Amit Goswami

Institute of Theoretical Science,University of Oregon, Eugene, OR, 97403

The Hard Questions: View from A Science within Consciousness

Abstract

The hard questions of consciousness raised by Chalmers (1994) are extended and addressed from the point of view of an idealist science based on the ontology that consciousness is the ground of all being. It is shown that such a view, implicated in the context of quantum measurement theory, resolves not only the paradoxes of experience raised by Chalmers but also those of free will, the distinction of unconscious and conscious processing, and mind-brain identity.

Fall 1997 | Spring 1997 | Fall 1996 | Spring 1996 | Other Essays

I. Introduction

David Chalmers (1994) has designated the question of how experience comes about as the hard question of consciousness. While I agree that this is one of the hard questions, three others are as hard from a material realist view (the ontology that everything arises from matter and its correlates, energy and fields).

To begin, then, let me briefly state my four hard questions.

1. The question(s) of experience: experience consists of a perceived split of the world into one part (the subject, which may sometimes be implicit) that experiences the other part (object) as separate from it. How does the one world of matter separate into two, subject and object?

There are other related hard questions. An experience usually involves several brain areas, although the oneness of conscious experience is hard to debate. But how do we explain this oneness, the binding problem?

Although there is commonality in the intentionality of our experience, there is also undoubtedly a subjective quale. How can a subjective quale be explained from a science which is avowedly purely objective?

Similarly, although most of our experiences can be traced to the local senses, certain experiences--the creative, the psychic, and the mystical--smack of non-locality. The validity of these experiences is debated incessantly, reflecting a strong material realist prejudice against them. But the data concerning the non-locality of some experiences is as good as other cognitive data (Jahn, 1982).

The implicit or explicit subject of our local experiences is a local, personal "I" that we call the ego. But the implicit subject of the non-local experience is neither local nor personal; it is non-local and transpersonal. This two-level self-identity (Maslow, 1968; Assagioli, 1976) connected to experiences needs an explanation.

2. The question of downward causation: The material realist ontology assumes that all causation is upward; causal potency, in this view, ultimately rests with the reductive elementary particles of matter and their interactions. Yet, we experience real freedom when we are creative, when we are compassionate, when we make moral decisions. This implies downward causation--causal potency that originates with us, whatever us is. This is a very difficult question for material realist science.

3. The differentiation of unconscious and conscious processing: There is now data showing unconscious perception--a subject "sees" enough to act on the basis of it, but is not conscious of the seeing. Unconscious processing is also acknowledged as part of the creative process (Wallas, 1926), and much evidence exists in favor of it (Goswami, in press, b). But in material realist cognitive science, it is not easy to formulate a paradox-free distinction between the unconscious and conscious that also agrees with experimental data (McCarthy and Goswami, 1993).

4. Last but not least, the question of the mental as opposed to the physical aspect of an experience: Experiences consist of intentionality toward an object, but the physical object is not the only object. There is also a mental object in practically every event of conscious experience. A simple example is when I see a rose. I also concomitantly experience some such thought as: I see a rose. And I experience this thought not in the ordinary, public physical space of the rose but in a private, mental space that we call awareness. Can mental objects arise from the purely physical? Chalmers has included this question within the broad puzzle of experience. I will consider it separately for reasons that will be clear as we proceed.

Much debate exists about whether material realist science can or cannot answer these hard questions. Many philosophers think not. Thus, proponents try to do away with the questions rather than deal with them. The hardest question to banish is the question of experience (although even that has been tried, see Dennett, 1991), leading to Chalmers' strategy: let's challenge the material realist with this one question.

But a different view of science has begun in which all the questions, although hard, can be addressed. This is a science within consciousness. It is a science also based on a monistic ontology, but the ontological reality of the world is consciousness--consciousness is the ground of all being, not matter. This idealist science is crucially grounded in an interpretation of quantum mechanics and quantum measurement theory (Goswami, 1989, 1990, 1993, 1994). The advantage of this approach further obtains from the fact that it incorporates dualist approaches, such as those of Eccles (1993) and Stapp (1993), and even the materialist approach to consciousness and cognition (Goswami, in press, a).

Here, briefly, is the plan for the paper. I will first discuss quantum measurement theory from the monistic idealist view and discover the nature of consciousness from this vantage point. Next, I will show how all the different aspects of the question of experience can be addressed. I will also take up the questions of unconscious processing and downward causation. All this, so far, is a review of my earlier work. Finally, I will take up the question of mental as opposed to physical objects, and introduce some new ideas to incorporate the dualistic notions of Eccles (1994) within idealist science.

II. Quantum Measurement Theory and the Nature of Consciousness

The wave amplitude of an object in quantum mechanics (technically called a wave function or a coherent superposition) corresponds to a spread-out wave of possibility in potentia, but when we observe, we see the object localized like a particle. A quantum measurement corresponds to a discontinuous and non-local collapse of a wave-like state (many possibilities) into a localized particle-like state (one actuality). Who/what chooses which possibility is manifested in a particular measurement? If choice is involved, is consciousness? And if consciousness can collapse the quantum wave, can such a consciousness be made of matter, be an epiphenomenon of matter?

Quantum physicists have argued these matters of interpretation for decades without consensus. But in truth, the solution to the measurement paradox already exists--namely, to assert, as the mathematician John von Neumann (1955) originally did, that it is consciousness that collapses the quantum possibility wave. It is consciousness that chooses which possibility will manifest in actuality. All the objections (paradoxes) to this resolution are due to an almost universal misconception among scientists about the nature of consciousness. They tend to think that any positing of consciousness collapsing the possibility wave must refer to a dualistic consciousness--a consciousness separate from matter (Wigner, 1962, 1967). I (Goswami, 1989, 1990, 1993) have shown that if one understands consciousness as the ground of being (I call this ontology monistic idealism), then all objections find simple satisfying answers, as we will see below.

