July 21, 1998 LBNL-40722 Quantum Ontology and Mind-Matter Synthesis Henry P. Stapp Lawrence Berkeley National Laboratory University of California Berkeley, California 94720 \begin{abstract} The Solvay conference of 1927 marked the birth of quantum theory. This theory constitutes a radical break with prior tradition in physics, because it avers, if taken seriously, that nature is built not out of matter but out of knowings. However, the founders of the theory stipulated, cautiously, that the theory was not to be taken seriously, in this sense, as a description of nature herself, but was to be construed as merely a way of computing expectations about future knowings on the basis of information provided by past knowings. There have been many efforts over the intervening seventy years to rid physics of this contamination of matter by mind. But I use the reports at this Symposium to support the claim that these decontamination efforts have failed, and that, because of recent developments pertaining to causality, the time has come to take quantum theory seriously: to take it as the basis for a conception of the universe built on knowings, and other things of the same kind. Quantum theory ensures that this conception will yield all the empirical regularities that had formerly been thought to arise from the properties of matter, together with all of those more recently discovered regularities that cannot be understood in that mechanical way. Thus I propose to break away from the cautious stance of the founders of quantum theory, and build a theory of reality by taking seriously what the incredible accuracy of the predictions of the formalism seems to proclaim, namely that nature is best understood as being built around knowings that enjoy the mathematical properties ascribed to them by quantum theory. I explain why this idea had formerly been incorrectly regarded as untenable, due to a failure to distinguish signals from influences: relativistic quantum field theory ensures both that signals cannot travel faster than light, but that influences, broadly conceived, cannot be imagined to enjoy that property. Failure to recognize this fact had made a realistic interpretation of quantum theory seem impossible. I then explain how our conscious knowings can play a causally efficacious and binding role in brain dynamics without violating the statistical rules of quantum theory, and describe how these features provide a foundation for understanding how consciousness could have evolved by natural selection from primitive beginnings. \end{abstract} Invited Paper: The X-th Max Born Symposium ``Quantum Future''. 1. Introduction. The modern era was created probably as much by Descartes' conceptual separation of mind from matter as by any other event. This move freed science from the religious dogmas and constraints of earlier times, and allowed scientists to delve into the important mathematical regularities of the observed physical world. Descartes himself allowed interaction between mind and matter to occur within the confines of a human brain, but the deterministic character of the physical world specified later by Newtonian mechanics seemed to rule out completely, even within our brains, any interference of mind with the workings of matter. Thus the notion of a completely mechanical universe, controlled by universal physical laws, became the new dogma of science. It can readily be imagined that within the milieu dominated by such thinking there would be stout opposition to the radical claims of the founders of quantum theory that our conscious human knowings should be taken as the basis of our fundamental theory of nature. Yet the opposition to this profound shift in scientific thinking was less fierce than one might suppose. For, in the end, no one could dispute that science rests on what we can know, and quantum theory was formulated in practical human terms that rested squarely on that fact. Hence the momentous philosophical shift was achieved by some subtle linguistic reformulations that were inculcated into the minds of the students and practitioners of quantum theory. The new thought patterns, and the calculations they engendered, worked beautifully, insofar as one kept to the specified practical issues, and refrained, as one was instructed to do, from asking certain ``meaningless'' metaphysical questions. Of course, there are a few physicists who are dissatisfied with purely practical success, and want to understand what the practical success of these computational rules is telling us about ourselves and the nature of the world in which we live. Efforts to achieve such an understanding are proliferating, and the present work is of that genre. Historically, efforts to achieve increasingly coherent and comprehensive understandings of the clues we extract from Nature have occasionally led to scientific progress. The outline of the present work is as follows. In section 2, I document the claim made above that the orthodox Copenhagen interpretation of quantum theory is based squarely and explicitly on human knowings. The aim of the paper is to imbed this orthodox pragmatic epistemological theory in a rationally coherent naturalistic ontology in a minimalistic way that causes no disruption of anything that orthodox quantum theory says, but merely supplies a natural ontological underpinning. In the special case of processes occurring in human body/brains this ontological structure involves human conscious knowings that enter into the brain dynamics in a manner that accounts for the way that these knowings enter into the orthodox interpretation of quantum theory. In section 3 I discuss another interpretation, which is probably the common contemporary interpretation of the Copenhagen interpretation. It is coarse in that it is imprecise on essential theoretical points. Because it is common and coarse I call it the Vulgar Copenhagen Interpretation. In section 4 the unusual causal structure of quantum theory is discussed, and is used to justify, in the context of trying to understand the role of mind in nature: 1) the rejection of the classical ontology, 2) the reasonableness of attempting to ontologicalize the orthodox interpretation of quantum theory, and 3) the expectation that our knowings involve non-local aspects. Section 5 is entitled ``All roads lead to Solvay 1927''. The 1927 Solvay conference, seventy years ago, marked the birth of the orthodox Copenhagen interpretation of quantum theory. In this section I review this Symposium from a certain point of view, namely the viewpoint that many of the highlights of the Symposium confirm the basic message of the orthodox interpretation, namely that the only reasonable way to make rational sense out of the empirical data is to regard nature as being built out of knowings. I argue that the experience of the last seventy years suggests the reasonableness of taking this interpretation seriously: more seriously than the founders of quantum theory took it. Basically, they said, cautiously, that the mathematical formalism is a useful tool for forming expectations about our future knowings on the basis of our past ones. That claim has been now been abundantly confirmed, also in fields far beyond the narrow confines of atomic physics. But the founders scrupulously avoided any suggestion that this mathematical formalism corresponded to reality. They either discouraged us from asking questions about what is really happening, or, if pressed, looked for reality not in their own knowledge-based formalism, but in terms of more conventional physical terms. This reluctance to take their own formalism seriously was, I think, the result partly of an inertial carry-over from classical physics, which shunned and excluded any serious consideration of mind in physics, and partly of a carry-over of an idea from the special theory of relativity. This is the idea that no influence or signal could propagate faster than light. However, in quantum theory there is a sharp distinction between signal and influence, because it can be proved both that no signal can be transmitted faster than light, and that this property cannot be imagined to hold for influences. The distinction between signal and influence has to do with the difference between the causal structure of the deterministic evolution of the {\it statistical predictions of the theory} and the causal structure of something that has no analog in classical mechanics, namely the {\it selection process} that acts within the deterministic structure that is the analog of the classical deterministic structure, but that is not fully determined by that structure. In cosmological solutions in general relativity there is usually a preferred set of advancing spacelike surfaces that provide a natural definition of instantaneousness. Also, there is the empirical cosmological preferred frame defined by the background black-body radiation. So the idea of special relativity that there is no preferred frame for the universe, although it may indeed hold for the formulation of the general local-deterministic laws, is not as compelling now as it was in 1905, or even 1927: that idea could very well break down in our particular universe at the level of the selection of particular individual results (knowings). Indeed, I believe it {\it must} break down at that level. (Stapp, 1997) So I propose to take seriously the message of Solvay 1927, that nature be understood as built out of knowings. But we must then learn how better to understand knowings, within the mathematical framework provided by the quantum formalism. In section 6 I distinguish the two different components of the quantum mechanical evolutionary process, the unitary/local part and the nonunitary/ nonlocal part, and note that our conscious knowings, as they occur in the quantum description, enter only into the latter part. But that part is eliminated when one takes the classical approximation to the quantum dynamics. Thus from the perspective of quantum mechanics it would be irrational to try to find consciousness in a classical conception of nature, because that conception corresponds to an approximation to the basic dynamics from which the process associated with consciousness has been eradicated. I note there also that the ontologicalization of the quantum mechanical description dissolves, or at least radically transforms the mind-matter dualism. The reason is this: in the classical theory one specifies at the outset that the mathematical quantities of the theory represent the physical configuration of matter, and hence one needs to explain later how something so seemingly different from matter as our conscious knowings fit in. But in the quantum case one specifies from the outset that the mathematical quantities of the theory describe properties of knowings, so there is no duality that needs explaining: no reality resembling the substantive matter of classical physics ever enters at all. One has, instead, a sequence of events that are associated from the outset with experiences, and that evolve within a mathematically specified framework. Section 7 lays out more explicitly the two kinds of processes by showing how they can be considered to be evolutions in two different time variables, called process time and mathematical time. Section 8 goes into the question of the ontological nature of the ``quantum stuff'' of the universe. In the sections 9 and 10 I describe the proposed ontology. It brings conscious knowings efficaciously into quantum brain dynamics. The basic point is that in a theory with objectively real quantum jumps, some of which are identifiable with the quantum jumps that occur in the orthodox epistemological interpretation, one needs three things that lie beyond what orthodox quantum theory provides: \noindent 1. A process that defines the conditions under which these jumps occur, and the possibilities for what that jump might be. \noindent 2. A process that selects which one of the possibilities actually occurs. \noindent 3. A process that brings the entire universe into concordance with the selected outcome. Nothing in the normal quantum description of nature in terms of vectors in Hilbert space accomplishes either 1 or 2. And 3 is simply put in by hand. So there is a huge logical gap in the orthodox quantum description, if considered from an ontological point of view. {\it Some extra process, or set of processes, not described in the orthodox physical theory, is needed.} I take a minimalistic and naturalistic stance, admitting only the least needed to account for the structure of the orthodox quantum mechanical rules. In appendix A I show why the quantum character of certain synaptic processes make it virtually certain that the quantum collapse process will exercise dominant control over the course of a conscious mind/brain processes. \vskip .1in \noindent {\bf 2. The subjective character of the orthodox interpretation of quantum mechanics.} In the introduction to his book ``Quantum theory and reality'' the philosopher of science Mario Bunge (1967) said: ``The physicist of the latest generation is operationalist all right, but usually he does not know, and refuses to believe, that the original Copenhagen interpretation --- which he thinks he supports --- was squarely subjectivist, i.e., nonphysical.'' Let there be no doubt about this. Heisenberg (1958a): ``The conception of objective reality of the elementary particles has thus evaporated not into the cloud of some obscure new reality concept but into the transparent clarity of a mathematics that represents no longer the behavior of particles but rather our knowledge of this behaviour.'' Heisenberg (1958b): ``...the act of registration of the result in the mind of the observer. The discontinuous change in the probablitity function... takes place with the act of registration, because it is the discontinuous change in our knowledge in the instant of registration that has its image in the discontinuous change of the probability function.'' Heisenberg (1958b:) ``When old adage `Natura non facit saltus' is used as a basis of a criticism of quantum theory, we can reply that certainly our knowledge can change suddenly, and that this fact justifies the use of the term `quantum jump'. '' Wigner (1961): ``the laws of quantum mechanics cannot be formulated...without recourse to the concept of consciousness.'' Bohr (1934): ``In our description of nature the purpose is not to disclose the real essence of phenomena but only to track down as far as possible relations between the multifold aspects of our experience.'' In his book ``The creation of quantum mechanics and the Bohr-Pauli dialogue'' (Hendry, 1984) the historian John Hendry gives a detailed account of the fierce struggles by such eminent thinkers as Hilbert, Jordan, Weyl, von Neumann, Born, Einstein, Sommerfeld, Pauli, Heisenberg, Schroedinger, Dirac, Bohr and others, to come up with a rational way of comprehending the data from atomic experiments. Each man had his own bias and intuitions, but in spite of intense effort no rational comprehension was forthcoming. Finally, at the 1927 Solvay conference a group including Bohr, Heisenberg, Pauli, Dirac, and Born come into concordance on a solution that came to be called ``The Copenhagen Interpretation''. Hendry says: ``Dirac, in discussion, insisted on the restriction of the theory's application to our knowledge of a system, and on its lack of ontological content.'' Hendry summarized the concordance by saying: ``On this interpretation it was agreed that, as Dirac explained, the wave function represented our knowledge of the system, and the reduced wave packets our more precise knowledge after measurement.'' Certainly this profound shift in physicists' conception of the basic nature of their endeavour, and the meanings of their formulas, was not a frivolous move: it was a last resort. The very idea that in order to comprehend atomic phenomena one must abandon physical ontology, and construe the mathematical formulas to be directly about the knowledge of human observers, rather than about the external real events themselves, is so seemingly preposterous that no group of eminent and renowned scientists would ever embrace it except as an extreme last measure. Consequently, it would be frivolous of us simply to ignore a conclusion so hard won and profound, and of such apparent direct bearing on our effort to understand the connection of our knowings to our bodies. Einstein never accepted the Copenhagen interpretation. He said: ``What does not satisfy me, from the standpoint of principle, is its attitude toward what seems to me to be the programmatic aim of all physics: the complete description of any (individual) real situation (as it supposedly exists irrespective of any act of observation of substantiation).'' (Einstein, 1951, p.667) and ``What I dislike in this kind of argumentation is the basic positivistic attitude, which from my view is untenable, and which seems to me to come to the same thing as Berkeley's principle, {\it esse est percipi}. (Einstein, 1951, p. 669). Einstein struggled until the end of his life to get the observer's knowledge back out of physics. But he did not succeed! Rather he admitted that: ``It is my opinion that the contemporary quantum theory...constitutes an optimum formulation of the [statistical] connections.'' (ibid. p. 87). He referred to: ``the most successful physical theory of our period, viz., the statistical quantum theory which, about twenty-five years ago took on a logically consistent form. ... This is the only theory at present which permits a unitary grasp of experiences concerning the quantum character of micro-mechanical events.'' (ibid p. 81). One can adopt the cavalier attitude that these profound difficulties with the classical conception of nature are just some temporary retrograde aberration in the forward march of science. Or one can imagine that there is simply some strange confusion that has confounded our best minds for seven decades, and that their absurd findings should be ignored because they do not fit our intuitions. Or one can try to say that these problems concern only atoms and molecules, and not things built out of them. In this connection Einstein said: ``But the `macroscopic' and `microscopic' are so inter-related that it appears impracticable to give up this program [of basing physics on the `real'] in the `microscopic' alone.'' (ibid, p.674). The examination of the ``locality'' properties entailed by the validity of the predictions of quantum theory that was begun by Einstein, Podolsky, and Rosen, and was pursued by J.S. Bell, has led to a strong conclusion (Stapp, 1997) that bears out this insight that the profound deficiencies the classical conception of nature are not confinable to the micro-level. This key result will be discussed in section 4. But first I discuss the reason why, as Mario Bunge said: ``The physicist of the latest generation is operationalist all right, but usually he does not know, and refuses to believe, that the original Copenhagen interpretation --- which he thinks he supports --- was squarely subjectivist, i.e., nonphysical.'' \vskip .1in \noindent{\bf 3. The Vulgar Copenhagen Interpretation.