The Science of Freedom: An Intriguing Perspective, Questioning Determinism Through Philosophy, Cognitive Neuroscience & Quantum physics (Popular Science)

So how can we know if human beings really have free will? The question of whether or not we have free will (libertariani

597 60 4MB

English Pages 592 [248] Year 2018

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Science of Freedom: An Intriguing Perspective, Questioning Determinism Through Philosophy, Cognitive Neuroscience & Quantum physics (Popular Science)

Citation preview

The Science of Freedom / Michael Abraham Copyright © 2018 Michael Abraham All rights reserved; No parts of this book may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, taping, or by any information retrieval system, without the permission, in writing, of the author.

Translation from the Hebrew: Avi Woolf Contact: [email protected]

Dedicated with love to my wife Dafna and my children Nachman, Bruria, Moshe Gershon, Shlomo, Yossi and Rivkah.

When you have eliminated the impossible, whatever remains, however improbable, must be the truth? – Sherlock Holmes The path of people who say we are masters of our deeds, what we do, we do from knowledge. I cannot say so, everything I do would seem that if I had control, I would not have done it. I will return to the matter and relate what happened after the opening was opened before me and I entered the house, we were in one room, which as most simple houses in our city did not have a corridor but rather a man opens the door which opens out into the street and finds himself standing in one of the rooms of the house. The room was neither small nor large nor even medium, but one of those rooms whose dimensions are hard to estimate due to their structure. – Shai Agnon, Tachrich Shel Sipurim

Contents Prologue Preface PART ONE DEFINING TERMS: FREEDOM, LIBERTY, AND CAUSALITY Chapter One The Debate on Free Will: An Outline Chapter Two A Phenomenology of Libertarianism Chapter Three On Morality and Fatalism Chapter Four Determinism: A Priori Considerations Chapter Five On the Origins and Meaning of Causality Chapter Six On Reasons, Causes and Freedom of Will Chapter Seven A Mathematical Model of the Three Fundamental Mechanisms PART TWO THE WHOLE WORLD IS PHYSICS? Chapter Eight Between the Mental and the Physical Chapter Nine Gaps in Physicalism: Chaos Chapter Ten Gaps in Physicalism: Quantum Theory Chapter Eleven Materialism: A Priori Considerations PART THREE FREEDOM OF WILL AND NEUROSCIENCE Chapter Twelve Neuroscience Chapter Thirteen Brain-Psyche Relations: Emergence Chapter Fourteen The Libet Experiments and Moral Decisions Chapter Fifteen The Wonders of the Damaged Brain PART FOUR ENDING: FREEDOM OF WILL AND DUALISM FROM A BIRD’S EYE VIEW Chapter Sixteen Test Yourself Chapter Seventeen Summary Epilogue Appendix

Introduction to the English edition “The Science of Freedom,” is the first of my books to be translated into English. I wrote it out of the feeling that the subject of freedom of will is not getting proper treatment in the popular literature, and in the professional literature as well, (in the fields of philosophy and neuroscience), its treatment is inadequate and includes many inaccuracies. My purpose in this book is to defend the concept of freedom of will, an approach that has become increasingly unacceptable in scientific and philosophical circles, mainly as the result of the rapid development of the neurosciences. To my delight, the book was a great success in Israel, and I received quite a bit of feedback that it fills a void that exists in the literature dealing with freedom of will and especially with regard to recent findings of the present generation of studies of brain research. My recognition of this deficiency and the need and importance of filling it convinced me to take the expensive and complex step of translating this book, in particular of all of my books, into English. I hope and trust that this English edition will be as well accepted throughout the world. My thanks to Yedioth Books in Israel for their assistance in publishing and distributing the book in Hebrew, and of course to the translator Avi Wolf, and to all of the staff of the English publisher, eBookPro and to their head Benny Carmi, and to Kim Ben Porat and Barry Rosenfeld, for their dedicated and professional work without which I could not have made it up to now. It is also my pleasure to thank Oren Margalit for his invaluable help and initiative. I would be very pleased to receive feedback and comments from readers to my email: [email protected]

Prologue1 COGITO ERGO SUM — “I THINK, THEREFORE I AM”

Rene Descartes (Cartesius in Latin) was a French philosopher, scientist, and mathematician who lived in the sixteenth and seventeenth centuries. His main claim to fame was his classical philosophical argument known as the “cogito,” based on the Latin phrase Cogito Ergo Sum, which is usually translated as “I think, therefore I am.” Descartes searched desperately for a solid foundation for the philosophy and science he wished to develop and which he ultimately did succeed in forming. He was greatly troubled by the lack of certainty in philosophy — that is, the fact that almost any argument (and certainly factual argument) could be accepted or rejected without there being a contradiction, and this means that every last thing we “know” is uncertain. As a result, the basis for every philosophical, theological, and scientific approach seemed shaky to him. It’s important to understand that Descartes lived and operated in the time before the growth of the British empiricism of Locke, Hume, and Berkeley which emerged very much as a counterpoint to his own rationalist form of thought. We should therefore not be surprised that as far as he was concerned, the supreme and primary tool for ensuring certainty was the intellect, that is thought, and not necessarily observation. As far as he was concerned, observation as such was not a reliable source for certain truths unless it is properly established — that is, based on truths derived from the intellect. For this reason, he could not rely on observation in trying to find a certain Archimedean point for his philosophy. He had to search for an argument which stood on its own — that is, which was not based on observation — of the sort which could not be doubted or questioned. His mode of thought is described in a number of the books which he wrote, and his thought process is a little different in every one of them with the formulation of his conclusions being appropriately different as well. Here we will point to the main outlines of his argument, as well as one difference between the various sources in his writings which will serve us below.

