntly: their failure to appreciate the world as it really is cuts no ice with science, Their ideas are at fault, not the scientific method, Materialists presume that there is some sort of identity statement for consciousness. (Of course, we don’t actually have one yet, but for those of us who are not consciousness-mysterians, we fee! certain that one is in the offing.) Hence, the sceptics can’t really imagine possible worlds in which consciousness is not whatever we ultimately discover it to be because they aren’t 66 V.G. HARDCASTLE imagining consciousness in those cases (or, they aren't imagining properly). But never- theless, what can I say to those who insist that they can imagine consciousness as beyond science’s current explanatory capacities? I think nothing, for they can claim that I am conceptually confused as well. Agreeing to disagree is perhaps appropriate. I suppose we have reached a stand-off of sorts. I say materialism and mechanism entail an identity statement for consciousness, just as we get one for water and we expect one for life. Consciousness is no more mysterious to me than the wetness of water or the aliveness of life. That is to say, I find all of the phenomena interestingly weird, and the identity statements that science produces marvelously curious. But all are on a par. The sceptics do not share my intuitions. So be it. However, I feel no more inclined to try to convince them otherwise than I do trying to convince the religious that souls don’t exist. I recognize hopeless projects, Our antecedent intuitions simply diverge too much to engage in a productive dialogue. mM But perhaps again I am not being fair. The reason water-mysterianism seems implausible is that we are able to embed our understanding of water and HO in the sophisticated larger framework of molecular chemistry and sub-atomic physics. We just know an awful lot about how atoms and molecules interact with one another and the corresponding micro- and macro-properties, Life-mysterianism seems implausible to those for whom it seems implausible for similar reasons. We don’t know as much about biological history as we do about molecular chemistry, but we do know enough at least to gesture toward a suitable framework in which to embed a decomposition of life. But consciousness might be different. We have far, far to go before we can claim to understand either cognitive or brain processes with any surety. Pethaps there just isn’t a suitable larger framework in which to embed an understanding of consciousness; hence, any scientific model we try to construct will appear strained and stilted at best. And perhaps this is what really drives the explanatory gap — we don't yet know what we are talking about when we claim that consciousness is a natural phenomenon. Suppose this argument is correct (though I am not sure that it is, for reasons I explain below). What follows from it? It can’t be that a theory of consciousness is not possible, nor even that consciousness is fundamentally odd, Rather, all we can say is that we have to wait and see what else we learn about the mind and brain before a decomposition and localization of consciousness can be intuitively satisfying. Consciousness might very well be C, but our informed intuitions lag behind. (An aside: Can we really say what would happen if my neural circuits are replaced by silicon isomorphs? Maybe it is reasonable to think that your experiences would not be affected. But, in the same vein, it is reasonable to believe that the world is Euclidean — though it isn’t, of course — and it used to be reasonable to burn witches at the stake — though it is no longer. What seems reasonable at first blush often isn’t once the para- meters of the problem are made sufficiently clear; moreover, our intuitions change as our perspective on the world changes. At present, we simply don’t know enough about the explanatory currency of the brain to hypothesize intelligently about what will happen if THE WHY OF CONSCIOUSNESS: A NON-ISSUE FOR MATERIALISTS 67 we push on it in various ways. Intuition pumps only work if we have robust and well-founded intuitions in the first place.) All we can say at this point is that an antecedent commitment to materialism means that an understanding of consciousness will someday be embedded in some larger mind-brain framework. We are just going to have to wait until that time before our intuitions concerning what counts as a satisfactory identification for phenomenal experience will be useful (or even usable). Nevertheless, though there is a great deal we don’t know about the mind and the brain, there is still a lot that we do. Indeed, within the broader framework of currently accepted psychological and neurophysiological theories, we have found striking parallels between ‘our phenomenal experiences and activities in the brain, Chalmers points to some in his paper; others are more basic. E.g. removing area MT is correlated with phenomenal blindness; ablations in various regions of cortex are correlated with inabilities to perceive shapes, colours, motion, objects; lesions surrounding the hippocampus are correlated with the loss of episodic memory. Or, for less invasive results, consider what happens when various chemicals are added to our brains. We decrease pain, increase sensitivity, induce hallucina- tions, alter moods, and so on. Data such as these should (someday) allow us to locate conscious experiences both within our information processing stream and within the head. Perhaps more data, better constructed scientific models, and more agreement among the scientists themselves about the details, would alter the intuitions of the sceptics, but I doubt it. For the difference between someone like Chalmers and me is not in the details; itis in how we understand the project of explaining consciousness itself. It is a difference in how we think of scientific inquiry and what we think explanations of consciousness are supposed to do. Explanations are social creatures. They are designed for particular audiences asking particular questions within a particular historically determined framework. (See van Fraassen, 1980, for more discussion of this point.) Materialists are trying to explain to each other what consciousness is within current scientific frameworks. Their explana- tions are designed for them. If you don’t antecedently buy into this project, including its biases, history, context, central questions, possible answers, and relevant actors, then a naturalist’s explanation probably won’ t satisfy you. It shouldn't. But that is not the fault of the explanation, nor is it the fault of the materialists. If you don’t accept the rules, the game won’t make any sense. If you do accept the rules, then the explanations will follow because they are designed for you as a member of the relevant community. (This is not to say that you will agree with explanations, just that they will seem to be of the right sort of thing required for an answer.) Who's in and who's out is a matter of antecedent self-selection. I opt in; the sceptics opt out. Because we don’t agree on the rules, my explanations don’t make sense to them, and their explanations don’t make sense to me. Explanation for the cognitive and biological sciences just is a matter of uncovering the appropriate parallels between the phenomena and the physical system. Huntington’s 4 T note that in each of these cases, there is evidence that such patients still process at least some of the information unconsciously. For example, prosopagnosics claim that they can no longer recog- nize faces upon visual inspection. However, their galvanic skin response changes in the presence of caretakers or loved ones in a manner consistent with their in fact knowing and recognizing the people. For a review of this literature, see Hardcastle (1995). 