} Let me call the original subjectivist, knowledge-based Copenhagen interpretation the ``strict'' Copenhagen interpretation. It is pragmatic in the sense that it is a practical viewpoint based on human experience, including sensations, thoughts, and ideas. These encompass both the empirical foundation of our physical theories and the carrier of these theories, and perhaps all that really matters to us, since anything that will never influence any human experience is, at least from an anthropocentric viewpoint, of no value to us, and of uncertain realness. Nevertheless, the prejudice of many physicists, including Einstein, is that the proper task of scientists is to try to construct a rational theory of nature that is not centered on such a small part of the natural world as human experience. The stalwarts of the Copenhagen interpretation were not unaware of the appeal of that idea to some of their colleagues, and they had to deal with it in some way. Thus one finds Bohr(1949) saying, in his contribution `Discussion with Einstein' to the Schilpp(1951) volume on Einstein: \noindent ``In particular, it must be realized that---besides in the account of the placing and timing on the instruments forming the experimental arrangement---all unambiguous use of space-time concepts in the description of atomic phenomena is confined to the recording of observations which refer to marks on a photographic plate or similar practically irreversible amplification effects like the building of a water drop around an ion in a cloud-chamber.'' \noindent and, \noindent ``On the lines of objective description, it is indeed more appropriate to use the word phenomenon to refer only to observations obtained under circumstances whose description includes an account of the whole experimental arrangement. In such terminology, the observational problem in quantum physics is deprived of any special intricacy and we are, moreover, directly reminded that every atomic phenomena is closed in the sense that its observation is based on registrations obtained by means of suitable amplification devices with irreversible functioning such as, for example, permanent marks on a photographic plate, caused by the penetration of electrons into the emulsion. In this connection, it is important to realize that the quantum mechanical formalism permits well-defined applications referring only to such closed phenomena.'' These are carefully crafted statements. If read carefully they do not contradict the basic thesis of the strict Copenhagen interpretation that the quantum formalism is about our observations described in plain language that allows us to ``tell others what we have done and what we have learned.'' On the other hand, it seems also to be admitting that there really are events occurring `out there', which are we are observing, but which do not derive their realness from our observations of them. Heisenberg (1958) says something quite similar: \noindent ``The observation, on the other hand, enforces the description in space and time but breaks the determined continuity of the probability function by changing our knowledge of the system.'' \noindent "Since through the observation our knowledge of the system has changed discontinuously, its mathematical representation also has undergone the quantum jump, and we speak of a `quantum jump' .'' \noindent ``A real difficulty in understanding the interpretation occurs when one asks the famous question: But what happens `really' in an atomic event?'' \noindent ``If we want to describe what happens in an atomic event, we have to realize that the word `happens' can apply only to the observation, not to the state of affairs between the two observations. It [ the word `happens' ] applies to the physical, not the psychical act of observation, and we may say that the transition from the `possible' to the `actual' takes place as soon as the interaction of the object with the measuring device, and therefore with the rest of the world, has come into play; it is not connected with the act of registration of the result in the mind of the observer. The discontinuous change in the probability function, however, occurs with the act of registration, because it is the discontinuous change in our knowledge in the instant of recognition that has its image in the discontinuous change in the probability function.'' All of this is very reasonable. But it draws a sharp distinction between the quantum formalism, which is about knowledge, and a world of real events that are actually occurring `out there', and that can be understood as transitions from the `possible' to the `actual', closed by irreversible processes when the interaction between the object and the measuring device, and hence the rest of the world, comes into play. Yet the extreme accuracy of detailed theoretical calculations [one part in a hundred million in one case] seems to make it clear that the mathematical formalism must be closely connected not merely to our knowledge but also to what is really happening `out there': it must be much more than a mere representation of our human knowledge and expectations. I call this natural idea---that the events in the formalism correspond closely to real ``physical'' events out there at the devices---the Vulgar Copenhagen Interpretation: vulgar in the sense of common and coarse. This vulgar interpretation is I think the common interpretation among practicing quantum physicists: at this symposium some important experimentalists were, as Mario Bunge suggested, unwilling to believe that the quantum mechanical formalism was about `our knowledge'. And it is coarse: the idea of what constitutes an `irreversible' process is not carefully specified, nor is the precise meaning of `as soon as the interaction with the object with the measuring device comes into play'. My aim in this paper is to reconcile the strict and vulgar interpretations: i.e., to reconcile the insight of the founders of quantum theory that the mathematical formalism of quantum is about knowledge with the demand of Einstein that our basic physical theory be a theory of nature herself. The main obstacle to a rational understanding of these matters is the faster-than-light action that the quantum formalism seems to entail, if interpreted at a physical level. If one takes literally the idea that the quantum event at the device constitutes a real transition from some physical state of `possibility' or `propensity' to a state of `actuality' then---in the `entangled states' of the kind studied by Schroedinger, by Einstein, Podolsky, and Rosen, and by Bell and others---it would seem that the mere act of making a measurement in one region would, in certain cases, instantly produce a change in the physical propensities in some far-away region. This apparent faster-than-light effect is dealt with in the strict Copenhagen interpretation by denying that the probability function in the formalism represents anything physical: the formalism is asserted to represent only our knowledge, and our knowledge of far-away situations can be instantly changed---in systems with correlations---by merely acquiring information locally. This fact that the strict Copenhagen interpretation ``explains away'' the apparent violations of the prohibition [suggested by the theory of relativity] of faster-than-light actions is a main prop of that interpretation. So the essential first question in any attempt to describe nature herself is the logical status of the claimed incompatibility of quantum theory with the idea---from the theory of relativity in classical physics---that no influence can act backward in time in any frame of reference. It is of utmost importance to progress in this field that we get this matter straight. \vskip .1in \noindent {\bf 4. Causality, Locality, and Ontology.} David Hume cast the notion of causality into disrepute. However, when one is considering the character of a putative law of evolution of a physical system it is possible to formulate in a mathematically clean way a concept of causality that is important in contemporary physical theory. In relativistic physics, both classical and quantum mechanical, the idea of causality is introduced in the following way: We begin with some putative law of evolution of a physical system. This law is specified by picking a certain function called the Lagragian. A key feature of the possible Lagrangians is that one can modify them by adding a term that corresponds to putting in an extra-force that acts only in some small spacetime region R. The evolution is specified by the ``law'' specified by the chosen Lagrangian, plus boundary conditions. Let us suppose that boundary condition is specified as the complete description of ``everything'' before some ``initial time'' $T_{in}$. The laws then determine, in principle, ``everything'' for all later times. In classical mechanics ``everything'' means the values of all of the physical variables that are supposed to describe the physical system that is being considered, which might be the entire physical universe. In quantum mechanics ``everything'' means all of the ``expectation values'' of all of the conceivable possible physical observables, where ``expectation value'' means a predicted average value over an (in principle) infinite set of instances. To bring in the notion of causality one proceeds as follows. It is possible, both in classical and quantum theory, to imagine changing incrementally the Lagrangian that specifies the law of evolution. The change might correspond to adding extra terms to the forces acting on certain kinds of particles if they are in some small spacetime region R that lies later than time $T_{in}$. Such a change might be regarded as being introduced whimsically by some outside agent. But, in any case, one can compare the values of ``everything'' at times later than time $T_{in}$ in the new modified world (i.e., the world controlled by the new modified Lagrangian) to the values generated from the laws specified by the original Lagrangian. If one is dealing with an idealized world without gravity, or at least without any distortion of the `flat' Minkowsky spacetime, then it is a mathematical property of relativistic field theories, both classical and quantum mechanical, that ``nothing'' will be changed outside the forward light cone of the region R in which the Lagrangian was changed! In other word, ``everything'' will be exactly the same in the two cases at all points that cannot be reached from the spacetime region R without moving faster than the speed of light. This property of relativistic field theories is called a causality property. The intuition is that this change in the Lagrangian can be regarded, or identified, as a ``cause'', because it can be imposed whimsically from outside the physical system. The mathematical property just described says that the effects of this ``cause'' are confined to its forward light cone; i.e., to spacetime points that can be reached from the spacetime region R of the cause without ever traveling at a speed greater than the speed of light. Relativistic field theories are formulated mathematically in such a way that this causality property holds. This means that insofar as it is legitimate to imagine that human beings can ``freely choose'' [i.e., can act or not act upon a physical system without there being any cause {\it from within that physical system} of this act] to do one thing or another in a region $R$ [e.g., to exert or not exert a force on some physical particles of the system in region $R$] then ``everything'' outside the forward light cone of $R$ will be independent of this choice: i.e., ``everything'' outside this forward light cone will be left unaltered by any change in this choice. This relativistic causality property is a key feature of relativistic field theories in flat Minkowsky spacetime: it is all the causality that the orthodox pragmatic quantum philosophy calls for. But notice that by ``everything'' one means, in the quantum case, merely the ``expectation values'', which are averages over an (in principle) infinite ensemble of instances. Now, one might think that since this relativistic causality property holds for these averages it ought to be {\it at least conceivably possible} that it could hold also for the individual instances. But the amazing thing is that this is not true! It is not logically possible to impose the no-faster-than-light condition in the individual instances, and maintain also the validity of certain simple predictions of quantum theory. The point is this. Suppose one considers an experimental situation involving two experimental regions that are spacelike separated from each other. This means that no point in either region can be reached from any point in the other without traveling faster than the speed of light. In the first region there is an experimenter who can freely choose to do one experiment or another. Each of these two alternative possible experiments has two alternative possible outcomes. There is a similar set up in the second region. Each possible outcome is confined to the associated experimental region, so that no outcome of an experiment in one region should be able to be influenced by the free choice made by the experimenter in the other region. One single instance is considered, but with the two free choices of the two experimenters being treated as two free variables. Thus the one single instance under consideration, uniquely fixed at all times earlier than the earliest time in either of the two experimental regions, will go into one or another of altogether $(2\times 2)=$ four alternative possible evolutions of this system, depending on which of the two alternative possible choices is made by each of the two experimenters. There can then be further branchings that are specified by which of the possible outcomes nature selects for whichever experiments are performed. The particular experimental details can be arranged so that the assumed validity of the predictions of quantum theory for that particular arrangement entails the {\it nonvalidity} of at least one of the three following locality conditions: {\bf LOC1:} It is possible to impose the following condition: If in each of the two regions the first of the two possible experiments were to be performed, and a certain result $r$ appeared in the first region then if this very same experiment were to be performed in the first region then this same result $r$ would appear there even if the experimenter in the second region were to elect at the last moment to do the other measurement. The rationale for this locality condition is that a free choice of what to do in one place cannot---relativity theory leads us to believe--- affect, at a speed faster than the speed of light, what occurs elsewhere: making a different choice in one region should not be able to force what appears (at the macroscopic, observable level) in the other region to be different. Indeed, in some frame of reference the outcome in the first region has already occurred before the experimenter in the second region makes his free choice of which experiment he will perform. But, according to ideas from relativity theory, what someone has already seen and recorded here at some earlier time cannot be disturbed by what a faraway experimenter freely chooses to do at some later time. Notice that LOC1 requires only that it be possible to impose this condition. The point is that only one of the two possible experiments can actually be performed in the second region, and hence nature herself will make only one choice. So what would actually appear in the first region if the experimenter in the other (far away) region were (at some future time) to make a different choice in not physically well defined. Thus this is a theoretical investigation: the question is whether the predictions of QT are compatible with the notion that nature evolves in such a way that what one observer sees and records in the past can be imagined to be fixed independently of what another person will freely choose to do in the future. {\bf LOC2:} Suppose, under the condition that the first of the two possible measurements were to be performed in the first region (with no condition imposed on what the outcome there is) that one can prove from LOC1 and the predictions of quantum theory, the truth of a statement $S$ that pertains exclusively to what experimenters