The measurement problem is squarely resolved when we turn the meta physics of science upside down and posit consciousness as the ground of a ll being, and thus, having causal efficacy--downward causation. What els e can we say about a consciousness that collapses the quantum possibility into actuality? All objects are quantum objects. Therefore, any machine, such as the ones called "measurement apparatuses" that we use to amplify a quantum phenomenon, itself becomes a possibility wave (a superposition of macroscopically distinguishable possibilities) when in contact with micro-quantum possibilities. This includes our brain-mind. Consciousness can collapse the whole conglomerate because it transcends the material universe. Does such collapse constitute mind over matter? No, consciousness transcends both matter and mind, which exist as possibilities within consciousness before collapse. Collapse consists of recognition and choice.

If two people simultaneously make an observation, whose choice counts (a slight variant of this paradox is sometimes called the paradox of Wigner's friend)? Neither's. Consciousness is one, unitive (see also, Blood, 1993). Our separateness is only an apparent one (see later).

So, if a tree falls in the forest, is there a sound if nobody is t here to hear it? Centuries ago, Bishop Berkeley said that God is always in the "quad," in the forest, to hear the sound, so the sound is there. But not so with quantum measurement. If transcendent consciousness is al ways looking and collapsing, quantum possibilities would never develop an d all the wonderful phenomena of quantum physics that give us the technologies of computers, lasers, and superconductors would be impossible. The solution is to realize that consciousness collapses the possibility wave only in the presence of an immanent observer. The measurement is tangle d-hierarchical and produces self-reference. An example of a tangled hierarchy is the self-referential sentence, I am a liar. Neither the subject nor the predicate of the sentence is the top-level, each qualifies the other (Hofstadter, 1980). This tangled hierarchy causes the self-reference of the sentence (for further details, see Goswami, 1993).

Out of the self-referential measurement itself simultaneously arise a subject--which I call the quantum self--that measures, that chooses, that observes, and object(s) that are observed. Notice how, in this description, dualism is avoided because ultimately there is oneness (the division is only an appearance), allowing subjects and objects to be treated on the same footing. Objects have upward causation by virtue of the laws of quantum possibility dynamics that they follow. The subject has downward causation that comes from its freedom of choice to collapse actuality from possibility, creating manifestation.

What is quite reassuring is that these properties of consciousness --transcendence, unity, and self-reference, derived from the requirement that consciousness collapse the quantum wave function without raising any new paradox--are also the characteristics of consciousness that mystics from every age have declared based on their direct realization. This is the reason that the ontology used here, monistic idealism, has also been termed perennial philosophy.

The Search for the Quantum Mind

Crucial to the monistic idealist resolution is the idea of self-reference in the brain-mind-- the simultaneous co-arising of the choosing subject and the experienced object defines the collapse; thus there is no dualism involved. This self-reference is also the most important bra in-mind paradox--how is it that we can refer to ourselves? The two paradoxes, self-reference and quantum measurement, find simultaneous resolution under the idealist ontology if we posit additionally that the brain has quantum machinery in addition to the neuronal machinery that act as amplifying measurement apparatuses for the quantum.

That the human brain may contain a quantum system in addition to its "classical" neuronal system is a decades-old idea (Walker, 1970; Bass , 1975; Stuart et al, 1978; Stapp, 1982, 1993; Wolf, 1984; Eccles, 1986, 1994; Lockwood, 1989; Donald, 1990; Goswami, 1990; Burns, 1991; Hameroff, 1994; Penrose, 1994). What follows is a brief summary.

How does a laser get its special intensity in a narrow pencil beam or a superfluid its special flow characteristic? The answer in each case is the phase coherent motion (as in dancing in step) of the constituent s. One type of model of the quantum in the brain-mind posits a superfluid-like coherence in the movement of a constituent matrix (Stuart et al, 1 978; Lockwood, 1989) arising from the interaction dynamics of the many-body system. The latest entry in this field is the work of Hameroff (1994) who sees this coherent build-up in the structure of microtubules within the brain cells (see also, Penrose, 1994).

What prevents ordinary macro-objects from displaying significant quantum uncertainties of movement is their mass. A second type of model f or the quantum in the brain-mind attempts to find carriers of relatively small mass so that their movement is adequately quantum to create the required ambiguity for consciousness to operate upon.

How does an electrical impulse pass from one neuron to another ac ross a synaptic cleft? Conventional theory says that the synaptic transmission must be due to chemical neurotransmitters. E. H. Walker (1970) th inks that the synaptic cleft is so small that quantum tunneling of electrons may play a crucial role in the transmission of nerve signals. Eccles (1986, 1994) has discussed a similar mechanism for invoking the quantum in the brain--his "microsites" that mediate the quantum connection between neurons at the synapses do seem to satisfy the small-mass requirement o f quantum behavior (Herbert, 1993).