Descartes’ Skeptical Dive Descartes began his philosophical inquiry by doubting anything and everything which could be doubted. He was not willing to accept any arguments as given, even those which seemed obvious. He raised the possibility that a demon was deceiving him or perhaps some mental illness was forcing him to arrive at the wrong conclusions. At this skeptical stage, he was not willing to accept that what he sees actually exists, including the argument that God exists (something which he considered obvious). It’s important to understand that Descartes did not really believe in such a radically skeptical position. His skeptical attack was a deliberate philosophical line of thought whose purpose was to ultimately arrive at a rock solid argument which would serve as a solid basis for philosophy and science, which Descartes never stopped believing in throughout his inquiry. To find this foundation, he chose a methodological approach of doubting anything which could be doubted. He waited to see what was left after he removed all such

arguments and beliefs, in the hope that he would find his desired point of certainty, which to him was the only worthy solid foundation for his philosophy and science.

The Cogito Principle At the end of the process, Descartes arrived at the surprising conclusion that such an argument does indeed exist.2 There is an argument which cannot be dismissed in any way — that is, it cannot be doubted. His argument is that the very existence of man cannot be doubted by man himself. The consideration which led him to this conclusion is his famous cogito argument: “I think, therefore I am.” This is the formulation in his book Discourse on the Method 3 (p. 48). Another source in his writings4 contains a slightly different formulation: “I cast doubt, therefore I am.” If I doubt my very existence, this is a sign that there is someone doing the doubting (or someone who thinks it). Therefore it cannot be that I will doubt my very existence. If I doubt, then I (that is, the one doing the doubting) exist, and therefore the doubt itself disappears. Descartes is effectively arguing here that my ability to cast doubt is the proof of my existence. I can therefore doubt or question almost anything, but not my very presence.

The Innovation of the Cogito As many of his interpreters pointed out,5 this is a partial and inaccurate presentation of the argument. Every logical argument bases its conclusions on an assumption or assumptions, and if we reject one of these assumptions then we are not obligated to accept the conclusion. In our example, if someone comes and says “It isn’t true that I doubt” or “It isn’t true that I think,” then he is naturally exempt from the conclusion that “I am.” After all, our attitude towards the conclusion depends on the truth of the assumptions. However, if we base ourselves on any assumption, then we can propose a parallel assumption that goes as follows: “I walk, therefore I am.” Obviously the conclusion “I am” derives from the assumption “I walk,” in a manner no less necessary than the assumption “I think.” After all, if I don’t exist, how can I walk? So what’s so special about the assumption Descartes chose? Why did he decide to arrive at his conclusions from the assumption “I think” rather than, say, “I walk”? Here lies the essence of the argument: Descartes’ primary innovation was not in the logical derivation of the conclusion “I am” from the assumption “I think.” His primary innovation was the insight that “I think” is in and of itself a necessary argument, as opposed to “I walk.” If I think that I don’t think, that’s also a thought, and therefore I am thinking once again. In other words, “I think” is a necessary statement which cannot be denied. To deny it is as good as demonstrating its truth. Of course, one cannot say the same thing regarding the sentence “I walk.” If I think that I don’t walk, or doubt this assumption, this will not lead me to any contradiction. In other words, there is no necessary truth in the argument “I walk” (in philosophical terms, this would be called a contingent statement). In sum, Descartes’ primary innovation is not the derivation of the conclusion “I am” from the assumption “I think,” but rather the insight that the argument “I think” is in and of itself a necessary argument. The derivation of the conclusion “I am” from this assumption is simpler, and in fact contains little that is new.6 Obviously, if we logically derive any conclusion from a certain assumption, the conclusion will also necessarily be certain. Therefore, what is particularly important regarding this argument is the certainty of the assumption.

A Note on the Difference Formulations of the Cogito In Descartes’ main work, Meditations,7 on page 42, he proves the necessity of the argument “I think,” and only incidentally notes that this assumption leads to the conclusion “I am.” As I showed here, this is the most accurate logical presentation of the cogito argument. Let us now look at another formulation of the cogito: “I doubt, therefore I am.” The assumption “I doubt” is not at all necessarily true. Its refutation does not necessarily lead to a contradiction. Only the claim “I am” is necessary, as if I doubt my existence then it is clear that I exist (and if not — then I certainly exist). In other words, the formulation which derives “I am” from the assumption “I doubt my existence,” is a formulation by which the conclusion — “I am” — is what is certain rather than the assumption. The assumption “I doubt” is not necessarily true, and in this sense it is like the argument which bases existence on the assumption “I walk.” By contrast, the formulation which derives the conclusion from the assumption “I think” is based on an assumption which is certainly true, and therefore the conclusions which derive from this conclusion are also logically necessary. “I am” is the first conclusion from the foundation of certainty, and Descartes sees it as the beginning of the path to the other conclusions as well as the rest of his

scientific and philosophical thought.