can observe under various possible conditions of their own making in the second region. Then this locality condition asserts that it is logically possible to demand that $S$ remain true under the condition that the experimenter in the first region freely chooses (say in the future) to perform there, instead, the second possible measurement. The rationale is that, according to certain ideas from the theory of relativity, the truth of a statement that pertains to macroscopic conditions that refer exclusively to one space-time region should not depend on what someone far away freely chooses to do later. {\bf LOC3} This is another form of LOC1: Altering the free choice in R leaves any outcome in L undisturbed. [See Stapp, 1997] The validity of the predictions of quantum theory in correlation situations like this are being regularly borne out. (...Most recently in a highly publicized experiment using the Swiss telephone company optical fibers to connect experimental regions that were 14 km apart, with the intent of making important practical applications.) Thus it can, I believe, be confidently assumed that the pertinent quantum predictions are valid. But in that case one of the ``locality conditions'' described above {\it must} fail. Before drawing any conclusions one must consider the impact or significance of the assumption that the experimenters' choices can be treated as ``free variables''. It is part of the orthodox quantum philosophy that the experimenters' choices can and should be considered to stand outside the physical system that is being examined. Bohr and Heisenberg argued that biological systems in general lie outside the domain covered by the pragmatic framework. But in any case, one thing is certain: the beautiful and elegant quantum formalism is naturally suited to the idea that it represents a system that is part of a bigger system that can extract information from it, where the nature of the information being extracted from the subsystem is controlled by things outside that subsystem, namely the observer and his instruments of observation. But even at a more intuititive level it seems that the decision-making process of human experimenters are so complex and delicate, and so insulate-able in principle, prior to the time of the examination, from the system that they are about to examine, as to render their choices as to what to look for effectively free, under appropriate conditions of isolation, from any influence upon them {\it by the system they are about to examine}. So it would seem to be safe, under appropriate conditions of prior isolation, to treat these choices {\it as if they were free from such influences} even in a strictly deterministic universe. In a quantum universe this move is even more reasonable, because the choices could be governed by a quantum process, such as the decay of a radio-active nucleus. Within the quantum theoretical framework each such decay appears as a spontaneous random event. It is free of any ``physical'' cause, where ``physical'' means something that is part of the physical world as that world is described by the physical theory. Thus within both the deterministic and stochastic contexts it seems reasonable to treat the choices to be made by the experimenters as if they were free, in the sense of not being influenced by the physical properties of the system that is about to be examined. One caveat. The arguments require that meaning be given to a condition such as: ``If the experimenter in region one performs experiment one, and the outcome that occurs there is outcome one''. This condition is nonsensical in the Everett many-minds interpretation, because {\it every} outcome occurs. I have excluded that interpretation from consideration on other grounds, which are described in section 5. The apparent failure of the locality condition has three important consequences: 1. It gives a solid basis for the conclusion of the founders of quantum theory that no return to the notions of classical mechanics (relativistic field theory) is possible: the invalid locality property certainly holds in relativistic classical mechanics. 2. It makes reasonable the attempt to ontologicalize the orthodox interpretation. It had formerly been believed that this was a nonsensical thing to try, because ontologicalization immediately entails faster-than-light transfer of information on the individual-instance level. Such transfers had seemed unacceptable, but are now seen to be unavoidable even in a very general framework that maintains merely the validity of the {\it predictions} of quantum theory, and the idea that the experimenters' choices can be considered to be ``free'', in the weak sense discussed above. 3. Because the nonlocal effects enter into orthodox quantum theory specifically in connection with the entry of our knowings into the dynamics there is prima facie evidence that our knowings may be associated with the nonlocal aspect of nature. It is worth noting that these effects are not confined to a microscopic scale: in the Swiss experiment the effect in question extended over a separation of 14km. And, according to quantum theory, the effect does not fall off at all with distance. In my proposal each of our knowings is associated with a brain event that involves, as a unit, a pattern of brain (e.g., neuronal) activity that may extend over a large part of the brain. The collapse actualizes this whole pattern, and the associated knowing is an expression of the functional properties of this pattern. Once the reality is recognized to be knowledge, rather than substantive matter, the nonlocal connections seem less problemmatic: nothing but knowledge about far-away knowings is changed by nearby knowings. \vskip .1in \noindent {\bf 5. All Roads Lead to Solvay 1927.} The Solvay conference of 1927 marks the birth of (coherently formulated) quantum theory. Two of the many important papers delivered there stand out. Born and Heisenberg presented a paper on the mathematical formalism and proclaimed that the essential features of the formalism were complete and not subject to further revision. Dirac gave a paper on the interpretation, and claimed that ``the wave function represents our knowledge of the system, and the reduced wave packets our more precise knowledge after measurement.'' These two parts, the mathematical formalism and its interpretation in terms of knowledge, meshed perfectly: that was the logical basis of the Copenhagen interpretation. This was an epic event in the history of human thought. Since the time of the ancient Greeks the central problem in understanding the nature of reality, and our role in it, had been the puzzling separation of nature into two seemingly very different parts, mind and matter. This had led to the divergent approaches of idealism and materialism. According to the precepts of idealism our ideas, thought, sensations, and other experiential realities should be taken as basic. But then the mathematical structure carried by matter was difficult to fathom in any natural way. Materialism, on the other hand, claimed that matter was basic. But, if one started with matter then it was difficult to understand how something like your experience of the redness of a red apple could be constructed out of it, or why the experiential aspect of reality should exist at all if, as classical mechanics avers, the material aspect is dynamically complete by itself. There seemed to be no rationally coherent way to comprehend the relationship between our experiences of the reality that exists outside our thoughts, and the nonexperiential-type material substance that the external reality was claimed to be made of. Yet at the Solvay meeting, physicists, of all people, had come up with a perfect blending, based on empirical evidence, in which the mathematical structure needed to account for all of the empirical regularities formerly ascribed to substantive matter, was present without there being anything like substantive matter: the mathematical structure was a property of knowings! What an exhilerating moment it must have been. Driven simply by the need to understand in a rational way the empirical facts that nature had presented to us, scientists had been led to a marvelous resolution of this most fundamental of all philosophical problems. It was a tremendous achievement. Now, seventy years later, we are able to gather here at the X-th Max Born Symposium to celebrate the unbroken record of successes of that profound discovery, and to hear about its important new triumphs. So now, the end of our Symposium, I take this opportunity to review briefly some of its highlights from the perspective of the Solvay breakthough. Probably the most exciting reports were from experimentalists who are now performing experiments that could only be imagined seventy years ago. Yet the thinking of the founders of quantum theory did involve ``gedanken'' experiments designed to confirm the rational coherency of the framework. Today these ``thought'' experiments involving preparations and measurements on small numbers of individual atoms are being carried out, and the results invariably confirm all of the ``quantum weirdness'' that the Copenhagen interpretation predicted. But do these successes really confirm the radical ideas of Solvay 1927? Time has eroded the message of Solvay to the extent that the scientist performing the experiments hardly recognize the Solvay insights in the interpretation of their work, though they give lip service to it. One must probe into the rational foundations of the subject to see the import of their results on this deep question. I cite first the report of Omnes. There had been hope that some way around the Copenhagen interpretation would emerge from the studies of decoherence and consistent histories that have been so vigorously pursued of late. No one has pursued these ideas more deeply than Omnes. His verdict is that these methods amount to ``the Copenhagen interpretation `done right' ''. He said similar things in his book (Omnes, 1994). And such prominent proponents of ``decoherence'' as Zurek(1986) and Joos(1986) have said similar things: Zurek concluded that the study of decoherence ``constitutes a useful addition to the Copenhagen ...a clue pointing at a still more satisfactory resolution of the measurement problem...a hint about how to proceed rather than the means to settle the matter quickly.'' Joos asks at the beginning of his article ``Is there some way, at least a hint, how to understand... '' and at the end says ``one may hope that these superselection rules can be helpful in developing new ideas ..[about].. measurement processes.'' So they both stressed that decoherence effects do not resolve the deep problems. Indeed, decoherence is rather the {\it cause} of the problem: decoherence effects make it virtually impossible to empirically determine whether quantum collapses are occurring outside our brains or not. It is precisely {\it because} of decoherence effects that we cannot tell, empirically, whether or not collapses actually do occur ``when the interaction of the object with the measuring device, and hence the rest of the world, comes into play''. The decoherence-consistent-histories approach had originally been pursued within the Everett framework, and indeed was sometimes called the `post-Everett' approach to stress that it was being pursued within that framework, rather than the Copenhagen framework, which it sought to unseat. But Omnes put his finger on the fatal flaw in the Everett approach when he said that it did not explain the transition from ``and'' to ``or''. In the evolving wave function of Everett the various branches do evolve independently, and hence might naturally be imagined to have different ``minds'' associated with them, as Everett suggests. But these branches, and the minds that are imagined to be properties of these branches, are all simultaneously present. Hence there is no way to give meaning to the notion that one mind is far more likely to be present at some finite time than the others. It is like waves on a pond: the big waves and the small ones are all present simultaneously. So one needs something else, perhaps like a surfer that will be pushed into one branch or the other, to define the ``or'' that is logically needed to define the notion of the probabilities of the different ``alternatives''. Yet the Everett interpretation allows nothing else besides the wave function and its properties. So all the minds are simultaneously present because all the corresponding properties of the various branches are simultaneously present. The idea of the surfer being pushed by the wave is exactly the idea behind the model of David Bohm that was so ably expounded here by D. Duerr and by F. Faisal. But the model has not been consistently extended to the relativistic case of quantum electrodynamics, or to quantum chromodynamics, which are our premiere quantum theories. The model has other unpleasant features. One is the problem of the empty branches. Each time a ``good measurement'' is performed the wave function must separate into different ``branches''. These branches are parts of the wave function such that the full wave function is a sum (i.e., superposition) of these branches, and each branch is nonzero only in a region (of the 3n-dimensional space in which these wave functions live) that overlaps none of the other regions. Here n is the number of particles in the universe. If two branches separate then the `surfer' (which in the Bohm model would be the entire classically described physical world) must end up in just one of these branches. But all the other branches (which are regarded as physically real) must continue to evolve for all eternity without ever having any effect upon the `surfer', which is the only part of reality that is directly connected to human experience, according to the model. This seems wildly extravagant! If the surfer is the important thing then the effort of nature to continue to evolve these ineffectual branches for all eternity seems to be a gigantic waste of effort. But if the surfer is not important then why is this tiny part of reality there at all? It does nothing but get pushed around. There is a perhaps bigger problem with the initial conditions. The model is predicated on the premise that the single real classical world is a random element in a statistical ensemble of possibilities. The idea of a statistical ensemble makes good sense when we have the possibility of repeated preparations of similar situations. But when we are speaking about the entire universe it does not seem to make sense to speak of a particular statistical ensemble of universes with some particular density (weight) function if only one of them is ever created. Or are we supposed to think that a whole ensemble of real classical worlds is created, and that ``our'' real world is just one of them? That would seem to be the more natural interpretation. But I asked David Bohm about that, many years ago, and he insisted that there was, according to his thinking, only one universe. Bohm was stimulated to construct his model by conversations with Einstein. Yet Einstein rejected the model, calling it ``too cheap''. I asked Bohm what he thought about Einstein's evaluation, and he said he completely agreed. Indeed, at the end of his book with Hiley about his model, after finishing the part describing the model, he added two chapters about going beyond the model. He motivated those chapters by references to the efforts that I was making, and that Gell-mann and Hartle were making, to go beyond the Copenhagen interpretation. Gell-mann and Hartle were pursuing the decoherence-consistent-histories approach mentioned above, which has led back to Solvay, and I had proposed a theory of events. The events were real collapses of a wave function that was considered to be ontologically real. This brings me to the talk of Rudolf Haag. Haag described his theory of events, and mentioned that it still needed twenty years of work. In his written account Haag(1996) mentions that I had proposed essentially the same theory in the seventies, some twenty years ago (Stapp, 1975, 1977, 1979). {\it My} twenty years of work on this idea has lead back to Solvay 1927. The problem is always the same: if one wants to make natural use of what nature has told us, namely that the beautiful mathematical formalism works to high precision, then one is led to ascribe to that formalism some ontological reality. But then the condition for the collapses must be spelled out in detail. It is natural for physicists to try to find {\it purely physical} conditions. But in the end there are no adequate natural conditions of this kind: the possibilities are all unnatural and ad hoc. Von Neumann said it all when he showed, back in the thirties, that one could push the boundary between the world described by the quantum formalism and the world described in terms our classical concepts all the way to the boundary between brain and mind without disrupting the predictions of quantum theory, and noted that there is no other natural place to put the boundary, without disrupting the integrity of the theory. In fact, it is, in principle, {\it only} if one pushes the boundary all way to the brain-mind interface that one obtains, strictly, the prediction of orthodox quantum theory: otherwise there are rogue collapses that are not associated with knowings. Of course, pushing the boundary all the way to mind brings mind into our theory of nature. But why on earth should we try to keep mind out--- bottled up, ignored, and isolated from the physical world---when we know it is present, and seemingly efficacious, particularly when the intense struggle of physicists to find a rational way of accounting for the observed phenomena led them to the conclusion that the theory of physical reality has the form of a theory about knowings, not the form of a theory about matter. Our aim should be not to bring back moribund matter, which we are well rid of, but to learn how better to understand knowings, within the mathmatical framework provided for them by the quantum formalism. \vskip .1in \noindent {\bf 6. The Two Quantum Processes.