Stapp (1993) also thinks that quantum processes play a key role in the release of neurotransmitters from vesicles into a synaptic junction. An action potential pulse opens channels for diffusion of calcium ions into the vesicular release sites. But the calcium ions are of small enough mass, and thus their diffusion is quantum. Whereas for "classical" diffusion there is only one possible path of flow (for all practical purposes) for the ions, in quantum diffusion the ions simultaneously flow in many possible paths, creating a substantial ambiguity. Many such quantum interactions occur in possibility at many synaptic sites. This gives the brain, upon amplification by other neural processes, a macroscopic possibility structure until one component in that coherent superposition of possibilities corresponds to a state of macroscopic cognitive meaning that consciousness recognizes. At this point consciousness collapses that component of the uncollapsed coherent superposition, all the neurons involved in that meaningful state simultaneously fire, and a perception arises (a long with a subject).

So perhaps the brain has a quantum system that produces ambiguities, sharing the job with classical measurement apparatuses that amplify th e ambiguities in a macroscopic state of coherent superposition of possibilities. At our current state of knowledge, we cannot discern between the various models of the quantum machinery, but one thing is certain: consciousness is needed to make actuality out of the possibilities that the du al quantum system/classical measurement apparatus(es) present. This is t he idea that I (Goswami, 1990) have adapted into an idealist model of consciousness, quantum measurement, and self-reference called quantum functionalism.

To see how a tangled hierarchy arises in the brain-mind, let us examine a crude model of the brain-mind's response to an ambiguous stimulus (Goswami, 1993). The stimulus is processed by the sensory apparatus an d presented to the dual quantum system/classical measurement apparatus. The state of the quantum machinery expands as a coherent superposition, and all the classical measurement apparatuses that couple with it also become coherent superpositions. Of particular importance is the fact that some of the amplifying role is played by apparatuses that are basically of the same size as the quantum system (as particularly emphasized by Stapp ), leading to a blurring of the distinction between the quantum system an d the amplifying apparatus. This makes a tangled hierarchy. When consciousness collapses the possibility waves of this tangled-hierarchical system of the brain mind, self-reference, the quantum-self experience, arises 2E

Note that I am not introducing a classical/quantum dichotomy. Ultimately, all objects are quantum objects; therefore, a classical measurement apparatus can never really "measure" a quantum object. It itself is described by possibility waves and thus cannot resolve another object's possibility wave into actuality. However, because of its large mass, we c an simultaneously assign both position and momentum to it, albeit approximately; this is Bohr's correspondence principle. This property allows us to use a classical measurement apparatus to amplify the quantum possibilities for us to "see" and to make a memory of what we see.

Two Self-Identities

Experiences lead to learning, one aspect of which is developmental changes in the brain-mind's classical substructure--the memories and re presentations of experience. Additionally, I believe that something profound also takes place in the self-referential quantum/classical coupled system.

There is a difference between an ordinary quantum measurement, such as the measurement of an electron by a Geiger counter, and the quantum measurements that take place in the brain. The former is a one-shot affair; but, because of memory replay, the quantum machinery of the brain-min d can be measured repeatedly for the same stimulus by its measurement apparatuses (this gives rise to secondary-awareness events upon collapse). The quantum equation of the self-referential system of the brain-mind mus t continually be modified by the repeated measurement interactions. Mark Mitchell and I (Mitchell and Goswami, 1992) have proposed that the modified equation is nonlinear (technically called a nonlinear Schrodinger equation) as opposed to the ordinary quantum equation (the Schrodinger equation) which is linear.

Non-linearity means that the current value of the possibility wave function depends in a complicated way on its value at a previous time. We find that, as a result of the feedback from this non-linearity, the quantum possibility waves of a self-referential system gradually become conditioned; the probability of actualizing formerly experienced states gradually gains greater weights. (Similar conditioning has been theorized for photons in a resonant cavity, also, due to the non-linearity of their Schrodinger equation; Carmichael, 1993.)

A well-known characteristic of learning is that learning a performance reinforces the probability of the same subsequent performance. This is the effect we see here. In essence, learning increases the likelihood that, after the completion of measurement, the quantum-mechanical state s of the tangled-heirarchical quantum-system/measurement-apparatus will correspond to a prior learned state. In other words, learning biases the quantum dynamics of the brain-mind and thus reduces the access to its full potentia. Before learning, the possibility pool from which consciousness chooses its states spans the mental states common to all people at all places at all times. With learning, certain responses gradually gain greater weight over others, responses that we call personal.

The especially interesting thing that Mitchell and I found is that in the limit of infinite conditioning of a task, the likelihood that a conditioned stimulus involving that task will trigger a conditioned response approaches one-hundred percent. In this limit, it seems that the behavior of the brain-mind's dual quantum-system/measuring-apparatus approaches classical determined behavior. This is the brain-mind analog of the correspondence principle. In this limit, quantum functionalism approaches conventional cognitive science (Goswami, 1995).

Fairly early in our physical development, learning accumulates an d conditioned response patterns begin to dominate the brain-mind's behavior, despite the fact that the versatility of the quantum system is always available for new creative experiences of what I call primary awareness, which involves the quantum-self modality. When the creative potency of the quantum system is not engaged, when the primary awareness events are not attended, the secondary-awareness processes of memory-replay dominate ; the tangled hierarchy of the systems of the brain-mind, in effect, becomes a simple hierarchy of the learned programs--the representations of pa st experiences. At this stage, the creative uncertainty as to "who the chooser is" of a conscious experience involving the quantum self diminishes. Then we begin to identify with a separate, individual self, the ego, that perceives apparent continuity in the form of a stream of consciousness, that thinks it chooses on the basis of its past experiences, that presumably has "free will." But in truth, this so-called free will of the e go-identity exists only to the extent that the conditioning can only approach 100%, never reach it.