68 V.G. HARDCASTLE chorea is explained by a disruption in the GABA-ergic loop. Equilibrium in neurons is explained in terms of the influx and efflux of ions across the cell membrane. Perceptual binding is explained (maybe) in terms of 40 Hz neuronal oscillations, The withdrawal reflex in Aplysia is explained in terms of patterns of activation across the motor system. Echolocation is explained in terms of deformed tensor networks. So: find the parallels between brain activity and phenomenal experience and you will have found a naturalistic account of consciousness. Denying the project and devising different criteria for explanation is a perfectly legitimate move to make, of course. ‘There is always room for more. Winning converts though is something else. I wish Chalmers well in that enterprise, for how to do that truly is the gap that remains. References Baars, B.J. (1988), A Cognitive Theory of Consciousness (Cambridge: Cambridge University Press). Block, N. (1995), ‘On a confusion about a function of consciousness’, Behavioral and Brain Sciences, 18 (2), pp. 227-47. Chalmers, D.J. (1995), ‘Facing up to the problem of consciousness’, Journal of Consciousness Studies, 2 (3), pp. 200-19 (reprinted in this volume). CChurchland, P.M. (1984), Matter and Consciousness (Cambridge, MA: The MIT Press). Crick, F. and Koch, C. (1990), ‘Toward a Neurobiological Theory of Consciousness’, Seminars in the Neurosciences, 2, pp. 263-75. Flanagan, O. (1992), Consciousness Reconsidered (Cambridge, MA: The MIT Press). Hardcastle, V.G. (1995), Locating Consciousness (Amsterdam and Philadelphia: John Benjamins). Hardin, C.L. (1988), Color for Philosophers: Unweaving the Rainbow (New York: Hackett). Lycan, W.G. (1987), Consciousness (Cambridge, MA: The MIT Press). McGinn, C. (1991), The Problem of Consciousness (Oxford: Blackwell). Nagel, T. (1974), ‘What is It Like to be a Bat?" Philosophical Review, 83, pp. 435-50. Searle, J. (1992), The Rediscovery of Mind (Cambridge, MA: The MIT Press). Shoemaker, S. (1975), ‘Functionalism and Qualia’, Philosophical Studies, 27, pp. 291-315. Shoemaker, S. (1981), ‘Absent qualia are not possible — A reply to Block,’ Philosophical Review, 90, pp. 581-99. van Fraassen, B. (1980), The Scientific Image (Cambridge, MA: The MIT Press). Kieron O’Hara and Tom Scutt There Is No Hard Problem of Consciousness I: Introduction Cognitive psychology has made many important advances over the past couple of decades. Conceived as a response to the hollow dogma of behaviourism, as a discipline it has clearly shown itself to possess greater explanatory power, and has been widely accepted as an important paradigm for the description and modelling of a number of psychological phenomena. For example the work of Marr and Hildreth (1980), not only outlined the higher level functions (such as edge detection) that the visual system would need to be capable of, but also predicted many of the lower-level neural mechanisms for computing such functions, which have since been revealed by neurophysiological and psychophysical research. ‘One reason for the success of cognitive psychology is its technological aspect. Its methodology is distinctive for its concentration on computational and formal aspects of psychological phenomena, When confronted with some problematic behaviour, the typical response of a cognitive psychologist is to ask: How is that done? He or she typically lives in a world of simulations and models. This concentration on computation provides cognitive psychology with important links to the engineering discipline of artificial intelligence. In AI, the engineer asks: How could that be done? And we can see that the explanations in cognitive psychology will find a ready audience in AI, ready to test the psychological hypotheses to the limit. Because of the important links between cognitive psychology and engineering, how- ever, many thinkers find its theories unsatisfactory as general explanations of human psychological phenomena. The cognitive psychologist, by concentrating on computa- tional aspects of such phenomena, will, according to these critics, miss out on precisely those aspects which are not computational. The cognitive psychologist’s reply is that anything that is literally true about psychological phenomena will be fully accounted for by aconcentration on their computational aspects, and his or her research programme will be an attempt to delve for satisfactory computational or formal explanations. The result is a stand-off, with the critic emphasising the aspects of the phenomena which are prima facie independent of any possible computational account, and the cognitive psychologist standing by the record of the discipline in past investigations. In This paper was originally published in the Journal of Consciousness Studies, 3, No.4 (1996), pp. 290-302. 70 K. O'HARA AND T. SCUTT any particular area, the matter can only be resolved convincingly by the success or failure of the cognitive psychologist in providing successful theories. One such area is that of consciousness. The question is how any straightforward physicalist account, such as would be provided by cognitive psychology, can show how such apparently non-physical phenomena as awareness are possible. The explananda of such an account are not often described as exactly as one might hope, as we shall discuss in section three, but there is a general consensus that, however they are described, they are problematic. In Nagel’s formulation of the problem (1974), if something has con- sciousness, then it is like something to be that thing (it has a subjective aspect). But surely — the critic adds — it isn’t like anything to be a computer program, or formal theory. How can the cognitive psychologist ever succeed in reducing the sui generis phenomenon of consciousness to a series of physical phenomena formally described? This leads to the distinction drawn by the critic between the easy problems and