} There have been many attempts by physicists to `get mind back out of physics': i.e., to reverse the contamination of physics brought in by Bohr, Heisenberg, Dirac, Pauli and company in 1927. I believe those decontamination efforts have failed, even though I myself have worked hard to achieve it. So I am taking here the other tack, and trying to build a coherent ontology around the orthodox ideas. In particular, I am accepting as basic the idea that there are knowings, and that each such knowing occurs in conjunction with a collapse of the wave function that reduces it to a form concordant with that knowing. I assume that knowings are not associated exclusively with human body/brains. But I shall focus here on these particular kinds of knowings because these are the ones we know most about. A fundamental fact of orthodox quantum theory is that the evolution of the state of the physical system between the collapse events is mathematically very different from the evolution of this state associated with the collapses: the former are ``unitary'' and ``local'', whereas the latter are neither. The ``unitarity'' property means several things. On the one hand, it means that the evolution is in some sense no change at all: the internal or intrinsic structure of the state is unaltered. One can imagine that only the `mode of description' of the state is changed, not the state itself. Indeed, that point of view is very often adopted in quantum theory, and is the one I shall adopt here. (See the next section.) The ``unitarity'' property also means that the transformation operator that changes the state at an earlier time to the state at a later time does not depend on that initial (or final) state: there is, in this sense, in connection with the unitary part of the process of evolution, {\it no self reference!} According to the orthodox interpretation, there is no experiential reality associated with the unitary part of the evolution, which is the part between the observations: there is no essential change, and no self reference, and hence, reasonably enough, no experience. Experiences are associated only with the nonunitary parts of the evolution: the part associated with observations. For that part there {\it is} essential change, and the transformation operator analogous to the one defined for the unitary case would depend on the state upon which it acts. Thus there would be, in this sense, self-reference. This self reference (nonlinearity) plays a key role in the dynamics associated with observation. It is a special kind of self reference that has no counterpart in classical mechanics. In the classical approximation to the quantum dynamics only the unitary part of the dynamical evolution survives. So from a quantum mechanical point of view it would be nonsensical to look for mind in a system described by classical physics. For classical physics is the result of an approximation to the full dynamical process of nature that eliminates the part of that process that orthodox quantum theory says is associated with our experiences. \vskip .1in \noindent {\bf 7. The Two Times: Process Time and Mathematical Time} The distinctions between the two processes described above is central to this work. It can be clarified, and made more vivid, by explaining how these two processes can be considered to take place in two different times. In quantum theory there are two very different kinds of mathematical objects: vectors and operators. Operators operate on vectors: the action of an operator on a vector changes it to another (generally different) vector. Given an operator, and a vector that represents a state of a physical system (perhaps the entire universe), a number is formed by first letting the operator act on the vector, and then multipling the resulting vector by the (complex conjugate of the) original vector. This number is called the ``expectation value of the operator in the state represented by that vector''. Modern field theories are generally expressed in the so-called Heisenberg picture (rather than the so-called Schroedinger picture). I shall follow that practice. In ordinary relativistic quantum field theory each spacetime point has a collection of associated operators. (I gloss over some technicalities that are not important in the present context.) Consider the collection of operators $C(t)$ formed by taking all of the operator associated with all of the spacetime points that lie at fixed time $t$. This set $C(t)$ is ``complete'' in the sense that the expectation values of all the operators of $C(t)$ in a state $S$ determine all the expectation values of the all the operators in $C(t')$ in the state $S$, for every time $t'$. The operators in $C(t)$ are related to those in $C(t')$ by a {\it unitary} transformation. Whether one represents the state $S$ by giving the expectation values in this state of all the operators in $C(t)$, or of all the operators in $C(t')$, is very much like choosing to use one coordinate system or another to describe a given situation: it is just a matter of viewpoint. The unitary transformation that relates the collection of operators $C(t)$ to the collection of operators $C(t')$ is essentially the unitary transformation associated with the Schroedinger-directed temporal evolution. It is in this sense that the unitary transformation that generates evolution in the ``mathematical time'' $t$ is relatively trivial. It is deterministic, continuous, invertible, and independent of the state $S$ of the physical system upon which the operators act. But giving the complete set of all the operators associated with all the points in spacetime says nothing at all about the evolution of the state! Saying everything that can be said about the operators themselves, and about evolution via the unitary part of the transformation has merely fixed the mode of description, and the connections between different modes of description. It has not said anything about the all-important evolution of the state. The state undergoes a sequence of abrupt jumps: $$...S_i \longrightarrow S_{i+1} \longrightarrow S_{i+2}... .$$ The situation can be displayed graphically by imagining that $i$ is the imaginary part of the complex time $t$: the evolution proceeds at constant imaginary part of $t$ equal $i$, and at constant $S_i$, with the real part of $t$ increasing until it reaches a certain `jump time' $t_i$, whereupon there is an abrupt quantum jump to a new constant state $S_{i+1}$, and a new constant imaginary part of $t$ equal to $i+1$, and the evolution then again proceeds with increasing real part of $t$ until the next `jump value' $t_{i+1}$ is reached, and then there is another jump up to a new value, $i+2$, of the imaginary part of $t$. Thus the full process is represented in complex time as a line having the shape of a flight of steps. The horizontal segments where the real part of time is increasing represent the trivial unitary parts of the process, which correspond merely to changing the viewpoint, or mode of description, with the state remaining fixed, and with no associated experience. The vertical segments correspond to increases in `process time'. These are the parts associated with experience. (This identification of the vertical axis with imaginary time is purely pedagogical) The present endeavour is to begin to fill in the details of the process associated with the increases in the vertical coordinate, process time, which is the time associated with the nontrivial part of the evolutionary process, and with experience. The final phase of each vertical segment is the fixing of a new knowing. But some process in Nature must bring about this particular fixing: this process is represented by motion along the associated vertical segment. \vskip .1in \noindent {\bf 8. Quantum Ontology} What is the connection between the our experiences and the physicists' theoretical description of the physical world? The materialist position is that each experience {\it is} some aspect of the matter from which the physicists say the world is built. But the physical world certainly is not built out of the substantive matter that was postulate to exist by classical mechanics. Such stuff simply does not exist, hence our experiences cannot be built out of it. The quantum analog of physical reality, namely the quantum state S of the universe, is more like information and ideas than like the matter of classical physics: it consist of accumulated knowledge. It changes when human knowledge changes, and is tied to intentionality, as I shall explain presently. Orthodox classical mechanics is naturally complete in itself: the physical world represented in it is dynamically complete, and there is no hint within its structure of the existence of anything else. Orthodox quantum mechanics is just the opposite: the physical world represented by it is not dynamically complete. There is a manifest need for a process that is not represented within the orthodox description. In orthodox quantum mechanics the basic realities are our knowings. The dynamics of the physical world represented in the orthodox quantum formalism is not internally complete because there is, in connection with each knowing, a collapse process that appears in the orthodox theory as a ``random choice'' between alternative possibilities: contemporary quantum theory provides no description of the process that selects the particular knowing that actually occurs. This collapse process, which is implemented by a nonunitary/nonlocal transformation, must specify two things that the contemporary machinery of quantum theory does specify: \noindent 1. It must specify an experience E, associated with a corresponding projection operator P(E), such that the question is put to Nature: ``Does E occur?'' \noindent 2. It must then select either the answer `yes', and accordingly change the current state (i.e., density matrix) S to the state PSP, or select the answer `no', and accordingly replace S by (1-P)S(1-P). The probability of answering `yes' is Trace PSP/TraceS; the probability of answering `no' is Trace (1-P)S(1-P)/Trace S. In the orthodox pragmatic interpretation the step 1 is achieved by a human experimenter's putting in place a device whose observed response will determine whether the system that is being examined has a certain property specified by P(E): the occurrence of experience E will confirm, basically on the basis of past experience, that future experiences will be likely to conform to the answer ``Yes, the system has property P(E).'' According to the orthodox viewpoint, the experimenter stands outside the quantum system being examined, and the device is regarded as an extension of himself. Step 2 is then achieved by appeal to a random selection process that picks the answer `Yes' or `No' in accordance with a statistical rule. This selection process (also) is not represented within the orthodox Hilbert space description. How can these two steps be comprehended in a rational, minimalistic, naturalistic way? \vskip .1in \noindent {\bf 9. Von Neumann's Process I.} The first step in the nonunitary process is what von Neumann called Process I, in contrast to his Process II, which is the normal unitary evolution. Process I consists of ``posing the next question". We can suppose that the possible answers are $Yes$ or $No$. Nature will then answer the question. The crucial requirement is that the answer $Yes$ must be recognizably different from the answer $No$, which includes no recognizable answer at all. In practice a human being creates the conditions for Process I, and it is he who recognizes the positive response: this recognition is a knowing. For example, the observer may know that he is seeing the pointer on the device---that he himself has set in plac---resting definitely between the numbers 6 and 7 on the dial. This is a complex thing that he knows. But knowings can be known, at least in part, by later knowings. This is the sort of knowing that science is built upon. Of course, all one can really know is that one's experiences are of a certain kind, not that there really is a pointer out there. So we expect the knowings to correspond in some way to a brain activity of some sort, which under normal circumstances would be an effect of something going on outside the brain. Von Neumann accepts the statistical character of the theory, and his Process I is statistical in character: his Process I covers merely the posing of the question, and the assignment of a statistical weight to each of the recognizably different alternative possible answers. It does not cover the subsequent process whereby Nature delivers an answer. My basic commitment here is to accept the quantum principles as they are, rather than to invent new principles that would allow us to exclude mind from Nature's dynamics. So I accept here, ontologically as well as pragmatically, that {\it the possibilities singled out in Process I are defined by different `possible knowings'}. Two important features of the von Neumann Process I are: 1) It produces an abrupt increase in entropy. If the state of the universe prior to the process is well defined, so that the entropy (with no coarse graining) is zero, then if, for example, the Process I gives a statistical mixture with $50\%\ Yes$ and $50\%\ No$, the entropy will jump to $ln 2$. 2) It is quasi-local. There will be nonlocal aspects extending over the size of the examined system, but no long-range nonlocal effects of the kind mentioned in section 3. That is, there will be, for the Process I associated with a human knowings, brain-sized nonlocal effects associated with defining the question, but no nonlocal effects extending outside the body/brain. Thus Process I is, for human knowings, a human process, not a global one. [Technically, the reason that there is no effect on far-away systems is that such an effect is computed by performing a `trace' over the degrees of freedom of the nearby system (e.g., the brain/body), but von Neumann's Process I is achieved by dropping out interference terms between the alternative possible answers, and that operation leaves this trace unaltered.] Process I lies at the root of measurement and mind-body problems. In approaches that try to explain Process I in purely physical terms, with knowings not mentioned, but rather forced to follow from physically characterized processes, the answers tend to assert either that:\\ 1), the wave function of a particle occasionally just spontaneously reduces to a wave function that is essentially zero except over a small region, or that\\ 2), what is not measurable in practice (i.e., via some practicable procedure) does not exist in principle: if it is impractical to detect an interference term them it does not exist. This latter sort of rule is certainly justified in a pragmatic approach. But most physicists have been reluctant to accept such rules at the ontological level. Hence the pragmatic approach has won by default. From the present standpoint, however, the basic principle is that Nature responds only to questions that are first posed, and whose answers are possible knowings, or are things of the same general ontological type as possible knowings. [The needed generalization will be discussed later, after the knowings themselves have been discussed.] But the important immediate point is that the quantum dynamics is organized so as to put knowings, and their possible generalizations, into the central position. All such knowings contribute to the general self knowledge of the universe, which is represented by the (Hilbert-space) state $S$ of the universe. \vskip .1in \noindent {\bf 10. Origin of the Statisical Rules} \vskip .1in Without loss of generality we can suppose that each posed question is a single question answered with a single reply, {\it Yes} or {\it No}. Then the usual (density matrix) formalism allows the reduction process to be formalized in the following way. The state of the universe is represented by the density matrix (operator) $S$. The question is represented by the projection operator $P$: $P^2 = P$. Then the von Neumann Process I is represented by $$ S\equiv [PSP+(1-P)S(1-P)+PS(1-P)+(1-P)SP] \longrightarrow PSP + (1-P)S(1-P). $$ The subsequent completion of the reduction is then represented by $$ [PSP + (1-P)S(1-P)] \longrightarrow PSP \mbox{ or }(1-P)S(1-P) $$ where the fractions of the instances giving the two results are: $$ (Tr PSP)/(Tr PSP + Tr (1-P)S(1-P)) \mbox{\ for\ } PSP $$ \noindent and $$ (Tr (1-P)S(1-P))/(Tr PSP + Tr(1-P)S(1-P)) \mbox{\ for\ } (1-P)S(1-P). $$ Here Tr represents the trace operation, which instructs one to sum up the diagonal elements $$ of the matrix $$ that represents the operator, for some complete orthonormal set of states $|i>$. [The value of the trace does not depend upon which complete orthonormal set is used, and, for any two (bounded) operators $A$ and $B$, $Tr AB = Tr BA$. Using this property, and $P^2 = P$, one sees that the denominator in the two equations just given reduces to $Tr S$. A partial trace is given by the same formula, but with the vectors $|i>$ now forming a complete orthonormal basis for {\ part} of the full system] I believe it is perfectly acceptable to introduce an unexplained random choice or selection in a pragmatically formulated theory. But in a rational ontological approach there must be some sufficient cause or reason for a selection to pick out $Yes$ rather than $No$, or vice