Incidentally, the experiments of neurophysiologist Benjamin Libet and his collaborators (1979) have demonstrated that there is almost half a second of time delay between the primary event of quantum collapse and our verbal awareness of the event. This half-second is the time taken for secondary awareness processing.

III. Qualia

Suppose you have several people experience the sight, sound, and t ouch of a red car with its engine running. Suppose also that you have available to you the right combination of some super technology and high-power mathematics so that you are able to make a complete description of th e neuronal states of the brains of your subjects, even one for your own b rain, upon experiencing the car. Except for minor differences, you would expect the neuronal configurations of all the brains, including yours, t o be identical.

And yet, you know that in the case of your brain, something is left out, something that the objective neuronal configurations cannot possibly describe, and that is your subjective experience as observer. You might say that this comprises something special in the neuronal configuration of your observer brain compared to all the observed brains. But then you would be admitting (barring solipsism) that your conscious experience of your brain state changes your supposedly objective brain state. The alternative is to admit that the neuronal configuration does not provide a complete description of the experience (Goswami, 1994).

This is the paradox of self-reference back again, and without addressing it, it is impossible to address the issue of the qualia of experience. The theory of quantum functionalism above, having addressed the paradox of self-reference, thus also successfully eradicates the paradox of the qualia of experiences. In this theory, the processing of the incoming stimulus involves quantum processes and their amplification at every stage, leading to a macroscopic coherent superposition of possibilities until consciousness supervenes. Because consciousness of the experience transcends the brain-state of the quantum/classical ensemble, the latter is clearly an incomplete description of the experience. Yet the subject consciousness of the experience (the subject pole with the qualia of experience) arises co-dependently and tangled-hierarchically with the chosen bra in-state (the object pole), both of which exist only as possibility until the collapse, and no dualism is involved. Nor is it necessary to invoke an operational negation of the authenticity of qualia (Dennett, 1991).

The qualia of the primary experience is basically universal, and thus objective in some sense, but secondary subjective and personal qualities arise from secondary-awareness processing--the reflection from the mirror of individual brain memory. The idea of consciousness self-referentially collapsing both the object pole of the experience and the subject pole, where the quality of the experience lies, also resolves the thorny issue of the oneness of a conscious experience--the idea that we can be consciously aware at any given instant of only one particular thing. It is well-known that all attempts by psychologists and neuro-physiologists to split the unity of a conscious experience (for example, by surgically splitting the brain hemispheres) have failed. In the quantum functionalist scenario there is only one collapse at a given time, and that defines the event in awareness. However, the secondary-awareness events that follow each of two consecutive primary events can overlap in time. This is responsible for the fact that we can be peripherally aware of several things in awareness at the same time.

In this way we see that an understanding of the qualia of experience can be obtained within the extended context of quantum functionalism. We do not need to resort to operationalism and negate qualia altogether, as Dennett has done.

IV. Unconscious Perception

The quantum theory distinguishes between conscious and unconscious perception. Ordinary perception consists of the collapse of a possibility wave by consciousness (via recognition and choice) in the presence of awareness. But in unconscious (subliminal) perception, in which consciousness but not awareness is present, there should be no collapse of the w ave, according to our quantum model. Unconscious processing is found to be of crucial importance in the creative process, for which a quantum explanation has been given (Goswami, in press).

Cognitive experiments using polysemous words seem to verify this aspect of the quantum model. In a representative experiment, Tony Marcel (1980) used strings of three words in which the middle word was polysemous; his subjects watched a screen as the three words were flashed one at a time at intervals of either 600 milliseconds or 1.5 seconds between flashings. The subjects were asked to push a button when they consciously recognized the last word of the series. The original purpose of the experiment was to use the subject's reaction time as a measure of the relations hip between congruence (or lack of it) among the words and the meanings assigned to the words in such series as hand-palm-wrist (congruent), clock -palm-wrist (unbiased), tree-palm-wrist (incongruent), and clock-ball-wrist (unassociated). For example, the bias of the word hand followed by th e flashing of palm may be expected to produce the hand-related meaning of palm, which then should improve the reaction time of the subject for recognizing the third word wrist (congruence). But if the biasing word is t ree, then the lexical meaning of palm as a tree would be assigned and the meaning-recognition of the third word wrist should take a longer reaction time (incongruous). And indeed, this was the result.

However, when the middle word was masked by a pattern that made i t impossible to see with awareness though unconscious perception continue d, there was no longer any appreciable difference in reaction time between the congruent and the

incongruent cases. This is surprising because presumably both meaning s of the ambiguous word were available, regardless of the biasing context , yet neither meaning was chosen over the other. Apparently, choice, and therefore quantum collapse, is a concomitant of conscious experience but not of unconscious perception. It is our consciousness that chooses--we choose, therefore we are--but we choose only when awareness is present.

If this quantum explanation of the Marcel experiment is correct, then the experiment also demonstrates the existence of coherent superpositions in the brain-mind. Before choice, the quantum description of the ambiguous state of the brain-mind subject to a pattern-masked, polysemous word-stimulus is none other than a coherent superposition.

It is certainly not ruled out, however, that an explanation of the Marcel experiment can be given using computer models of cognition (called connectionism; see Rummelhart et al, 1985), although such a model will probably have to use the unattractive idea of consciousness being a central processing unit (smacking of a homunculus) in order to distinguish between conscious and unconscious perception (Posner and Klein, 1973; Posner and Schneider, 1975 a, b). Thus the question arises, Is there an unambiguous way to discern between a quantum and a classical computer model of cognition? 