the hard problem of consciousness. The easy problems are the problems that are clearly amenable to the methodology of cognitive science: Chalmers suggests as examples the ability to discriminate, categorize and react to environmental stimuli; the integration of informa- tion by a cognitive system; the reportability of mental states; the ability of a system to access its own internal states; the focus of attention; the deliberate control of behaviour; and the difference between wakefulness and sleep. All these phenomena — and doubtless others — will be susceptible to the methodology of cognitive psychology. Contrasted with these problems is the hard problem, which is that of explaining consciousness itself, how consciousness is possible at all. Answers to the easy problems would fail to address the hard problem, which is the subjective aspect of consciousness, because the easy problems are clearly approachable using only objective terms. The sum of all the (objective) answers to all the easy problems would never be an explanation of subjectivity itself. In this paper, we will defend the claims of cognitive psychology against the critic we have described, by arguing that the hard problem (or HP as we shall call it) is not a serious problem. For a problem to be a genuine problem, some sort of idea of a solution must be available (e.g. some way of recognizing a solution when one stumbles across one); whereas all discussion of HP seems to preclude any sort of answer being given. Our argument has two components. The first component is a pragmatic claim, to the effect that, since we know how to address the easy questions (at least in rough), and we haven't the foggiest idea of what to do about HP, we would be wasting our time if we didn’t explore the easy problems first; we establish this in section two. The second component is a philosophical claim that we do not know what HP is as yet, and that the only possible way of approaching HP is via the easy problems; this is the business of section three, It may be thought that the existence of attempted answers to HP precludes the possibility that HP cannot sensibly be addressed; in section four we discuss some attempted answers to HP (the theories of Edelman, Crick and Chalmers), to show why they fail to address any actual problem, Finally, in section five, we will defend our position against charges that we have eliminated, or would like to eliminate, the notion of consciousness from scientific discourse. THERE IS NO HARD PROBLEM OF CONSCIOUSNESS n Il: Methodological Reasons to Ignore HP In the venerable joke, little Johnny is peering hard at the lawn. ‘Why Johnny,” says his mother, ‘what are you doing?’ Johnny points across the garden and says, ‘I lost my pocket money over there.’ ‘If you lost your money over by the garage, why are you looking by the pond? asks his mother; and little Johnny duly replies: ‘Cos the grass is shorter over here.’ One should never analyse humour, but since this joke isn’t actually very funny, we will, just this once. Why is it a joke that Johnny is looking where the grass is short, and not where the grass is long? Well, he thinks that he will save effort by looking where the grass is short, because such a search is easier in short grass than in long, but we know that he will expend greater effort, because the money is almost certainly concealed in the long grass, and furthermore, little Johnny has told us that himself. So, although a smaller amount of effort would be expended in the short grass per blade (or whatever unit is appropriate here), all that effort is guaranteed to be in vain, since the money is somewhere else entirely. And since he is aware of all that, and has simply failed to draw the obvious conclusion, little Johnny is just being stupid, The analogy to consciousness is evident. The critic we discerned in the opening section argues that the answer to HP is to be found in the long grass over by the garage, and that cognitive psychologists who mess about with the easy questions by the pond will — to the extent that they see themselves as thereby addressing HP — be wasting their time. However, approaching consciousness via the easy problems is only open to the same criticisms as little Johnny's approach to retrieving his money if two conditions hold, to render the analogy suitably watertight. In the first place, in the joke, we know that the money is in the long grass. For Johnny to be really stupid for the purposes of the joke, he must know that the money is in the long grass too, but the main point is that there is a laugh to be had as long as we in the audience know roughly where the money is, And in the second place, it must be the case that Johnny could look in the long grass, that he has a procedure for finding the money such that, if he employed it in the long grass, he would achieve his goal. When we come to examine the extent to which the joke is analogous to the situation in the study of consciousness, we find that neither of these two conditions holds. We do not know where to look for the answer to HP, and in particular, we don’t know that we are necessarily wrong to approach the question via the easy problems. The joke would not be funny if little Johnny knew only that he had lost his money somewhere in the garden; in that event, he would be rational, not stupid, to begin to look where least effort would be expended. For the second condition to hold, it would have to be the case that there was an alternative approach to consciousness to the one provided by the cognitive psychologi- cal investigation of the easy problems. But as itis, no-one really has any well-worked-out plan for approaching the problem any other way. If, in the joke, the long grass was comp- letely and necessarily impenetrable, Johnny could not be blamed for not looking there. ‘We have, in the study of consciousness, an obvious problem which is that addressing the easy problems is the only game in town. We do not know that addressing the easy problems will automatically fail to provide insight into HP, We do not know any other way of addressing HP. We do know how to address the easy problems — indeed, 2 K, O'HARA AND T. SCUTT Chalmers seems to hold this as definitive of the easy problems. Pragmatically, it does look as if looking into the easy problems is as likely to be successful as any other way of approaching consciousness, and therefore it would seem to be as rational a way of proceeding as any. Consider the alternative; the only alternative to the easy way is to develop a new paradigm of psychological investigation. For example — and this is an extreme, if genuine, example — take Penrose’s (1994) suggestion that a new branch of physics is required to explain consciousness. Pragmatically, is there any reason at all why one should take the Penrose route to a solution to HP? Addressing the easy problems is hard enough, and