versa. In view of the manifestly nonlocal character of the reduction process, there is, however, no reason for this selection to be determined locally. Quantum theory does not specify what this selection process is, and I do not try to do so. But given our ignorance of what this process is, it is highly plausible that it should give statistical results in accord with the rules specified above. The reason is this. If the selection process depends in some unknown way on things outside the system being examined then the fractions ought to be invariant under a huge class of unitary transformations $U$ of the state $S$ that leave $P$ invariant, for these transformations are essentially the trivial rearrangements of the distant features of the universe: $$ S\longrightarrow USU^{-1} \ \ \ \ \ \ \ U^{-1}PU = P. $$ Since the statistical description after the Process I has occurred is essentially similar to the classical statistical description one should expect $S$ and $P$ (or $(1-P)$) to enter linearly. But the trace formulas are the only possibilities that satisfy these conditions, for all $U$ that leave $P$ invariant. The point here is only that if the actual selection process depends in a complicated and unknown way on distant uncontrolled properties of $S$ then the long-term averages should not be sensitive to basically trivial rearrangements made far away. This assumption is quite analogous to the assumption made in classical statistical analysis---which has a deterministic underpinning---that in the absence of information about the full details one should integrate over phase space {\it without any weighting factor other than precisely one} in those degrees of freedom about which one has no information. Thus the quantum statistical rules need not be regarded as some mysterious property of nature to have unanalysable {\it tendencies} to make sudden random jumps: it is rational to suppose, within an ontological setting, that there is a causal, though certainly nonlocal, underpinning to these choices, but that we do not yet know anything about it, and hence our ignorance must be expressed by the uniquely appropriate averaging over the degrees of freedom about which we have no knowledge. The effective randomness of Nature's answers does not render the our knowings nonefficacious. Our knowings can enter the dynamics in a strongly controlling way {\it through the choice of the questions}, even though the answers to these questions are effectively random. The formation of the questions, in Process I, is human based, even though the selection of the {\it answers} is presumably global. This will be discussed presently. The theory is naturalistic in that, although there are knowings, there are no soul-like experiencers: each human stream of consciousness belongs to a human body/brain, which provides the structure that links the experiences of that stream tightly together. \vskip .1in \noindent {\bf 11. Brains and Experiences.} \vskip .1in The dynamics of the theory is organized around the collection of operators P(E) that connect experiences E to their effects on the state S of the universe. I describe here my conception of this connection, and of the dynamical differences between the quantum version of this connection and its classical analog. Each experience is supposed to be one gestalt that, like a percept, ``comes totally or not at all'', in the words of Wm. James (1987. p. 1061). This experience is part of a sequence whose elements are, according to James, linked together in two ways: each consists of a fringe that changes only very slowly from one experience to the next, and a focal part that changes more rapidly. The fringe provides the stable contextual framework. It is the background that provides both the contextual setting, within which the foreground is set, and the experience of a persisting historical self that provides both the backdrop for the focal part and the carrier of longer term motivations. The focal part has a sequence of temporally displaced components that, like the rows of a marching band that are currently in front of the viewing stand, consists of some that are just coming into consciousness, some that are at the center, and some that are fading out. The occurrence together, in each instantaneous experience, of this sequence of temporal components is what allows comparisons to be made within a conscious experience. Judgments about courses of events can be parts of an experiences. The experiences are organized in the first instance, around experiences of the person's body in the context of his environment, and later also around abstractions from those primitive elements. These matters are discussed in more detail in chapter VI of my book (Stapp ,1993). Each experience normally has a feel that includes an experience of a prolongation of the current sequence of temporal components: this prolongation will normally be a prolongation that is, on the basis of past experience, likely to be imbedded in the ``current sequence of temporal components'' of some later experience in the linked sequence of experiences. Each experience E induces a change of the state of the universe S--$>$ PSP. This change will, I believe, for reasons I will describe presently, be a specification of the classical part (see below) of the electro-magnetic field within the brain of the person. This specification will fix the activities of the brain in such a way as to produce a coordinated activity that will generally produce, via a causal chain in the physical world (i.e., via the causal evolution specified by the Schroedinger or Heisenberg equations of motion) the potentialities for the next experience, $E'$. That causal chain may pass, via the motor cortex, to muscle action, to effects on the environment, to effects on sensors, to effects on the brain, and finally to a set of potentialities for various possible prolongations of the current sequence of temporal components. Then a selection must be made: one of the potential experiences will become actual. But this description glosses over an essential basic problem: How do the possible experiences E and the associations E--$>$ P(E) get characterized and created in the first place. There is an infinite continuum of projection operators P such that S--$>$ PSP would generate a new state. Why are some particular P's given favored status, and why are these favored P's associated with ``experiences''? This favored status is this: some one of these favored P's will be picked out from the continuum of possibilities, in conjunction with the next phase of the dynamical process. This next phase is the putting to Nature of the question: Does the current state S jump to PSP or not? To provide some basis for getting the universe going in a way that tends to produce stable or enduring structure, instead of mere chaotic random activity, I assume that a basic characteristic of the underlying dynamics is to select only projectors P that impose a certain repetitiveness on the dynamical structure. These qualities of repetitiveness are assumed to be fundamental qualities of the projectors. But each such quality is a characteristic that is more general in its nature than any particular realization of it. These general qualities I call ``feels'': they encompass all human experiences, but extend far beyond. Thus the basic assumption is that certain projectors P have ``feels'', but most do not, where a ``feel'' is a generalized version of a human experience. Each feel is characterized by a quality of repetitiveness, and the actualization of this feel entails the actualization of some particular realization of that quality or pattern of repetitiveness within the dynamical structure that constitutes the universe. This actualization is expressed by the transformation S--$>$ PSP where P = P(E), and E is the feel: it is the quality of the repetitiveness that is being actualized. This general tendency to produce repetitive spatio-temporal patterns carries over to human experience, and will, I believe, be greatly enhanced by natural selection within the biological sphere. Thus the selection, from among the proferred potential experiences, of the next $E'$, will be such as to favor a sequences E--$>$ P(E)--$>$ $E'$ such that $E'$ is either the same as E, or at least the same as E in some essential way. Thus experiences, and their more general ontological cousins, feels, are tied to the generation of self-reproducing structures. This generation of regenerating/reverberating stable structures underlies quantum dynamics, in the form of the creation by the dynamics of stable and quasi-stable particles, and extends beyond human beings, both to biological systems in general, and even to the overall organization of the universe, according to the ideas being developed here. As regards this repetitiveness, it is undoubtedly pertinent that classical mechanics is formulated basically in space-time, with lawfulness expressed essentially by a static or quasi-static quality of momentum-energy. But the essence of the transition to quantum theory is precisely that this static quality of momentum-energy is replaced by a repetitive quality, by a characteristic oscillatory behavior': quantum theory is basically about repetitive regeneration. In line with all this, I assume that the projection operators P act by specifying the (expectation values of the) quantum elecromagetic field. There are many reason for believing that this is the way nature operates: 1. The EM fields naturally integrate the effects of the motions of the billions of ions and electrons that are responsible for our neural processes. Thus examining the EM fields provide a natural way of examining the state of the brain, and selecting a state of the EM field of the brain provides a natural way of controlling the behavior of the brain. 2. The EM field has marvelous properties as regards connections to classical physics. The bulk of the low-energy EM state automatically organizes itself into a superposition of ``coherent states'', each of which is described by a classical electromagnetic field, and which enjoys many properties of this classical elecromagnetic field. These ``classical'' states are brought into the dynamical structure in a natural way: the condition that each actually realized state will correspond to essentially a single one of these classically describable coherent states is what is needed to deal effectively, in a physically realistic way, with the infra-red divergence problem in quantum electro-dynamics. [See Stapp (1983), and Kawai and Stapp (1995)] 3, These ``classical'' states (coherent states) of the quantum EM field are robust (not easily disrupted by the thermal and random noises in a warm wet brain): they are ideal for use in generating self-reproducing effects in a warm, wet, noisy enviroment. [See Stapp (1987), (1993, p.130), and Zurek (1993)] 4. These classical states are described by giving the ampitudes in each of the oscillatory modes of the field: spacetime structure arises from phase relationships among the different oscillatory modes. Although the theory being developed here maintains a close connection to classical physics, its logical and ontological structure is very different. In classical physics the dynamics is governed entirely by myopic local rules: i.e., by rules that specify the evolution of everything in the universe by making each local variable respond only to the physical variables in its immediate neighborhood. Human experiences are thus epiphenomenal in the sense that they do not need to be recognized as entities that play any dynamical role: the local microscopic description, and the local laws, are sufficient to specify completely the evolution of the state of physical universe. Experiential gestalts can regarded as mere effects of local dynamical causes, not as essential elements in the causal progession. But the most profound lesson about nature learned in the twentieth century is that the empirically revealed structure of natural phenomena cannot be comprehended in terms of any local dynamics: natural phenomena are strictly incompatible with the idea that the underlying dynamics is local. The second most profound lesson is that the known observed regularities of natural phenomena can be comprehended in terms of a mathematical model built on a structure that behaves like representations of knowledge, rather than representations of matter of the kind postulated to exist in classical mechanics: the carrier of the structure that accounts for the regularities in nature that were formerly explained by classical physical theory is, according to contempory theory, more idealike than matterlike, although it does exhibit a precise mathematical structure. The third essential lesson is that this new description, although complete in important practical or pragmatic ways, is, as an ontological description, incomplete: there is room for additional specifications, and indeed {\it an absolute need for additional specifications} if answers are to be given to questions about how our experiences come to be what they are. The presently known rules simply do not fix this aspect of the dynamics. The purpose of work is to make a first stab at filling this lacuna. One key point, here, is that brains are so highly interconnected that it will generally be only large macroscopic structures that have a good chance of initiating a causal sequence that will be self-reproductive. So each possible experience E should correspond to a P(E) that creates a macroscopic repetitiveness in the states of a brain. A second key point is that our knowings/experiences can be efficacious not only in the sense that they select, in each individual case, what actually happens in that case, but also in the statistical sense that the rules that determine which questions are put to Nature, can skew the statistical properties, even if the answers to the posed questions follow the quantum statistical rules exactly. I turn now to a discussion of this point and its important consequences. \vskip .1in \noindent {\bf 12. Measurements, Observations, and Experiences.} \vskip .1in A key question is whether, in a warm wet brain, collapses associated with knowings would have any effects that are different from what would be predicted by classical theory, or more precisely, by a Bohm-type theory. Bohm's theory yields all the predictions of quantum theory in a way that, like classical mechanics, makes consciousness epiphenomenal: the flow of consciousness is governed deterministically (but nonlocally) by a state of the universe that evolves, without regard to consciousness, in accordance with local deterministic equations of motion. Bohm's theory, like classical physics, tacitly assumes a connection between consciousness and brain activity, but the details of this connection are not specified. The aim of the present work is to specify this connection, starting from the premise that the quantum state of the universe is essentially a compendium of knowledge, of some general sort, which includes all human knowledge, as contrasted to something that is basically mechanical, and independent of human knowledge, like the quantum state in Bohmian mechanics. I distinguish a ``Heisenberg collapse'', S--$>$ PSP or S--$>$ (1-P)S(1-P), from a ``von Neumann collapse'' S--$>$[PSP + (1-P)S(1-P)]. The latter can be regarded as either a precursor to the former, or a representation of the {\it statistical} effect of the collapse: i.e., the effect if one averages, with the appropriate weighting, over the possible outcomes. This latter sort of averaging would be pertinent if one wanted to examine the observable consequences of assuming that a certain physical system is, or alternatively is not, the locus of collapses. This issue is a key question: Are there possible empirical distinctions between the behaviors of systems that are---or alternatively are not--- controlled by high-level collapses of the kind that this theory associates with consciousness. Can one empirically distinguish, on the basis of theoretical principles, whether collapses of this kind are occurring within some system that is purported to be conscious. This question is pertinent both to the issue of whether some computer that we have built could, according to this theory, be conscious, and also to the issue of whether our own behavior, as viewed from the outside, has aspects that reveal the presence of the sort of quantum collapses that this theory associates with consciousness. This question about differences in behaviour at the statistical level feeds also into the issue of whether being conscious has survival value. If behaviour has, on the average, no dependence on whether or not collapses occur in the system then the naturalistic idea that consciousness develops within biological systems due to the enhancement of survival rates that the associated collapses provide would become nonsense. Indeed, that idea is nonsense within classical physics, for exactly this reason: whether conscious thoughts occur in association with certain physical activities makes absolutely no difference to the microlocally determined physical behavior of the system. There are certain cases in which a von Neumann collapse, S--$>$ [PSP + (1-P)S(1-P)], would produce no observable effects on subsequent behavior. To understand these conditions let us examine the process of measurement/observation. If one separates the degrees of freedom of the universe into those of ``the system being measured/observed'', and those of the rest of the universe, and writes the state of the universe as $$ S = |\Psi >< \Psi | $$ with $$ |\Psi> = \sum_i \phi_i \chi_i, $$ where the $\phi_i$ are states of ``the system being measured/observed'', and the $\chi_i$ are states of the rest of the universe, then since we observers are