The Possibility of Quantum Interference in the Brain-Mind

Is the play of probabilities in the brain-mind classical or quantum? This question, I think, can be answered with an even more decisive objective experiment that can discern between classical and quantum models of the mind.

There is a distinct mathematical difference between classical and quantum probabilities. In quantum mechanics, the probability amplitudes are added algebraically before squaring to find the net probability, for example, of the result of the passage of electrons through a double-slit arrangement. In classical physics, on the other hand, the probabilities simply add. It is the addition of amplitudes before squaring them that gives the phenomenon of quantum interference.

Suppose we ask subjects to look at a screen on which a polysemous word (such as palm in Marcel's experiment) is flashed that has two possible interpretations, A and B. The quantum state of the subjects' brains would then become a coherent superposition in response to the ambiguous signal, fifty percent for recognizing A and fifty percent for B (assuming equal probability for the two possible responses to the picture). Now suppose we go one step further. Suppose, next, the subjects are shown another ambiguous word, again with two different interpretations C and D. Suppose the probability of choosing C after having chosen A is P1 and after having chosen B is P2 (P1 and P2 can be measured by repeated experiments with several subjects).

Then we bring in a new sample of subjects and flash them first th e ambiguous word with interpretations A and B, followed quickly by the on e with interpretations C and D. Only after seeing both words are they asked for their interpretations. Since probabilities are multiplicative, i f the probabilities are classical, then the total probability for choosing C for all subjects will be .5(P1 + P2). Any deviation from this estimate will tell us about quantum interference and thus about the quantum nature of the probabilities and, thereby, of choice (Woo, 1981).

However, there are some uncertainties in doing such an experiment ; for example, how do we guarantee that a subject retains the dichotomy ( i.e., doesn't make a choice) between seeing the first word and the second ? In fact, if the idealist interpretation of quantum mechanics is correct, then each sight-recognition collapses the ambiguity of the stimulus.

But we can think of a very plausible way to carry out a successful experiment. We can take advantage of the fact that the unconscious min d does not choose unless there is an observation with awareness, as discussed previously. If we mask the ambiguous words with a pattern, then the subjects can see the words only with unconscious awareness (as in Marcel 's experiment). This guarantees that no choice is made after seeing the words, no collapse of the dichotomy, until we ask, and there should be interference. A thorough analysis (McCarthy and Goswami, 1993) shows that, indeed, if the two ambiguous words are shown simultaneously and with a pattern mask, then, due to quantum interference, the recognition times for the target word can be drastically different from what is predicted by connectionist models. Thus this experiment should be able to establish beyond any reasonable doubt the existence of quantum coherent superpositions in the brain-mind.

V. Quantum Non-locality in the Brain-Mind 

One of the principal aspects of quantum functionalism is non-locality. Evidence for quantum non-locality of our experience abounds in the literature of paranormal phenomena (see, for example, Jahn, 1982).

The recent experiment by the Mexican neurophysiologist Jacobo Grinberg-Zilberbaum and his collaborators directly supports the idea of non -locality in human brain-minds--this experiment is the equivalent for bra ins of the objective Aspect et al's (1982) experiment. Two subjects are instructed to meditate together for a period of twenty minutes in order t o establish a "direct communication"; then they enter separate Faraday chambers (metallic enclosures that block electromagnetic signals) while maintaining their direct communication for the duration of the experiment. One of the subjects is now shown a light flash that produces an evoked potential (an electro-physiological response produced by a sensory stimulus measurable by an EEG) in the stimulated brain. But amazingly, in about one in four cases, the unstimulated brain also shows an electrical activity, a "transferred" potential quite similar in shape and strength to the evoked potential. (Control subjects never show any transferred potential 2E) The straightforward explanation is quantum non-locality--the two brain-minds act as a non-locally correlated quantum system. In response to a stimulus to only one of the correlated brains, consciousness collapses identical states in the two brains, hence the similarity of the brain potentials (Grinberg-Zylberbaum et al, 1994).

Clearly, a radical hypothesis is dictated by the experimental data on telepathy, distant viewing, transferred potential, and the like. How can local signals perceived by one observer be perceived also by another without some other local signals? Because consciousness may choose to collapse identical possibility waves simultaneously in two correlated locally-separated observers.

There is a striking similarity between correlated brains (as in th e Grinberg-Zilberbaum experiment) and correlated photons as in the Aspect experiment, but there is also a striking difference. The similarity is that in both cases the initial correlation is produced by some "interaction." But the difference is that in the former case, as soon as the wave function is collapsed by measurement, the objects become uncorrelated; but in the case of the correlated brains, consciousness maintains the corr elation over the one-hundred light flashes that are needed to get the ave rage evoked potential. This difference is highly significant. The non-locality of correlated photons, although striking in terms of demonstrating the radicalness of quantum physics, cannot be used to transfer information, according to a theorem attributed to Philippe Eberhard. But in the case of the correlated brains, since consciousness is involved in making and maintaining the correlation, Eberhard's theorem does not apply, and message transfer is not forbidden.

Non-local consciousness collapsing correlated quantum wave functions at different brain areas, simultaneously giving rise to an event of fe lt experience, is the simplest answer to the binding problem.

VI. A New Psychophysical Parallelism

So far, I have dealt with only one of the two ontological problems with the conventional approach to consciousness. Conventionally, Wester n philosophers attribute properties of consciousness--experience and choice--to the mind. This has been corrected in quantum functionalism in which consciousness is defined to transcend both matter and mind.