work in those cognitive psychological fields is likely to produce interesting and important results (as no-one denies), even if the end result is not an answer to HP. ‘The advantage of such an approach is that we know where to start, there is a body of work to build on, there are surrounding theories in which such work can be embedded and located. But developing a new branch of physics is the sort of paradigm-shifting, epoch-making task that, unless one was Nobel Prize material, one should probably avoid. Worse, one is not even guaranteed success with respect to the HP, since any such suggestion remains tentative, though the Nobel Prize ought to be consolation enough. It is difficult to see why anyone who is neither seriously irrational nor seriously clever, and who wishes to contribute to a solution to HP, would not take the pedestrian route and go for the easy problems. We call this argument the pragmatic argument. Ill: Philosophical Reasons to Ignore HP Surprisingly, in this debate, pragmatic considerations carry little weight. Although the pragmatic argument seems to us to be sufficient to consign HP to the dustbin reserved for historical curiosities, the critic of cognitive psychological approaches to consciousness will certainly not agree. More a priori reasons are required here, since all the arguments to be met are firmly unencumbered with empirical detail. Therefore, in addition to the pragmatic argument, we will now add two compelling a priori arguments to leave HP alone. Both are concerned with the lack of understanding of the nature of HP that precedes satisfactory answers to the easy problems. The first a priori argument, which we shall refer to as the context argument, simply makes the obvious point that an understanding of a concept may be facilitated, at the very minimum, by the investigation of contexts in which that concept is important. Take the example of the concept of life. The analogy of HP in this context is the question of what life is (call this LHP). The analogy of the easy problems would be various biological questions to do with the processes that go on in things that are alive. Many major philosophers addressed the LHP, but very few addressed the easy problems. Indeed, between Aristotle and Harvey it is not obvious that anyone did. We can also agree that, remarkable as they were, Aristotle's discussions of the easy problems were inadequate. Yet as the discipline of biology grew up in the modern era, more and more easy problems fell to the advancing forces. Now, what happened to LHP? Since the vitalists, no-one has seriously attempted to answer the question; indeed, very few thinkers think the question is worth asking at all. Those who do still wonder about LHP are increasingly likely to THERE IS NO HARD PROBLEM OF CONSCIOUSNESS B assimilate it with one of the easy problems. The answers to the easy problems gradually undermined the air of mystery that underlay LHP, and suddenly LHP didn’t really seem like a problem any more. Consciousness may or may not be analogous to life in this way. It may be the case that answering all the easy problems completely and satisfactorily would still not get rid of HP. But that cannot be decided a priori. Clearly, answers to the easy problems are bound to lead to adjustments in the concept of consciousness. For example, there is no doubt that the research on epileptics who had had their corpus callosum severed (Gazzaniga, 1970; Nagel, 1971), so-called ‘split-brain’ patients, has led to a serious readjustment of the notion of the unity of consciousness, and no philosopher or psychologist investigating in this field can afford to ignore such results. But surely no-one would claim that the phenomena of consciousness thereby revealed would have been discovered a priori, ‘without empirical research into this problem (presumably an ‘easy’ problem). Hence, had this easy problem been neglected in favour of HP, any putative solution to HP would quite simply have been wrong, because it would have failed to account for important properties of consciousness (in which case it couldn’t have described consciousness fully). Further- more, the only way to provide criticisms of the putative solutions to HP would be to perform the research into the easy problems to discover the properties of consciousness overlooked by HP-research. So the very notion of consciousness we are dealing with will change as more answers to easy problems roll in (right and wrong answers). Older notions of consciousness will be seen to have been mistaken, or simply wither on the vine as their role in scientific research diminishes, Hence, addressing the easy problems will seriously affect any answers to HP that we get, via their effect on the notion of consciousness whose nature is the topic of HP. The lesson is clear: we cannot be confident of any putative solution to HP until we have a reliable picture of the topography of the ground via the less ambitious research into the easy problems. Not all easy problems would have to be solved, but a well-supported psychological consensus will, without a shadow of a doubt, precede any serious assault on HP. That consensus is not yet with us, and will only follow a long period of dedicated research into the easy problems. The second a priori argument, which we shall call the epistemological argument, is an adaptation of a recent argument from Colin McGinn (1989). McGinn’s argument has been widely criticized, and with good reason. But we feel that there is a kernel of truth in it, which is highly germane to the issue at hand. McGinn argues that we will never understand consciousness (i.e. we will never solve HP), and argues innovatively on epistemological grounds. His argument goes as follows. We are all good physicalists, so any explanation of consciousness (solution to HP) will consist of some physical predicate, P, satisfaction of which by an entity will entail that that entity has consciousness. But, argues McGinn, it is inconceivable that we would ever recognize that a mere physical predicate confers consciousness in this way; we just do not understand consciousness in the right terms. So even if we discovered P, we would never recognize that it was the solution to HP. ‘There are many dubious points in McGinn's argument (which in addition we have necessarily caricatured). It is not obvious to us which side of our particular fence McGinn sits; he would probably advocate the pursuit of the easy problems on the ground that HP 4 K. O'HARA AND T. SCUTT will be forever out of our reach, and therefore would probably see the point of our methodological argument in section two. But he would also discount the possibility of HP fading away, or changing rapidly, as a result of the investigation of the easy problems, and in this he agrees with the critic of cognitive psychology we discerned in section one. ‘What McGinn has done is