parts of the rest of the universe it is reasonable to demand that if someone can have an experience E then there should be a basis of orthonormal states $\chi_i$ such that the corresponding projector P(E) is defined by $$ P(E) \phi_i = \phi_i $$ for all $i$, $$ P(E) \chi_i = \chi_i $$ for $i$ in $I(E)$, but $$ P(E) \chi_i = 0, $$ otherwise, where $I(E)$ is the set of indices $i$ that label those states $\chi_i$ that are compatible with experience E. A ``good measurement'' is defined to be an interaction between the system being measured and the rest of the universe such that the set of states $\phi_i$ defined above with $i$ in $I(E)$ span a proper subspace of the space corresponding to the measured system. In this case the knowledge that $i$ is in the set $I(E)$ ensures that the state of the measured system lies in the subspace spanned by the set of states $\phi_i$ with $i$ in $I(E)$. That is, experience E would provide knowledge about the measured system. Let P\_ be the projector that projects onto the subspace spanned by the set of states $\phi_i$ with $i$ in $I(E)$. Then a von Neumann collapse with P\_ in place of P would be identical to the von Neumann collapse S--$>$ [PSP + (1-P)S(1-P)]. But then the observer would be unable to determine whether a collapse associated with P\_ occurred in the system, unbeknownst to him, or whether, on the contrary, the definiteness of the observed outcome was brought about by the collapse associated with his own experience. This is essentially von Neumann's conclusion. But why should an actual collapse associated with the measured/observed system correspond in this special way to a subsequent experience of some human being? Why should an actually occurring P\_ be such as to ensure an equivalence between P\_ and a P(E)? Von Neumann's approach to the measurement problem suggests that such a connection would exist. In both the von Neumann and Copenhagen approaches the measuring device plays a central role. Different perceptually distinguishable locations of some ``pointer'' on the device are supposed to become correlated, during an interaction between the measured system and the measuring device, to different orthogonal subspaces of the Hilbert space of the measured system. This perceptual distinctness of the possible pointer positions means that there is a correlation between pointer {\it locations} and experiences. That connection must be explained by the theory of consciousness, which is what is being developed here. But why, ontologically, as opposed to epistemologically, should the projector P\_ in the space of the measured/observed system be to a state that is tied in this way to something outside self, namely the {\it location} of a pointer on a measuring device with which it might have briefly interacted at some earlier time. Von Neumann did not try to answer this question ontologically. If the real collapse were in the brain, and it corresponded to seeing the pointer at some one of the distinguishable locations, then from an epistemological point of view the effect of this collapse would be equivalent to applying P\_ to the state of the measured/observed system. If one works out from experiences and brains, in this way. one can formulate the collapses in terms of collapses out in the world, instead of inside the brain, and largely circumvent (rather than resolve) the mind-brain problem. Then the equivalence of the experience to the collapse at the level of the measured/observed system would become true essentially by construction: one defines the projectors at the level of the measured/observed system in a way such that they correspond to the distinct perceptual possibilities. But from a non-subjectivist viewpoint, one would like to have a characterization of the conditions for the external collapse that do not refer in any way to the observers. One way to circumvent the observers is to use the fact that the pointer interacts not only with observers but also with ``the environment'', which is imagined to be described by degrees freedom that will never be measured or observed. The representation of S given above will again hold with the $\phi_i$ now representing the states of the system being measured plus the measuring device, and the $\chi_i$ corresponding to states of the environment. The interaction between the pointer and the environment should quickly cause all the $\chi_i$ that correspond to different distinct locations of the pointer to become orthogonal. All observable projectors P are supposed to act nontrivially only on the states $\phi_i$: they leave unchanged all of the environmental states $\chi_i$.But then all observable aspects of the state S reside in tr S, where tr stands for the trace over the environmental degrees of freedom. Let $P_i$ be a projector onto an eigenstate of tr S. Suppose one postulates that each of the allowed projectors P\_ is a sum over some subset of the $P_i$, or, equivalently that each possible P\_ commutes with tr S, and is unity in the space of the degrees of freedom of the environment. This rule makes each allowed P project onto a statistical mixture of pointer locations, in cases where these locations are distinct. So it give the sort of P's that would correspond to what observers can observe, without mentioning observers. The P's defined in this way commute with S. But then the effect of any von Neumann reduction is to transform S into S: the von Neumann reduction has no effect at all. The collapse would have no effect at all on the average over the alternative possible answers to the question of whether or not the collapse occurs. This nondependence of the average is of course an automatic feature of classical statistical mechanics. The theory being described here is a development of von Neumann's approach in the sense that it gives more ontological reality to the quantum state than the Copenhagen approach, and also in the sense that it follows von Neumann's suggestion (or what Wigner describes as von Neumann's suggestion) of bringing consciousness into the theory as a real player. But it differs from the models discussed above that are based on his theory of measurement. For it does not associate collapses with things like positions of pointers on measuring devices. The projectors P(E) associated experiences E are in terms of classical aspects of the electromagnetic fields in brains of observers. That would be in line with von Neumann's general idea, but he did not go into details about which aspects of the brain were the pertinent ones. Rather he circumvented the issue of the mind-brain connection by centering his attention on the external devices and their pointer-type variables. The classical aspects of the EM field are technically different from pointers because their interaction with the environment is mainly their interaction with the ions and electrons of the brain, and these are the very interactions that both create these aspects of these fields, and that are in part responsible for the causal effects of the experiences E through the action of the projectors P(E). So what was formerly an uncontrolled and unobservable environment that disturbed the causal connections is now the very thing that creates the coherent oscillatory structure through which our experiences control our brains. The effects of this switch will be examined in the next section. \vskip .1in \noindent {\bf 13. Efficacy of Knowings.} \vskip .1in A formalism for dealing with the classical part of the the electro-magnetic field, within quantum electrodynamics (QED), has been developed in Stapp (1983) and Kawai and Stapp (1995), where it was shown that this part dominates low-energy aspects, and is exactly expressed in terms of a unitary operator that contains in a finite way the terms that, if not treated with sufficient precision, lead to the famous infrared divergence problem in QED. This classical part is a special kind of quantum state that has been studied extensively. It is a so-called coherent state of the photon field. Essentially all of the low-energy contributions are contained within it, and the effects of emission and re-absorption are all included. However, different classically conceived current sources produce different ``classical fields'', and hence the full low-energy field is a quantum superposition of these classical states. Each such classical state is a combination (a product) of components each of which has a definite frequency. All of the electrons and ions in the brain contribute to each of these fixed frequency components, with an appropriate weighting determined by that frequency. Thus the description is naturally in the frequency domain, rather than in spacetime directly: spatial information is encoded in quantum phases of the various fixed frequency components. Each value is represented, actually, by a gaussian wave packet centered at that value, in a certain space, and hence neighboring values are represented by overlapping gaussian wave packets. To exhibit a basic feature I consider a system of just three of these states. Suppose state 2 has all of the correct timings to elicit some coordinated actions. It represents in this simple model the state singled out by the projector P = P(E). Suppose it is dynamically linked to some motor state, represented by state 3: the dynamical evolution carries 2 to 3. Let state 1 be a neighbor of state 2 such that the dynamical evolution mixes 1 and 2. (I use here the Schroedinger picture, for convenience.) The transition from 2 to 3 will tend to depopulate the coupled pair 1 and 2. This depopulation of the system 1 and 2 will occur naturally whether or not any von Neumann collapse associated with P occurs. The question is: Can a von Neumann collapse associated with P affect in a systematic way the rate of depopulation from the coupled pair 1 and 2. The answer is ``Yes'': it can speed up the emptying of the amplitude in the system 1 and 2 into the system 3 that represents the motor action. This means that the effect of repeatedly putting to nature the question associated with P can have the effect of producing the motor action more quickly than what the dynamics would do if no question was put: putting the question repeatedly can effect the probabilities, compared Bohm's model, in which there are no collapses. The quantum rules regarding the probability of receiving a `Yes', or alternatively a `No', are stricly observed. To implement the dynamical conditions suppose the initial state is represented, in the basis consisting of our three states 1, 2, and 3, by the Hermitian matrix S with $S_{1,1} = x$, $S_{2,2} = y$, $S_{1,2}= z$, $S_{2,1}=z^*$, and all other elements zero. Suppose the coupling between states 2 and 3 is represented by the unitary matrix U with elements $U_{1.1} = 1$, and $$ U_{2,2}=U_{2,3}=U_{3,3}= -U_{3,2}= r = (2)^{-1/2}, $$ with all other elements zero. The mixing between the states 1 and 2 is represented by the unitary matrix M with $$ M_{1,1}=c, M_{1,2}= s, M_{2,1}= -s^*, M_{2,2}=c^*, M_{3,3}=1, $$ with all other elements zero. Here $c^* c + s^* s$ = 1. The initial probability to be in the state 2 is given Trace PS = y, where P projects onto state 2. The action of U depopulates state 2: $Trace PUSU^{-1}= y/2$. Then the action of the mixing of 1 and 2 generated by M brings the probability of state 2 to $$ Trace PMUSU^{-1}M^{-1} = (xs^* s)+ (yc^* c/2) -zcs^*r -z^*c^*sr, $$ where r is one divided by the square root of 2 For the case c = s = r this gives for the probability of state 2: $$ (xs^* s) + (yc^* c/2) -zcs^*r - z^*c^*sr = x/2 + y/4 - zr/2 -z^*r/2 $$ Since states 1 and 2 are supposed to be neighbors the most natural initial condition would be that the feeding into these two states would be nearly the same: the initial state would be a super position of the two states with almost equal amplitudes. This would make x = y = z = $z^*$. Then the probability of state 2 becomes $$ prob = y/2 + y/4 -yr $$ Then the effect of the mixing M is to decrease from y/2 the probability in the state 2 that feeds the motor action. If the question E, with P(E)= P, is put to nature before U acts, then the effect of the corresponding von Neumann reduction is to set z to zero. Hence in this case $$ prob = y/2 + y/4, $$ and the probability is now increased from y/2. Thus putting the question to Nature speeds up the motor response, on the average, relative to what that speed would be if the question were not asked. The point of this calculation is to establish that this theory allows experiences to exercise real control over brain activity, not only by making the individual choices between possibilities whose probabilities are fixed by the quantum rules, but also at a deeper level by shaping, through the choices of which questions are put to nature, those statistical probabilities themselves. This opens the door both to possible empirical tests of the presence of collapses of the kind predicated in this theory, and to a natural-selection-driven co-evolution of brains and their associated minds. \vskip .1in \noindent {\bf 14. Natural Selection and the Evolution of Consciousness.} \vskip .1in In a naturalistic theory one would not expect consciousness to be present in association with a biological system unless it had a function: nothing as complex and refined as consciousness should be present unless it enhances the survival prospects of the system in some way. This requirement poses a problem for a classically described system because there consciousness is causally non-efficatious: it is epiphenomenal. Its existence is not, under any boundary conditions, {\it implied} by the principles of classical physics in the way that what we call ``a tornado'' is , under appropriate boundary conditions, implied by the principles of classical physics. Consciousness could therefore be stripped away without affecting the behavior of the system in any way. Hence it could have no survival value. Consider two species, generally on a par, but such that in the first the survival-enhancing templates for action are linked to knowings, in the way described above, but in the second there is no such linkage. Due to the enhancement effects described in the preceding section the members of the first species will actualize their survival-enhancing templates for action faster and more often than the members of the second species, and hence be more likely to survive. And over the course of generations one would expect the organism to evolve in such a way that the possible experiences E associated with it, and their consequences specified by the associated projection operators P(E), will become ever better suited to the survival needs of the organism. \vskip .1in \noindent {\bf 15. What is Consciousness?} \vskip .1in When scientists who study consciousness are asked to define what it is they study, they are reduced either to defining it in other words that mean the same thing, or to defining it ostensively by directing the listener's attention to what the word stands for in his own life. In some sense that is all one can do for any word: our language is a web of connections between our experiences of various kinds, including sensations, ideas, thoughts, and theories. If we were to ask a physicist of the last century what an ``electron'' is, he could tell us about its ``charge'', and its ``mass'', and maybe some things about its ``size'', and how it is related to ``atoms''. But this could all be some crazy abstract theoretical idea, unless a tie-in to experiences is made. However, he could give a lengthy description of this connection, as it was spelled out by classical physical theory. Thus the reason that a rational physicist or philosopher of the ninteenth century could believe that ``electrons'' were real, and perhaps even ``more real'' than our thoughts about them, is that they were understandable as parts of a well-defined mathematical framework that accounted---perhaps not directly for our experiences themselves, but at least---for how the contents of our experiences hang together in the way they do. Now, however, in the debate between materialists and idealists, the tables are turned: the concepts of classical physics, including the classical conception of tiny electrons responding only to aspects of their local environment, absolutely cannot account for the macroscopic phenomena that we see before our eyes. On the contrary: the only known theory that does account for all the empirical phenomena, and that is not burdened with extravagent needless ontological excesses, is a theory that is neatly formulated directly in terms of our knowings. So the former reason for being satisfied with the idea of an electron, namely that it is part of a parsimonious mathematical framework that accounts quantitatively for the contents of our experiences, and gives us a mathematical representation of what persists during the intervals between our experiences, has dissolved insofar as it applies to the classical idea of an electron: it applies now, instead, to our knowings, and the stored compendium of all knowings, the Hilbert space state of the universe. To elicit intuitions, the classical physicist might have resorted to a demonstration of tiny ``pith balls'' that attract or repel each other due to (unseen) electric fields, and then asked the viewer to imagine much smaller versions of what he sees before his eyes. This would give the viewer a direct intuitive basis for thinking he understood what an electron is. This intuitive reason for the viewer's being satisfied with the notion of an electron as an element of reality is that it was a generalization of something very familiar: a generalization of the tiny grains of sand that are so common in our ordinary experience, or of the tiny pith balls. No things are more familiar to us than our own experiences. Yet they are