But there is a second part to the mind-body problem. It is about the "minding" that mind does--thinking, feeling, and so forth. Material realists assume that these mental properties emerge as higher order functions of brain-matter--mind is the software function of brain hardware. Even quantum functionalism implicitly goes along with this idea. But I no w think this is not enough. Dualists, to their credit, have always insisted that mind and brain are fundamentally different (Eccles, 1994). Their legitimate claim has been overlooked because the difficulties of dualism in the form of Cartesian interactionism have been considered to be insurmountable. I will now show how this difficulty can be overcome (in a manner quite different from Eccles (1994) and yet that captures his intention).

The truth is, except for quantum measurement, matter is law-like-- cause driven. The time development of matter is entirely given by the laws of quantum mechanics. Mind, on the other hand, is program-like--it is driven not only by causes but also by purpose. The programs of the mind can be simulated by computer algorithms, hence the temptation of assuming that mind is reducible to matter, "mind is brain". But then where does the purposiveness of mind come from? Logic dictates that only conscious ness can inject purposiveness in the world. Thus it makes more sense to hypothesize that consciousness "writes" the purposive mental programs in the brain. When we write software for our personal computer, we employ our mental picture of what we want to do in the programming. Similarly, consciousness must use a "mental body" to create the mental software of th e brain.

Thus the analysis and explanation of mentation calls for a radical hypothesis: there is, in reality, more than one substance-body. (The w ord substance here is to be understood in the broad sense of a mode of be ing.) Along with our material body, we also have a subtle body consisting of a mental substance that also obeys quantum possibility dynamics. In accord with conservation principles that conventional science has established, this subtle substance does not interact with the material substance in any direct way; we are not reviving Cartesian interaction dualism. And yet the subtle substances can communicate with the physical substance through the intermediary of consciousness. Consciousness can simultaneously collapse possibility waves in the correlated subtle mental and physical bodies of an individual.

In the philosophy of psychophysical parallelism, it is assumed that the physical body and the corresponding mental correlate go on simultaneously. Psychophysical parallelism obviously has the virtue that we do n o need the awkward assumption that mental states emerge from physical states. The mental body does what mind is for, thoughts and feelings. The physical brain does what it is made for: movement in neurons that causes physical action. Psychophysical parallelism avoids the problem of interaction between the two different substrata, but no mechanism is given for the parallelism to come about. The current model, with two separate substances that are connected via consciousness which simultaneously collapse s parallel actualities in both bodies, gives a mechanism for psychophysical parallelism. However, the new model is different from the old psychophysical parallelism in the sense that experience modifies both bodies as states of the two bodies become correlated by experience. We will dwell upon this point further with an example--perception (see below).

Note that the new hypothesis is postulating a new psychophysical parallelism, but firmly within a monistic idealist ontology. Both subtle and physical worlds remain in possibility until consciousness self-referentially collapses the possibility structure into actuality.

Perception

In the conventional cognitive-science model of perception, it is assumed that the brain makes a mental representation of a sensory object, an image, which is what we see. But there are many inherent difficulties in such a model. Some of these difficulties I have already discussed earlier in connection with experience.

To recap briefly, the main problems with the representational mode l of perception are these:

(1) The model implies a dualistic homunculus in our head watching the mental representational show offered to it; otherwise, how does the subject-object experience of the watching come about?

(2) The brain representation of a perceived object invariably involves many brain areas, but the experience of perception is one of unity; we don't perceive all the different aspects of imaging separately, they a re all integrated. With local mechanisms alone, it is hard to solve this "binding" problem (Feynman, 1981).

(3) With a strictly objective model of perception, it is hard to explain the subjective quality of experience referred to as "qualia" by philosophers.

(4) Still another difficulty is that the image in our brain after processing by the "higher" centers is not an exact replica of the object, and yet, somehow we are able to translate the image into the object, and in such a way as to form a consensus with other observers (although it i s by no means clear that what we see in consensus viewing is necessarily the object in its suchness, a point made by the philosopher Immanuel Kant ). As Eccles (1994) notes, How does spatio-temporal neuronal activity in the cerebral cortex evoke a perception of the object in the mind?

(5) And perhaps most importantly, the mental aspects of perception , the mental representation and associated awareness, thoughts, concepts, and other mental objects are internal and private in contrast to the physical aspect of perception, the external, public object which we share with other people.

This last problem has driven philosophers called naive realists to chuck the entire idea of mental representations. Instead, they assert t hat we see a thing by directly interacting with it--direct perception (Searle, 1994). Unfortunately, there is no model, no analog, of how physical interaction between the physical brain and a physical external object c an produce conscious perception and mental experience.

Now, the representational model would be more suited if perception is a matter of recognition, not cognition. The breakthrough in pattern recognition by computers came when it was realized that instead of programming computers to cognize a pattern, we can "teach" them to do so. In other words, a computer can easily recognize an object by comparing it wit h its "learned" representations. So who teaches the human biocomputer, the physical brain, its representations? Well, consciousness does--with the help of the states of the mental body. An unlearned stimulus produces an image in the physical brain in the form of possibilities of the quantum brain, but these possible images have no mental meaning. Consciousness ascribes mental meaning to the image with the help of mental states o f the mental body; only then, when consciousness recognizes and chooses a correlated pair of states of the physical brain and the mental body, is a meaningful representation made. The quantum measurement simultaneously also collapses the external object (which, before collapse, according to quantum mechanics, is also only "a pattern of disjoint tendencies") in physical space-time. In other words, there is also direct perception of t he object in space, as the naive realists theorize. Subsequently, once a representation is made, it is a computer-style operation. A learned stimulus elicits memory, which is reconstructed every time the stimulus is p resented and processed. When consciousness recognizes a learned state in its quantum possibilities of the physical brain, it also recognizes and chooses the correlated mental state. Thus in the process of perception, representations of the physical world are made not only in the physical b rain-mind but also in the mental body through modification of the probabilities of the experienced mental possibilities. In the reciprocal process of imagination, a physical brain-mind representation is made of subtle mental states. Mind and brain affect one another, as Wigner (1967) maintained. In this way, perception produces not only physical representations or memory in the physical brain but also a tendency in the mental body f or certain correlated states to collapse when a particular physical stimulus is presented. In this way the new psychophysical parallelism propose d here is quite different from previous forms. Actually, it is quite similar in this respect to Sheldrake's (1982) theory, a parallel that is being explored and will be published elsewhere.