highlighted the importance of our recognition that a solution to HP isa solution to HP. We do not agree that it is necessarily the case that we will never make such a recognition. But clearly some prerequisites for such a recognition must obtain before McGinn can be shown to be wrong. One such prerequisite is that we must know what we are talking about; until the nature of consciousness is more or less well-understood, we will never recognize that a theory solves HP. Note that this is independent of whether or not the theory actually does solve HP. Hence until HP is well-understood, it won’t be rational to address HP. ‘As a parallel, consider the work of ancient philosophers about the nature of matter. Their version of HP was MHP, what the nature of matter is. The easy problems in this realm are the way that matter reacts in various circumstances (in effect, the problems of chemistry), Now it tured out that Democritus, and later Lucretius, hit the nail on the head with their solution to MHP, which was an atomic theory. Although their theories weren't particularly detailed in the way in which modern atomic theories are detailed, they were more nearly right than Anaxagoras or Empedocles. But, because these authors were addressing MHP before the easy problems had been elucidated to any significant degree, it is fair to say that — from the point of view of chemistry at least — their answers to MHP are effectively worthless. However wonderful De Rerum Natura is as a piece of philosophical poetry, it is clear that its value is quite independent of the discovery that the right answer to MHP is that matter is atomic, as evinced by the poem’s high reputation during periods when it was thought that the answer to MHP was that matter is not atomic, And it is also clear that Democritus and Lucretius weren't right for the right reasons, They were neither cleverer, nor more rational, nor more well-informed than their rivals; itis simply that their shot in the dark happened to hit the target, while their rivals’ didn’t. Itis our contention that the thinkers who approach the notion of consciousness via HP are going to be subject to the same problems. It may indeed be the case that a new branch of physics is required to solve HP, as Penrose claims. It may be the case that information has an irreducible phenomenal aspect, as Chalmers suggests. But since neither of these authors has any idea of the detailed consequences of their solutions to HP, then neither of their answers can be considered as authoritative. And the reason that neither author is able to be authoritative is that work on the easy problems has barely commenced. A research programme, the programme of cognitive psychology, is underway to address the easy problems. That programme may fail completely, or may gain partial success, or may claim to provide complete solutions to the easy problems, When the programme has been carried through sufficiently to make an estimate of the extent of the success of the research, anyone who makes a stab at HP will be on much more solid ground. And even if one of the currently competing claims — all shots in the dark, as we claim — does hit the target, that will be coincidence, and will be rejected as such. THERE IS NO HARD PROBLEM OF CONSCIOUSNESS a IV: Edelman, Crick and Chalmers’ Attempts to Address HP The obvious reply to the arguments adduced in sections two and three above is that, although we have claimed that addressing HP will be at best very difficult and at worst impossible, there has been quite a number of serious attempts at solving HP. In turn, the obvious rejoinder to shat would be that these attempts fail in the ways that we have suggested that they will fail. ‘We have already discussed Penrose’s (1994) claim that a new branch of physics is required. This does not even look like an approach to HP, at least until the putative branch of physics is exhibited, Similarly, McGinn’s (1989) idea that a physical predicate P, true of the brain, will explain ‘the mind’, is hardly of use in the context of HP, given that he doesn't suggest what this predicate may be (because he thinks that we can never understand how the predicate will explain the mind), which makes his argument more like a testament to his faith in materialism than anything else. Any solution to HP must give us a handle on the problem. Other attempts simply do not have the right form. Eccles’ theory (Popper and Eccles, 1977) that random quantum effects will undermine physical determinism and ‘et in’ the mind will ultimately fail, because the effects of the mind are specifically not random. Our aim in this section will be to take a number of influential attempts on HP, and show that they succumb to the three arguments given above. Our aim is to criticize theories that are plausible, in the sense that Penrose’s, McGinn’s and Eccles’ are not, and furthermore which have gathered attention and support from the community. We hope they are sufficiently representative. Of course, criticizing existing theories can never be decisive — since another, better, theory may always come along later — but we expect that the exercise will demonstrate the plausibility of our arguments. To that end we consider the theories of consciousness of Gerald Edelman (1992), Francis Crick (Crick, 1994; Crick and Koch, 1990) and David Chalmers. 1. Edelman and the pragmatic argument Edelman distinguishes between Primary and Higher-Order Consciousness. Animals with primary consciousness have phenomenal experience, but are not aware that they have it. They have no notion of self, and therefore nothing to ‘attach’ the experience to. Primary consciousness consists of a re-entrant loop connecting value-category memory to current perceptual categorization (1992, p. 120). There would be a fleeting phenomenal aspect, but Creatures with primary consciousness, while possessing mental images, have no capacity to view those images from the vantage point of a socially constructed self (Edelman, 1992, p. 124). Edelman is very concerned — rightly, of course — to make sure that this notion of consciousness has a place in neuropsychology, and is not merely a metaphysical conceit. ‘What is the evolutionary value of such a system? Obviously, primary consciousness must be efficacious if this biological account is correct. Consciousness is not merely an epiphenomenon (Edelman, 1992, p. 121). ‘The question then is how this promise can be cashed out. And this is where what Chalmers terms the bait-and-switch takes place. Edelman lists the properties that make 16 K. O'HARA AND T. SCUTT consciousness useful (such as relating the creature's present inputs to previous dangers or rewards as experienced). But what actually happens is that Edelman lists the benefits of the re-entrant loop, the connection between value-category memory and current perceptual categorization, i.e. the thing he attached the