elusive: each of them disappears almost as soon as it appears, and leaves behind only a fading impression, and fallible memories. However, I shall try in this section to nail down a more solid idea of what a conscious experience is: it unifies the theoretical and intuitive aspects described above. The metaphor is the experienced sound of a musical chord. We have all experienced how a periodic beat will, when the frequency is increased, first be heard as a closely spaced sequence of individual pulses, then as a buzz, then as a low tone, and then as tones of higher and higher pitch. A tone of high pitch, say a high C, is not experienced by most listeners as a sequence of finely spaced individual pulses, but as something experientially unigue. The same goes for major and minor chords: they are experienced differently, as a different gestalts. Each chord, as normally experienced, has its own unique total quality, although an experienced listener can attend to it in a way that may reveal the component elements. One can generalize still further to the complex experience of a moment of sound in a Beethoven symphony. These examples show that a state that can be described physically as a particular combination of vibratory motions is experienced as a particular experiential quality: what we cannot follow in time, due to the rapidity of the variations, is experienced as a gestalt-type impression that is a quality of the entire distribution of energy among the sensed frequencies. According to the theory purposed here, the aspect of brain dynamics that corresponds to a conscious experience is a complex pattern of reverberating patterns of EM excitations that has reached a stable steady state and become a template for immediate further brain action. Its actualization by a quantum event initiates that action: it selects out of an infinite of alternative competing and conflicting patterns of neural excitations a single coherent energetic combination of reverberating patterns that initiates, quides, and monitors, an ongoing coordinated evolution of neural activities. The experience that accompanies this suddenly-picked-out ``chord'' of reverberations is, I suggest, the ``quality'' of this complex pattern of reverberations. Because the sensed combinations of EM reverberations that constitute the template for action is far more complex than those that represent auditory sounds, the quality of the former chord must be far more complex than that of the latter. But the most important quality of our experiencess is that they have meanings. These meanings arise from their intentionalities, which encompass both intentions and attentions. The latter are intentions to attend to---and thereby to update the brains representation of---what is attended to. These aspects of the experience arise from their self-reproducing quality: their quality of re-creating themselves. In the case of our human thoughts this self-reproductive feature has evolved to the point such that the present thought contains a representation of what will be part of a subsequent thought: the present experience E contains an image of a certain prolongation (projection into the future) of the current Jamesian sequence of temporal components that is likely, by virtue of the causal effect of E, namely S--$>$ PSP, with P = P(E), to be the current Jamesian sequence of a subsequent experience $E'$. Thus the meaning of the experience, through physically imbedded in the present state of the brain that it engenders, consists of the image of the future that it is likely to generate, within the context of its fringe. \vskip .2in {\bf Acknowledgements} This article is essentially a reply to detailed questions about earlier works of mine raised by Aaron Sloman, Pat Hayes, Stan Klein, David Chalmers, William Robinson, and Peter Mutnick. I thank them for communicating to me their dissatisfactions. I also thank Gregg Rosenberg and John Range for general support. \vskip .2in {\bf References.} Bohr, N (1934), {\it Atomic Theory and the Description of Nature} (Cambridge: Cambridge University Press). Bunge, M. (1967), {\it Quantum Theory and Reality} (Berlin: Springer). Einstein, A. (1951) {\it Albert Einstein: Philosopher-Scientist} ed. P.A. Schilpp (New York: Tudor). Fogelson, A.L. \& Zucker, R.S. (1985),`Presynaptic calcium diffusion from various arrays of single channels: Implications for transmitter release and synaptic facilitation', {\it Biophys. J.}, {\bf 48}, pp. 1003-1017. Feynman, R., Leighton, R., and Sands, M., (1965) {\it The Feynman Lectures in Physics.} (Vol. III, Chapter 21).(New York: Addison-Wesley). Haag, R. (1996) {\it Local Quantum Physics} (Berlin: Springer).p 321. Heisenberg, W. (1958a) `The representation of nature in contemporary physics', {\it Deadalus} {bf 87}, 95-108. Heisenberg, W. (1958b) {\it Physics and Philosophy} (New York: Harper and Row). Hendry, J. (1984) {\it The Creation of Quantum Theory and the Bohr-Pauli Dialogue} (Dordrecht: Reidel). Kawai, T and Stapp, H.P. (1995) `Quantum Electrodynamics at large distance I, II, III', {\it Physical Review}, {\bf D 52} 3484-2532. Joos, E. (1986) `Quantum Theory and the Appearance of a Classical World', {\it Annals of the New York Academy of Science} {\bf 480} 6-13. Omnes, R. (1994) {\it The Interpretation of Quantum Theory}, (Princeton: Princeton U.P.) p. 498. Stapp, H.P. (1975) `Bell's Theorm and World Process', {\it Nuovo Cimento} {\bf 29}, 270-276. Stapp, H.P. (1977) `Theory of Reality', {\it Foundations of Physics} {\bf 7}, 313-323. Stapp, H.P. (1979) `Whiteheadian Approach to Quantum Theory', {\it Foundations of Physics} {\bf 9}, 1-25. Stapp, H.P. (1983) `Exact solution of the infrared problem' {\it Physical Review}, {\bf 28} 1386-1418. Stapp, H.P. (1993) {\it Mind, Matter, and Quantum Mechanics} (Berlin: Springer), Chapter 6.\\ \& http://www-physics.lbl.gov/\~{}stapp/stappfiles.html Stapp, H.P. (1996) `The Hard Problem: A Quantum Approach', {\it Journal of Consciousness Studies}, {\bf 3} 194-210. Stapp, H.P. (1997) `Nonlocal character of quantum theory', {\it American Journal of Physics}, {\bf 65}, 300-304.\\ For commentaries on this paper see:\\ http://www-physics.lbl.gov/\~{}stapp/stappfiles.html\\ The papers quant-ph/yymmnnn cited there can be accessed at\\ quant-ph@xxx.lanl.gov by putting in the subject field the command:\\ get yymmnnn Stapp, H.P. (1997a) `Science of Consciousness and the Hard Problem', {\it J. of Mind and Brain}, vol 18, spring and summer. Stapp, H.P. (1997b) `The Evolution of Consciousness',\\ http://www-physics.lbl.gov/\~{}stapp/stappfiles.html Wigner, E. (1961) `The probability of the existence of a self-reproducing unit', in {\it The Logic of Personal Knowledge} ed. M. Polyani (London: Routledge \& Paul) pp. 231-238. Zucker, R.S. \& Fogelson, A.L. (1986), `Relationship between transmitter release and presynaptic calcium influx when calcium enters through disrete channels', {\it Proc. Nat. Acad. Sci. USA}, {\bf 83}, 3032-3036. Zurek, W.H. (1986) `Reduction of the Wave Packet and Environment-Induced Superselection', {\it Annals of the New York Academy of Science} {\bf 480}, 89-97 Zurek, W.H., S. Habib, J.P. Paz, (1993) `Coherent States via Decoherence', Phys. Rev. Lett. {\bf 70} 1187-90. \newpage \noindent {\bf Appendix A. Quantum Effect of Presynaptic Calcium Ion Diffusion.} Let me assume here, in order to focus attention on a particular easily analyzable source of an important quantum effect, that the propagation of the action potential along nerve fibers is well represented by the classical Hodgson-Huxley equation, and that indeed all of brain dynamics is well represented by the classical approximation apart from one aspect, namely the motions of the pre-synaptic calcium ions from the exit of the micro-channels (through which they have entered the nerve terminal) to their target sites. The capture of the ion at the target site releases a vesicle of neurotransmitter into the synaptic cleft. The purpose of the brain activity is to process clues about the outside world coming from the sensors, within the context of a current internal state representing the individual's state of readiness, in order to produce an appropriate ``template for action'', which can then direct the ensuing action (Stapp, 1993). Let it be supposed that the classically described evolution of the brain, governed by the complex nonlinear equations of neurodynamics, will cause the brain state move into the vicinity of one member of a set of attractors. The various attractors represent the various possible templates for action: starting from this vicinity, the state of the classically described body/brain will evolve through a sequence of states that represent the macroscopic course of action specified by that template for action. Within this classically described setting there are nerve terminals containing the presynaptic calcium ions. The centers of mass of these ions must be treated as quantum mechanical variables. To first approximation this means that each of these individual calcium ions is represented as if it were a statistical ensemble of classically conceived calcium ions: each individual (quantum) calcium ion is represented as a cloud or swarm of virtual classical calcium ions all existing together, superposed. This cloud of superposed virtual copies is called the wave packet. Our immediate interest is in the motion of this wave packet as it moves from the exit of a microchannel of diameter 1 nanometer to a target trigger site for the release of a vesicle of neurotransmitter into the synaptic cleft. The irreducible Heisenberg uncertainty in the velocity of the ion as it exits the microchannel is about $1.5$ m/sec, which is smaller than its thermal velocity by a factor of about $4 \times 10^{-3}$. The distance to the target trigger site is about $50$ nanometers. (Fogelson,1985;Zucker,1986) Hence the spreading of the wave packet is of the order of $0.2$ nanometers, which is of the order of the size of the ion itself, and of the target trigger site. Thus the decision as to whether the vesicle is released or not, in an individual instance, will have a large uncertainty due to the large Heisenberg quantum uncertainty in the position of the calcium ion relative to the trigger site: the ion may hit the trigger site and release the vesicle, or it may miss it the trigger site and fail to release the vesicle. These two possibilities, yes or no, for the release of this vesicle by this ion continue to exist, in a superposed state, until a ``reduction of the wave packet'' occurs. If there is a situation in which a certain particular {\it set of vesicles} is released, due to the relevant calcium ions having been captured at the appropriate sites, then there will be other nearby parts of the (multi-particle) wave function of the brain in which some or all of the relevant captures do not take place---simply because, for those nearby parts of the wave function, the pertinent calcium ions miss their targets---and hence the corresponding vesicles are not released. More generally, this means, in a situation that corresponds to a very large number N of synaptic firings, that, until a reduction occurs, all of the $2^N$ possible combinations of firings and no firings will be represented with comparable statistical weight in the wave function of the brain/body and its environment. Different combinations of these firings and no firings can lead to different attractors, and thence to very different macroscopic behaviours of the body that is being controlled by this brain. The important thing, here, is that there is, {\it on top of the nonlinear classically described neurodynamics}, a quantum mechanical {\it statistical effect} arising from the spreading out of the wave functions of the centers of mass of the various presynaptic calcium ions relative to their target trigger sites.The spreading out of the wave packet is unavoidable, because it is a consequence of the Heisenberg uncertainty principle. This spreading is extremely important, because it entails that every vesicle release will be accompanied by a superposed alternative situation of comparable statistical weight in which that vesicle is not released. This means that wave function of the entire brain must, as a direct consequence of the Heisenberg uncertainty principle, disperse into a shower of superposed possibilities arising from all the different possible combinations of vesicle releases or non-releases. Each possibility can be expected to evolve into the neighborhood of some one of the many different attractors. These different attactors will be brain states that will evolve, in turn, if no reduction occurs, into different possible macroscopic behaviors of the brain and body. Thus the effect of the spreadings of the wave functions of the centers of the presynaptic calcium ions is enormous: it will cause the wave function of the person's body in its environment to disperse, if no reduction occurs, into a profusion of branches that represent all of the possible actions that the person is at all likely to take in the circumstance at hand. The eventual reduction of the wave packet becomes, then, the decisive controlling factor: in any given individual situation the reduction selects---from among all of the possible macroscopically different large-scale bodily actions generated by the nonlinear (and, we have supposed, classically describable) neurodynamics---the single action that actually occurs. In this discussion I have generated the superposed macroscopically different possibilities by considering only the spreading out of the wave packets of the centers-of-mass of the pertinent presynaptic calcium ions relative to the target trigger sites, imagining the rest of the brain neurodynamics to be adequately approximated by the nonlinear classically describable neurodynamics of the brain. Improving upon this approximation would tend only to increase the quantum effect I have described. It should be emphasized that this effect is generated simply by the Heisenberg uncertainty principle, and hence cannot be simply dismissed or ignored within a rational scientific approach. The effect is in no way dependent upon macroscopic quantum coherence, and is neither wiped out nor diminished by thermal noise. The shower of different macroscopic possibilities created by this effect can be reduced to the single actual macroscopic reality that we observe only by a reduction of the wave packet. \newpage {\bf Appendix B. Knowings, Knowledge, and Causality.} I shall flesh out here the idea that Nature is built out of knowings, not matter. A typical knowing of the kind that quantum theory is built upon is a knowing that the pointer on the measuring device appears to lie between the numbers 6 and 7 on the dial. This is the sort of fact that all (or at least most) of science is built upon. It is quite complex. The idea that the appearance pertains to a dial on something that acts as a measuring device has a tremendous amount of education and training built into it. Yet somehow this knowing has this background idea built into it: that idea is a part of the experience. William James says about perceptions: ``Your acquaintance with reality grows literally by buds or drops of perception. Intellectually and upon reflection you can divide these into components, but as immediately given they come totally or not at all.'' This fits perfectly with Copenhagen quantum theory, which takes these gestalts as the basic elements of the theory. In the von Neumann/ Wigner type ontology adopted here there is, in association with this knowing, a collapse of the state vector of the universe. It is specified by acting on this state with a projection operator that acts on the degrees of freedom associated with the brain of the perceiver, and that reduces the state of the body/brain of the observer, and consequently also the state of the whole universe, to the part of that state that is compatible with this knowing. So a knowing is a complex experiential type of event that, however, according to the theory, occurs in conjunction with a correspondingly complex ``physical'' event that reduces the state of the the brain/body of the person to whom the experience belongs to the part of that state that is compatible with the knowing. [I shall use the word ``physical'' to denote the aspect of nature that is represented in the Hilbert-space description used in quantum theory: this aspect is the quantum analog of the physical description of classical physics.] That ``person'' is a system consisting of a sequence of knowings bound together by a set of tendencies that are specified by the state of the universe. This state is essentially a compendium of prior knowings. However, these knowings are not merely human knowings, but more general events of which human knowings are a special case. In strict Copenhagen interpretation quantum theory is regarded as merely a set of rules for making predictions about human knowledge on the basis of human knowledge: horses and pigs do not make theoretical calculations using these ideas about operators in Hilbert space, and their ``knowings'' are not included in ``our knowldge. But in a science-based {\it ontology} it would be unreasonable to posit that human knowledge plays a singular role: human knowings must be assumed to be particular examples of a general kind of ``knowings'' that would include ``horse knowings'' and ``pig knowings''. These could be degraded in many ways compared to human knowings, and perhaps richer in some other dimensions, but they should still be of the same general ontological type. And there should have been some sort of things of this general ontological kind even before the emergence of life. [In the section, ``What is Consciousness'', I have tried to provide an intuition about what a knowing associated with a