Why the mental image is private

What is the difference between gross physical and subtle substances? One big difference has to be the grossness of the macroworld of our shared perception in the physical domain.

Quantum objects obey the uncertainty principle--we cannot simultaneously measure both their position and momentum (mass times velocity) wit h utmost accuracy. Now in order to determine the trajectory of an object , we need to know where an object is now but also where it will be a litt le later; in other words, both position and velocity, simultaneously. So we can never determine accurate trajectories of quantum objects.

Although the macrobodies of our environment are made of the micro quantum objects that obey the uncertainty principle, because of their grossness, the cloud of ignorance that the uncertainty principle imposes on their motion is very small, so small that it can be ignored in most situations--this is called the correspondence principle. Thus macro bodies ca n be approximately attributed both position and momentum and therefore, trajectories. We need the intermediary of the macrobodies, a macro "measurement" apparatus, to amplify the micro quantum objects before we can observe them. This is the price we pay so that we have a shared reality of physical objects; everybody can simultaneously see the macrobodies.

Mental substance is subtle, it does not form gross conglomerates. In fact, as Descartes correctly intuited, mental substance is indivisible. For this substance, then, there is no reduction to smaller and smaller bits, there is no micro out of which the macro is made of.

So the mental world is a whole, or what physicists sometimes call an infinite medium. There can be waves in such infinite media, modes of movement that must be described as quantum possibility waves obeying a probability calculus.

We can directly verify that thoughts--mental objects--obey the uncertainty principle; we can never simultaneously keep track of both the content of a thought and where the thought is going--the direction of thought (Bohm, 1951). For thoughts, we can directly observe them without any intermediary, but the price is that thoughts are private, internal; we cannot normally share them with others.

VII. Comparison with Other Theories of the Same Ilk

As is well known, the idea of consciousness collapsing the quantum wave function was originally proposed by von Neumann (1955) who had some thing like psychophysical parallelism in mind. Subsequently, Wigner (196 2, 1967) took von Neumann's idea further and envisioned that mind and bra in should affect each other. The theory developed here, with the component of psychophysical parallelism, fulfills, I think, both von Neumann's and Wigner's visions but without the pitfalls of dualism.

Without the idea of psychophysical parallelism, the theory is closest in spirit to that of Stapp (1982, 1993). Stapp talks about felt experience arising from a quantum measurement, whereas I spell out how this c an be so (a tangled hierarchy of quantum measurement giving the subject-object split of self-reference). Stapp avoids the explicit monistic idealist ontology, perhaps because of the mystical connotation of this philosophy. But I think the mystical connection is a virtue, for the present development can be used to bridge science and spirituality (Goswami, 1993).

With psychophysical parallelism included, the theory is closest to that of Eccles (1994) except that some of the difficulties of the latter (for example, an implicit homunculus) have been removed via the idea of self-referential quantum measurement.

Stapp (1993) has criticized Eccels's (1986, 1994) theory on account of the following data. Certain brain damage eliminates all feeling in some part of the body. When told of the existence of an arm, for example , such a patient shows surprise. Stapp thinks this as evidence against t he existence of a separate mental body. It may, indeed, be evidence against a homunculus, as Eccles assumes, because the homunculus has the right representation still available to it. However, the situation is different in the present theory. With brain damage, the learned representation of that particular body part is gone. Thus, according to the present mod el, surprise is natural. The crucial question is, Can the patient learn the correct body image again? If there is a mental body, the answer must be yes.

I will end this paper by pointing out that the idealist science within consciousness reviewed and developed here is versatile and has already been applied to living cells (Goswami, 1994). As another application, a mental body which is conditioned by experience opens up the notion that some aspects of us may survive death. In other words, the doctrine that if the brain dies, the mind dies can be challenged. This subject is no w under study and the results will be published elsewhere (Goswami, 1995) 2E

I would like to thank Drs. Robert Thomkins and Jean Burns, and Kirsten Larsen for careful readings of the manuscript. Thanks are due to Hugh Harrison for helpful discussions and to Maggie Goswami for a careful editing.

References

Aspect, A., Dalibard, J., and Roger, G. (1982), 'Experimental test of Bell inequalities using time-varying analyzers', Physical Review Letters , 49, pp. 1804-06.

Assagioli, R. (1976), Psychosynthesis: A Manual of Principles and Techniques (New York: Penguin).

Bass, L. (1975), 'A quantum mechanical mind-body interaction',

Foundations of Physics, 5, pp. 155-172.

Blood, C. (1993), 'On the relation of the mathematics of quantum mechanics to the perceived physical universe and free will', preprint, Rutgers University, Camden, N. J.

Bohm, D. (1951). Quantum Theory. Englewood Cliffs, NJ: Prentice hall 2E

Burns, J. E. (1991), 'Contemporary models of consciousness: Part I', Journal of Mind and Behavior, 12, pp. 407-420.