consciousness label to in the first place. This is a common tactic: take some neuro-architectural item that seems to be doing the sort of thing we want consciousness to do; attach the consciousness label to it; and then show the benefits of consciousness by pointing to the assets of our newly labelled neuronal structure. We can see that Edelman has fallen foul of the pragmatic considerations we discerned in section two. The only way we can make progress is to tackle the easy problems; concentration on HP will always fail to provide a lead, Hence, when attempting to tackle HP, the temptation will always be to move covertly towards the easy problems. In Edelman’s case, the re-entrant loop is a phenomenon that is eminently open to study, and he applies his considerable expertise to that end, with interesting results, But the meth- odological difficulties involved in gaining a vantage point for HP will always lead the researcher to move via the easy problems, and we see why a common class of theories emerges, of which Edelman’s is an exemplar, in which a neuronal phenomenon is assimilated to the phenomena of consciousness. Chalmers is correct to say that such a strategy will always be unsatisfying for someone interested in HP. Incidentally, Edelman also exemplifies another methodological pitfall in assaulting HP. As Dennett puts it, Edelman’s theory is + an instructive failure. It shows in great detail just how many different sorts of question must be answered before we can claim to have secured a complete theory of consciousness, but it also shows that no one theorist can appreciate all the subtle- ties of the problems addressed by the different fields (Dennett, 1991, p. 268, n.1). There are so many easy problems to be understood, in so many fields, that the researcher who would take a crack at HP will be hard pressed to find (and argue for) a point of view that transcends them all. But such a point of view is vital if HP is to be regarded as a problem essentially distinct from the (sum of the) easy problems. Again, pragmatically, there are few who can move so easily between all the various disciplines that they can sift through the evidence provided by the various answers to the various easy problems and make correct judgments about them. 2. Crick and the context argument Francis Crick, rather than give an exact location for consciousness (he does give locations for certain phenomena, for example, free will, in his (1994), but these speculations are probably better overlooked; indeed, perhaps his tongue is in his cheek), points instead to an observable phenomenon — certain 35-75 hertz neural oscillations in the cerebral cortex (Crick and Koch, 1990). These oscillations appear to be correlated with the binding of information from various sense modalities, but it remains unclear whether Crick and Koch think that these oscillations form the basis for consciousness, are merely correlated with consciousness, or form an observable aspect of consciousness. Chalmers is right to criticize Crick and Koch on the grounds that suggesting a correlation between two phenomena in no way explains how one gives rise to the other (or indeed, which gives rise to which). THERE IS NO HARD PROBLEM OF CONSCIOUSNESS 7 Of course, concentration on particular observables amounts yet again to an assault on an easy problem. Once more, HP has been sidestepped, as Chalmers points out. But Crick and Koch are prepared to get excited over the properties of the easy problem they have concentrated on, We have suggested that one of the functions of consciousness is to present the result of various underlying computations and that this involves an attentional mechanism that temporarily binds the relevant neurons together by synchronizing their spikes in 40 hz oscillations (Crick and Koch, 1990, p. 272). Let us assume that they are right, for the sake of argument. Now, consciousness as we presently understand it is not like this at all. If Crick and Koch have got it right, then our view of consciousness will have to change, to conform to their results. But note that itis their concentration on a particular easy problem that has given them this perspective on HP. In that event, the form that HP has finally taken would have depended on the results of some investigation into an easy problem. On the other hand, if Crick and Koch have not provided us with any sort of handle on HP, as Chalmers says, then it is clear that they have uncovered some interesting results concerning the easy problems. No view of HP could fail to take these results into account. Good work on the easy problems will always have to be incorporated into HP. And such results are unlikely to be predictable in advance; recall our remarks concerning brain bisection and the unity of consciousness. In other words, as we claimed in section three, the easy problems, when solved, alter HP; hence HP could not be considered as reaching any sort of ‘final form’ until the easy problems were beginning to be cracked. They don’t all have to be solved, but we need to be comfortable enough in this area so that we do not expect to be surprised by any new discoveries. ‘Comfortable enough’ is a vague term, but its application seems clear. We are pretty familiar with the functions of the lungs, for example, and do not really expect to have our understanding of them turned upside-down (though of course that under- standing is always liable to be adjusted in the light of new discoveries). On the other hand, Big Bang theory is so new and untested that any speculation on the basis of its current form is probably going to be made redundant by massive changes in the theory itself. Our claim is that consciousness theory is more like Big Bang theory than lung theory, and that wwe are not yet comfortable enough with consciousness theory for it to be worth the risk in wasted resources to start speculating. 