nonbiological system might be like.] Science is an ongoing endeavor that is expected to develop ever more adequate (for human needs) ideas about the nature of ourselves and of the world in which we find ourselves. Newton himself seemed to understand this, although some of his successors did not. But the present stage of theoretical physics makes it clear that we certainly do not now know all the answers to even the most basic questions: physics is still very much in a groping stage when it comes to the details of the basic underlying structure. So it would be folly, from a scientific perspective, to say that we must give specific answers now to all questions, in the way that classical physics once presumed to do. This lack of certainty is highlighted by the fact that the Copenhagen school could claim to give practical rules that worked in the realm of human knowledge without paying any attention to the question of how nonhuman knowings entered into nature. And no evidence contrary to Copenhagen quantum theory has been established. This lack of data about nonhuman knowledge would make it presumptuous, in a science-based approach, to try to spell out at this time details of the nature of nonhuman knowings, beyond the reasonable presumption that animals with bodies structurally similar to the bodies of human beings ought, to the extent they also behave like human beings, to have similar experiences. But knowings cannot be assumed to be always exactly the kinds of experiences that we human beings have, and they could be quite different. The knowings that I mentioned at the outset were percepts: knowings that appear to be knowings about things lying outside the person's body. But, according to the von Neummann/ Wigner interpretation, each such knowing is actually connected directly to the state of the person's body/brain, after that event has occurred. This state of the body/brain will, in the case of percepts of the external world, normally be correlated to aspects of the state of the universe that are not part of the body/brain. But experienced feelings, such as the feelings of warmth, joy, depression, devotion, patriotism, mathematical understandings, etc. are not essentially different from percepts: all are experiences that are associated with collapse events that reduce the state of the body/brain to the part of it that is compatible with the experience.. I have spoken here of a body/brain, and its connection to an experience. But what is this body/brain? It seems to be something different from the knowing that it is connected to. And what is the nature of this connection? The body/brain is an aspect of the quantum mechanically described state of the universe. This Hilbert-space state (sometimes called density matrix) is expressed as a complex-valued function of two vectors, each of which is defined over a product of spaces, each of which corresponds to a degree of freedom of the universe. Any system is characterized by a certain set of degrees of freedom, and the state of that system is defined by taking the trace of the state of the universe over all other degrees of freedom, thereby eliminating from this state any explicit reference to those other degrees of freedom. In this way the state of each system is separately definable, and dependent only on its own degrees of freedom, even though the system itself is basically only an aspect of the whole universe. Each part (i.e., system) is separately definable, yet basically ontologically inseparable from the whole: that is the inescapable basic message of quantum theory. Each system has a state that depends only on its own degrees of freedom, and this system, as specified by its state, is causally pertinent, because each knowing is associated with some system, and the probabilities for its alternative possible knowings are specified by its own state, in spite of the fact that the system itself is fundamentally an inseparable part of the entire universe. It is the properties of the trace operation that reconciles these disparate requirements The state of the universe specifies only the probabilities for knowings to occur, and it generally undergoes an instantaneous global instantaneous jump when a new knowing occurs. But this probability, by virtue of the way it jumps when a new knowing occurs, and suddenly changes in regions far away from the system associated with the new knowing, and that it is formulated in terms of infinite sets of pssibilities that may never occur, is more like an idea or a thought than a material reality. Indeed, these properties of the state are exactly why the founders of quantum theory were led to the conclusion that the mathematical formalism that they created was about knowledge. The state of the universe is the preserved compendium of all knowings. More precisely, it is an aspect of that compendium that expresses certain statistical properties pertaining to the next knowing. There is presumeably some deeper structure, not captured by the properties expressed in the Hilbert-space mathematical structure, that fixes what actually happens. The knowings that constitute our experiences are the comings into being of bits of knowledge, which join to form the knowledge that is represented by the state of the universe. This gives an ontology based on knowings, with nothing resembling matter present. But the statistical causal structure of the sequence knowings is expressed in terms of equations that are analogs of the mathematical laws that governed the matter postulated to exist by the principles of classical mechanics. This connection to classical mechanics is enough to ensure a close similarity between the predictions of classical mechanics and those of quantum mechanics in many cases of interest, even though the two theories are based on very different mathematical structures. If one starts from the ontological framework suggested by classical mechanics the questions naturally arise: Why should experiences exist at all? And given that they do exist, Why should they be composed of such qualities as sensations of (experiential) colors and (experiential) sounds, and feelings of warmth and coldness, and perceptions of simple geometric forms that correspond more directly to the shapes of structures outside the body/brain than to structures (such as patterns of neural excitations that are presumably representing these various features) inside the body/brain. How do these experiential types of qualities arise in a world that is composed exclusively of tiny material particle and waves? The experiential qualities are not constructible from their physical underpinnings in the way that all the physical properties of a tornado are, according to classical mechanics, constructible from its physical constituents. Quantum theory allows one to get around these questions by eliminating that entire classical ontology that did not seem to mesh naturally with experiential realities, and replacing that classical ontology with one built around experiential realities. These latter realities are embedded in a specified way, which is fixed by the pragmatic rules, into a mathematical structure that allows the theory to account for all the successes of classical mechanics without being burdened with its awkward ontological baggage. A discussion of this appendix with cognitive scientist Pat Hayes can be found on my website: \\ (http://www-physics.lbl.gov/`tilde'stapp/stappfiles.html), \\ where `tilde' stands for the tilde symbol. \newpage \noindent {\bf Appendix C. Quantum Wholism and Consciousness.} One reason touted for the need to use quantum theory in order to accomodate consciousness in our scientific understanding of brain dynamics is the seeming pertinence of quantum wholism to the unitary or wholistic character of the conscious experience. I shall here spell out that reason within the framework of a computer simulation of brain dynamics. Suppose we consider a field theory of the brain, with several kinds of interacting fields, say, for example, the electric and magnetic fields, and a field representing some mass- and charge-carrying field. Suppose the equations of motion are local and deterministic. This means that the evolution in time of each field value at each spacetime point is completely determined by the values of the various fields in the immediate neighborhood of that spacetime point. Suppose we can, with good accuracy, simulate this evolution with a huge collection of computers, one for each point of a cubic lattice of finely spaced spatial points, where each computer puts out a new set of values for each the fields, evaluated at that its own spatial point, at each of a sequence of finely spaced times. Each computer has inputs only from the outputs of its nearest few neighbors, over a few earlier times in the sequence of times. The outputs are digital, and the equations of motion are presumed to reduce to finite-difference equations that can be readily solved by the stripped-down computers, which can do only that. Thus, given some appropriate initial conditions at some early times, this battery of simple digital computers will grind out the evolution of the simulated brain. Merely for definiteness I assume that the spatial lattice has a thousand points along each edge, so the entire lattice has a billion points. Thus our simulator has a billion simple computers. Now suppose after some long time the field values should come to spell out a gigantic letter ``M'': i.e., the fields all vanish except on a set of lattice points that have the shape of a letter ``M'' on one of the faces of the lattice. If the outputs are printed out at the location of the corresponding grid point then you or I, observing the lattice, would know that the letter ``M'' had been formed. But would the battery of dynamically linked but ontologically distinct computers itself contain that information explicitly? None of the computers has any information in its memory except information about numbers pertaining to its immediate neighborhood: each computer ``knows'' nothing except what its immediate environment is. So nowhere in the battery of computers, B, has the higher-level information about the global structure been assessed and recorded: the fact that an ``M'' has been formed is not ``known'' to the battery of computers. Some other computer C, appropriately constructed, could examine the outputs of the various elements of B, and issue a correct statement about this global properties of B, but that global information is not explicity expressed in the numbers that are recorded in B itself: some extra processing would be needed for that. Of course, brains examine themselves. So B itself might be able to do the job that C did above, and issue the statement about its own global property, and also record that information in some way in the configuration of values in the various simple computers: the existence of this configuration can be supposed to have been caused by the presence of the ``M'', and can be supposed to cause, under appropriate conditions, the battery of computers B to display on some lattice face the message: ``I did contain an `M' ". So the information about the global structure is now properly contained in the structure of B, as far as causal functioning is concerned. But even though the configuration of values that carries the information about the ``M'' is correctly linked causally to past and future, this configuration itself {\it is} no more than any such configurations was before, namely a collection of tiny bits of information about tiny regions in space. There is nothing in this classical conception that corresponds ontologically to the entire gestalt, ``M'', as a whole. The structure of classical physics is such that the present reality is specified by values located within in an infinitesimal interval centered on the present instant, without any need to refer to any more distant times. To bring relationships to the past and future events into the present evolving ontological reality would be alien to the ideas of classical physics. There is simply no need to expand the idea of reality in this way: it adds only superfluities to the ontologically and dynamically closed structure of classical physics. The situation changes completely when one quantizes the system. To make a computer simulation of the quantum dynamics one generalizes the spatial points of the classical theory to super-points. Each possible entire classical state is a super-point. In our case, each super-point is defined by specifying at each of the points in the lattice a possible value of each of the several (in our case three) fields. To each super-point we assign a super-computer. If the number of discrete allowed values for our original simple computers was, say, one thousand possible values for each of the three fields, and hence $10^9$ possible output values in all for each simple computer, then the number of allowed classical states would be $10^9$ raised to the power $10^9$: each of the $10^9$ simple computers can have $10^9$ possible values. Thus the number of needed super-computers would be $10^9$ raised to the power $10^9$. In the dynamical evolution each of these super-computers generates, in succession, one complex number (two real numbers) at each of the times in the finely spaced sequence of times. One can imagine that a collapse event at some time might make all of these complex numbers, except one, equal to zero, and make the remaining one equal to $1$. Then the state would be precisely one of the $10^9$ to the power $10^9$ classical states. It would then evolve into a superposition of possible classical states until the next collapse occurs. But the collapse takes the state to a ``whole'' classical world. That is, each super-computer is associated not just with some tiny region, but with the whole system, and the collapses can be to states in which some whole region of spacetime has a fixed configuration of values. Thus, for example, there would be a super-computer such that its output's being unity would mean that ``M'' appeared on one face. And the collapse to that single state would actualize that gestalt ``M''. The sudden selective creation of this gestalt is more similar to someone's experiencing this gestalt than any occurrence or happening in the classical dynamics, because in both the experience and the quantum event the whole body of information (the whole ``M'') suddenly appears. This intuitive similarity of collapse events to conscious events is a reason why many quantum theorists are attracted to the idea that conscious events are quantum events. Orthodox quantum theory rests on that idea. There is in the quantum ontology a tie-in to past and future, because if one asks what the present reality is, the answer can be either knowledge of the past, or potentialities for the future: the present is an abrupt transition from fixed past to open future, not a slice of a self-sufficient continuous reality. \newpage \noindent {\bf Appendix D. The Dilemma of Free Will.} \vskip 0.1in The two horns of this dilemma are `determinism' and `chance'. If determinism holds then a person seems reduced to a mechanical device, no more responsible for his acts than a clock is responsible for telling the wrong time. But if determinism fails then his actions are controlled in part by ``chance'', rendering him {\it even less} responsible for his acts. This argument can powerfully affect on our lives: it allows us to rationalize our own moral failings, and it influences the way we, and our institutions, deal with the failings of others. It might appear that there is no way out: either the world is deterministic or it's not, and the second possibility involves chance. So we get hung on one horn or the other. Quantum ontology evades both horns. The point is that determinism does not imply mechanism. The reason we say we are not responsible if determinism holds is that ``determinism'' evokes the idea of ``mechanism''; it evokes the idea of a clock. And, indeed, that's exactly what {\it is} entailed by the determinism of classical mechanics. According to the principles of classical mechanics everything you will do in your life was fixed and settled before you were born {\it by local `myopic' mechanical laws}: i.e., by essentially the same sort of local mechanical linkages that control the workings of a clock. If your thoughts and ideas enter causally into the physical proceedings at all, it is only to the extent that they are themselves completely controlled by these local mechanical processes. Hence the causes of your actions can be reduced to a huge assembly of thoughtless microscopic processes. But in quantum dynamics our knowings enter as the central dynamical units. What we have is a dynamics of knowings that evolve according to the rules of quantum dynamics. To be sure these dynamical rules do involve elements of chance, but these are no more problematic than the thermal and environmental noise that occurred in the classical case: our high-level structures cannot maintain total fine control over every detail. But there is, in spite of that important similarity, a huge difference because in the classical case everything was determined from the bottom up, by thoughtless micro processes, whereas in the quantum case everything is determined from the top down, by a dynamics that connects earlier knowings to later knowings. And these knowings are doing what we feel they are doing: initiating complex actions, both physical and mental, that pave the way to future knowings. No reduction to knowingless process is possible because each step in the dynamical processes is the actualization of a knowing that is represented mathematically as the grasping, as a whole, of a structural complex that is equivalent to the structure of the knowing. \end{document}