Carmichael, H. (1993), Private communication.

Dennett, D. C. (1991), Consciousness Explained (Boston: Little Brown).

Donald, M. J. (1990), 'quantum theory and the brain', Proceedings of Royal Society of London, 427A, p. 43.

Eccles, J. C. (1986), 'Do mental events cause neural events analogously to the probability fields of quantum mechanics?' Proceedings of the Royal Society of London, B227, 411-428.

Eccles, J. C. (1994), How the Self Controls Its Brain (Berlin: Springer Verlag).Feynman, R. P. (1981), 'Simulating physics with computers', International Journal of Theoretical Physics, 21, pp. 114-123.

Goswami, A. (1989), 'Idealistic Interpretation of Quantum Mechanics', Physics Essays, 2, pp. 385-400.

Goswami, A. (1990), 'Consciousness in quantum physics and the mind-body problem', Journal of Mind and Behavior, 11, pp. 75-96.

Goswami, A. (1993), The Self-Aware Universe: How Consciousness Creates the Material World (N.Y.: Tarcher/Putnam).

Goswami, A. (1994), Science within Consciousness (Sausalito, CA: Institute of Noetic Sciences Research Report CP-2).

Goswami, A. (1995), 'Monistic idealism may provide better ontology to cognitive science: a reply to Dyer', Journal of Mind and Behavior, 16, 13 5-150.

Goswami, A. (in press), 'Creativity and the quantum', Creativity Research Journal.

Grinberg-Zylberbaum, J., Delaflor, M., Attie, L., and Goswami, A. (199 4), 'The Einstein-Podolsky-Rosen paradox in the brain: the transferred potential', Physics Essays, 7, pp. 422-428.

Herbert, N. (1993), Elemental Mind (N.Y.: Dutton).

Hofstadter, D. (1980), Godel, Escher, Bach: An Eternal Golden Braid (N.Y.: Basic).

Jahn, R. (1982), 'The persistent paradox of psychic phenomena: an engineering perspective', Proceedings of the IEEE, 70, pp. 135-170.

Libet, B., Wright, E., Feinstein, B., and Pearl, D. (1979), 'Subjective referral of the timing for a cognitive sensory experience', Brain, 102 , 193.

Lockwood, M. (1989), Mind, Brain, and the Quantum (Oxford, U.K.: Black well).

Marcel, A. (1980), 'Conscious and preconscious recognition of polysemous words: locating the selective effects of prior verbal context', in Attention and Performance VIII, ed. R. S. Nickerson (Hilsdale, New Jersey: Erlbaum).

McCarthy, K., and Goswami, A. (1993), 'CPU or self-reference? Can we discern between cognitive science and quantum functionalist models of mentation?' Journal of Mind and Behavior, 14, pp. 13-26.Mitchell, M. and Goswami, A. (1992), 'Quantum mechanics for observer systems', Physics Essays , 5, pp. 526-529.

Penrose, R. (1994), Shadow of the Mind (Oxford, U.K.: Oxford University Press.

Posner, M. I., and Klein, R. (1973), 'On the functions of consciousness, in Attention and Performance IV, ed. S. Kornblum (New York: Academic Press).

Posner, M. I., and Snyder, C. R. R. (1975a), 'Attention and cognitive control', in Information Processing and Cognition: The Loyola Symposium, ed. R. L. Solso (New Jersey: Erlbaum Press).

Posner, M. I., and Snyder, C. R. R. (1975b), 'Facilitation and inhibit ion in the processing of signals', in Attention and Performance V, ed. P 2E M. A. Rabbit (New York: Academic Press).Rumelhart, D. E., McClelland, J. L., and the PDP Research Group (1986), Parallel Distributed Processing: Explorations in the Microstructure of Cognition (Volumes 1 and 2 ) (Boston: The MIT Press).

Searle, J. (1987), 'Minds and brains without programs', in Mind Waves, ed. C. Blackmore and S. Greenfield (Oxford: Basil Blackwell).

Searle, J. (1994), The Rediscovery of the Mind (Cambridge, MA: The MIT Press).

Sheldrake, R. (1981), A New Science of Life (L.A.: Tarcher).

Stapp, H. P. (1982), 'Mind, matter, and quantum mechanics', Foundation s of Physics, 12, pp. 363-398.

Stapp, H. P. (1993), Mind, Matter, and Quantum Mechanics (N.Y.: Springer).

Stuart, C. I. J. M., Takahashy, Y., and Umezwa, M. (1978), 'Mixed system brain dynamics', Foundations of Physics, 9, pp. 301-329.

von Neumann, J. (1955), The Mathematical Foundations of Quantum Mechanics (Princeton: Princeton University Press).

Walker, E. H. (1970), 'The Nature of consciousness', Mathematical Bio sciences, 7, pp. 131-178.

Wallas, G. (1926), The Art Of Thought (N.Y.: Harcourt, Brace, and World).

Wigner, E. P. (1962), In The Scientist Speculates, ed. I. J. Good (Kindswood, Surrey, U.K.: The Windmill Press).

Wigner, E. P. (1967), Symmetries and Reflections (Bloomington: Indiana University Press).Wolf. F. A. (1984), Starwave (N. Y.: MacMillan).

Woo, C. H. (1981), 'Consciousness and quantum interference', Foundations of Physics, 11, pp. 933-944.

Fall 1997 | Spring 1997 | Fall 1996 | Spring 1996 | Other Essays