3. Edelman, Crick, Chalmers and the epistemological argument The theories we have so far considered also provide evidence for our epistemological argument. Both Edelman and Crick are distinguished neurobiologists who have claimed to have located consciousness in the brain. In neither case is this fully explained, although in both cases it is work on the ‘easy’ problems which has led to a suggested solution of HP. We do not share Chalmers’ pessimism that work on the former can never help with our search for the latter. However, at present it is too early (indeed, it often looks plain ridiculous) to point to a part of the brain and say ‘this is where consciousness lies’. Diagrams of cognitive architecture in modern texts on the nature of mind often look like 7B K. O'HARA AND T. SCUTT medieval maps with labels saying ‘Here be Dragons’. Until we have a much clearer idea of what we are looking for, it is far too early to say where we have found it. Chalmers, on the other hand, takes a very different route, but will still fall foul of the epistemological argument. His position is founded on an a priori assumption that there can be no scientific account of consciousness (at least in physical terms as currently understood). He claims that there needs to be a non-reductive explanation of the hard problem of consciousness. Reductive explanation, according to Chalmers, won't get us anywhere. He gives an example of where such examples succeed and fail: If someone says, ‘I can see that you have explained how DNA stores and transmits hereditary information from one generation to the next, but you have not explained how it is a gene,’ then they are making a conceptual mistake . . . But if someone says, ‘I can see that you have explained how information is discriminated, integrated, and reported, but you have not explained how it is experienced,’ they are not making a conceptual mistake. This is a nontrivial further question (p. 13). Indeed. But that is because at the moment it is impossible to see how experience could be explained in terms of mere (sic) information processing — we lack the necessary bridging concepts to see one in terms of the other. In fact we are in much the same position as a geneticist would have been if such a term existed two hundred years ago. From our present viewpoint, a statement such as ‘I see how you have explained the nature and transmission of electromagnetic radiation in the range 400-700 nanometres, but you have not explained how it is light’, seems somewhat naive, But before these scientific facts about the electromagnetic nature of light were carefully laid out by that painstaking and haphazard endeavour we call science, the validity or otherwise of this sentence would have been open to debate (and a very ill-informed debate at that). At the end of the next millennium, a statement that consciousness can be explained in terms of X, ¥ and Z may be widely accepted as being beyond doubt. The same statement from our present viewpoint would be viewed with some scepticism, because we haven’t yet done the work (on the easy problems) which would justify this statement. Chalmers himself admits that ‘there is a direct correspondence between consciousness and awareness’ (p. 22). Wouldn’t it make sense, this being the case, to research awareness (the easy problem) to better understand consciousness? Is Chalmers really saying that under- standing the former doesn’t help us at all in understanding the latter? The naive question which began this paragraph is naive only so long as it is understood in a particular way (eg. from the standpoint of the twentieth century). Before, in Kuhnian terms, the paradigm shifted, it is a perfectly good question. For the paradigm with respect to consciousness to shift, however, a good deal of evidence has to be built up with no obvious home in any general theory — this is what provides the impetus for a paradigm shift. And this evidence will only be built up by addressing the easy problems. So what are we to make of Chalmers’ alternative to a physical account of conscious- ness? In the ‘dual-aspect theory of information’, it appears to us that ‘Information Space’ — the place where ‘experience might have a subtle kind of causal relevance in virtue of its status as the intrinsic nature of the physical’ (p. 28) — is the pineal gland of Chalmers’ account. However overstated some may have been in their heralding of Crick and Koch’s neural oscillations as an aspect of consciousness, at least such work on the easy problems ‘THERE IS NO HARD PROBLEM OF CONSCIOUSNESS 79 may get us nearer to a solution of HP (or nearer to a reasoned rejection of a physical account). It is not so clear that the non-reductive account of Chalmers offers any such reward. Itis, in our current epistemological state, impossible to understand how informa- tion can have a phenomenal aspect. It may turn out to be true, but we, in the twentieth century, have no idea what the implications of such a claim are. It is as unexplanatory to us as atomism was to the Ancient Greeks. V: Eliminativism with Respect to Consciousness and HP ‘There is one final point that we would like to make, and that is to rebut the charge that our insistence that concentration on the easy problems and consequent neglect of HP amounts to eliminativism with respect to consciousness. HP, such a charge would maintain, is the fundamental, definitive problem of consciousness. Anything else, such as the ability to discriminate environmental stimuli, is mere effect. Hence to avoid or reject HP amounts to rejecting the scientific investigation of consciousness. Such a strategy results either in the elimination of consciousness from our discourse, or the cheerful acceptance of its irreducibly occult nature. A sophisticated version of this charge would attempt to line us up with an eliminativist argument such as that of Kathy Wilkes (1988), where sceptical play is made with the notion that consciousness is not used to any great effect in natural science, is not rooted ‘in observation or experiment’ (1988, p. 38), cannot be seen as a natural kind, and has little or no explanatory power. The idea then is that, since the idea of consciousness fails to have any scientific use (although of course it may be the case that scientists have as a ‘goal the elucidation of consciousness — i.e. the investigation of HP), nothing would be lost were consciousness removed from scientific discourse altogether. First we note that this argument will not support the elimination of the notion of consciousness altogether, in a way that would effectively remove mental terms from discourse generally. Wilkes’ argument will only apply to scientific discourse, as she herself is more than willing to concede (1991). Hence she is advocating treating the term ‘consciousness’ rather like the term ‘dirt’, admitting that it has a ready usage in ordinary everyday talk, while keeping it out of more technical scientific discourse. Second, we note that if those who would support the investigation of HP are right in saying that HP is the only way of ‘getting at’ consciousness, then concentration on the easy problems and the removal of HP from the realm of scientific discourse would constitute an elimination of consciousness from science, As we argued in section three, we do not accept that they are right in so saying. Our main reply to this charge, however, is to side-step it, We are quite happy to agree with Wilkes that her conditional premiss is correct (with a small modification). That is we accept that if consciousness has no use in any practical context, then it should be eliminated from such contexts. We have substituted the term ‘practical’ for ‘scientific’. We do not want to define ‘practical’ here, but it certainly includes the scientific. Whether all practical contexts are scientific is a question we would like to leave open (many famous philosophers could be lined up both for and against that proposition). It follows from this that we would accept the eliminativist charge if we accepted ‘Wilkes’ non-conditional premiss that consciousness has no use in any practical context. 80 K. O'HARA AND T. SCUTT However, we don’t. There are a number of areas where some aspects of the notion of consciousness are used ineliminably. We think that the easy problems are examples of these (although this begs the question against our pro-HP critic). Perhaps the clearest case of a scientific (and therefore practical) context in which the notion of consciousness is centrally involved is the discipline of anaesthesiology, which is the science of making people lose conscious awareness. This is not the place to go into detail on this matter, but see Nikolinakos (1994) for a recent discussion of scepticism about consciousness in the context of anaesthesiology. We think that the elimination or otherwise of the notion of consciousness depends entirely on that notion being used practically, and would not wish our attachment to it to be merely sentimental. Fortunately, we can get serious work out of the concept, in a number of psychological and/or medical fields, and therefore, despite our rejection of HP, we are very happy to accept that the notion of consciousness has an important role to play in our scientific and practical discourse. Of course, the nature of that notion would depend on its use in those practical contexts — many aspects of the ‘common-sense’ or ‘folk’ notion of consciousness would turn out to be of little or no use in any practical context. But we think that this is a good discipline, and should help prevent metaphysical obscurities muddying the waters of practical discourse. VI: Let’s Take It Easy Our aim in this paper has been to establish that the easy problems of consciousness, as addressed by cognitive psychology, remain the most promising route to success. We adduced three arguments against approaching HP directly: © the pragmatic argument — we know how to address the easy problems, whereas we can only speculate with respect to HP; © the context argument — the elucidation of the easy problems will probably change the nature of our understanding of HP, and so any attempt to solve HP currently would be premature; © the epistemological argument — a solution to HP would only be of value if its success could be recognized because the problem was well-enough understood. We illustrated our claim with discussions of three prominent attempts to solve HP. Edelman, trying to avoid vacuousness, ends up conflating his answer to the easy problems with his answer to HP, showing the perils of ignoring the pragmatic argument. Crick, like the researchers into the split brain phenomena before him, doesn’t solve HP, but provides important input into HP, subtly changing its nature, as we would expect from the context argument. Because our understanding of HP is so limited, Edelman, Crick and Chalmers all fall foul of the epistemological argument. Our final point was to say that our preference for the easy problems and consequent neglect of HP does not amount to eliminativism with respect to consciousness. In this concluding section, we would like to expand a little on the nature of the easy problems and the hard problem. Part of the difficulty here is caused by the names given to the two groupings. Chalmers says: THERE IS NO HARD PROBLEM OF CONSCIOUSNESS 81 The easy problems of consciousness are those that seem directly susceptible to the standard methods of cognitive science, whereby a phenomenon is explained in terms of computational or neural mechanisms. The hard problems are those that seem to resist those methods (p. 9). Of course, rhetorically, easy problems don’t seem to be worthwhile, and an assault on the hard problem appears more heroic. No prominent player in the game, least of all Chalmers, would make the crass error of saying that the easy problems are actually easy. Equally, no-one would suppose that HP were not hard. But itis interesting to explore the reasoning behind calling the easy problems easy and HP hard. The easy problems are not easy, but we do have a handle on how to approach them. We have no guarantee of success, but at least we have a rough idea of where to go from here, What sets HP apart is the fact that we haven't the foggiest idea of how to begin. Compare, for example, Penrose’s solution with Chalmers’ or Crick's. They are not even remotely in the same ball park. We have three answers by three leading thinkers in three central disciplines, and they point in three mutually exclusive directions. Consider an analogy with bridge building. Easy bridge building involves constructing structures to cross small spans. Hard bridge building involves constructions over more challenging spans, in unfavourable conditions, These notions of ‘easy’ and ‘hard’ do not map onto the ones used in this debate, Suppose we termed bridge building ‘easy’ just in case we had a rough idea of how to start, no matter how problematic the project, and ‘hard’ just in case we did not know how to start, Then an easy project might be the construction of a bridge across the Atlantic, and a hard project might be the construction of a bridge into the fourteenth century. A hard bridge builder then might echo Penrose’s solution, that a new branch of physics needs to be developed before hard bridge building might begin in earnest. But the obvious response would be to say that, at our current state of ignorance, hard bridge building ought be be shelved, in favour of the easy sort, at least for the time being. Science is an incremental activity. For every Einstein there are thousands of Scutts and O’Haras, adding their two penn’ orth to the current state of knowledge. Ambition plays a role in science, but ambition needs to be tempered by a firm grasp of the possibilities available at any point. In particular, our understanding of the large questions often crucially depends on our answers to all the small and relatively insignificant questions. ‘When enough of the small questions have been answered, then an attempt to answer the hard questions becomes a possibility (unless the hard questions have been shown to be wrong-headed). ‘We all want to understand consciousness. But the way to achieve that understanding is not to divert intellectual resources toward HP. Because the small questions relating to consciousness remain unanswered, we simply do not understand HP sufficiently well to determine whether there is a serious question to answer (and if so, precisely what that question is), or whether the whole thing is some ghastly metaphysical error best avoided, We can have opinions on these questions, sure, but those opinions are fundamentally ill-informed, and will remain so until the lessons from the easy problems have been learned and absorbed.

You might also like