What “Is” Really Exists? There is another point which can very easily be missed. Did Descartes prove the existence of the body with his argument? The structure of the argument shows that what he proved was the existence of the spirit, our thinking self, and not the body. What this argument proves is that the “is” that thinks of himself (that is, intellect and consciousness) exists. The “I” whose existence we proved, is our spirit, not the body. Even after the cogito, the existence of our body is not necessarily demonstrated. Put differently: If we assumed that our intellect exists but we don’t have a body, we would arrive at no contradiction. This is an entirely consistent picture, and the cogito principle does not refute it. Thus, the statement “My body is” is not necessarily demonstrated. The statement whose necessity is proven by the cogito is the statement “My spirit exists.” Descartes himself felt this clearly in various places in his writings. In Meditations, pp. 42-43, there is a sharper formulation of the cogito: I need to decide and think confidently that this statement: “I am, I am present” is necessarily correct, so long as I express it or conceive of it in my spirit. We see here that the statement “I am” is considered by Descartes as necessarily true only when he expresses it or spiritually conceives of it. We see here the great dependence between thought and the existence of a thinking “is” (rather than a body). Therefore, if we wish to be careful with this argument, we must formulate the conclusion as follows: Existence is necessarily true only in those moments when it is thought. When I deal with the question of whether or not I am, then there, at that moment, I truly am. But in principle, it may be that I am not in other moments. The fact that I arrive at this conclusion whenever I ask this question does not necessarily demonstrate that I always am, only that I am precisely in those moments in which I ask the question.8

What is More Certain to Exist: Body or Spirit? We saw that the certain anchor of “I think” leads the cogito argument to its first necessary conclusion: “I (i.e., the intellect which thinks of itself) am.” But this is merely the beginning of his rationalist path. From here, Descartes continues to build his entire philosophy on this basis, as well as his theology (the existence of God), and even his science. More specifically, the existence of the body is also a direct consequence of the existence of the spirit. In other words, after I concluded that my spirit and consciousness exist, I am also ready to accept the contents of my consciousness — that is, the material world, including my body, as well as God. The next part of his analysis is far weaker, and it seems that his proofs of the existence of God and of the material world and the body are substantially less robust than the cogito argument itself which serves as the fundamental basis for all this. Ostensibly, this conclusion contradicts our most basic intuitions. If we ask someone what they think to be more certain — the existence of the body or the existence of the spirit — we are more likely, certainly in our generation, to receive the answer that the existence of the body is more certain. Throughout this book, we will see that the existence of the spirit is a matter of dispute, while the existence of the body is taken for granted. It’s not clear to people if there is such a thing as an intellect or will (even though no-one denies the existence of thought and desire, so far as I know), but it’s clear to everyone that we have a brain. By contrast, Descartes’ argument presents us with an entirely different picture: The existence of the thinking spirit is what is clear and entirely certain; the existence of the body is merely a conclusion (necessary in his view, though not mine) of the assumption that the spirit exists. According to him, after the existence of the spirit was proven — and only then — can we believe the observations which it brings to our knowledge such as the existence of the body. If we go further with this surprising conclusion, we can state in a more general manner that we have no way of knowing if, say, the chair next to us exists. What is clear is that there is an image of the chair in our consciousness (the result of data which reach us via our sense of vision and sense of touch). The conclusion that there is a “real” chair out there whose image we see is speculation whose credibility is not entirely clear. What is certainly before us in an unmediated manner is the conscious image of the chair and not the chair itself.9 We can therefore say that the existence of the chair is certainly less certain than the existence of the image, as the conclusion will never be more certain than the assumptions leading up to it. There’s no need to point out that this is true of any material object around us. In other words, even the existence of the brain and the physical processes occurring therein are conclusions deriving from the image created in the human consciousness observing both — the existence of these objects is dependent on the

recognition of our consciousness regarding them. What is certain is the consciousness which exists by us, while objective external existence is just a conclusion we derive from that. That conclusion is therefore dependent on assumptions regarding the credibility of the sense we used to arrive at it.10 It is therefore very strange that there are those who are still willing to deny the existence of a recognizing spirit but unquestionably accept what it recognizes (i.e., the existence of the body). The discussion here dealt with the question of materialism — that is, the question of the existence of the spirit or the materiality of the world. The main question the book will deal with is a different one, but as we will see, they converge to an extent: The question of determinism and the dilemma of whether or not man has free choice or will. In this context, there is also a principle parallel to that of the Cartesian one, which can be formulated thusly: cogito ergo eligo, or “I think therefore I choose.” This principle is the essence of this book, and its explicit formulation will be presented and discussed in the book’s Epilogue. 1 This philosophical prologue is a bit pedantic by necessity and can be skimmed over. The conclusion at the end should be read in full, however. 2 For an extensive analysis of the argument, as well as interpretations and critiques thereof, cf. A.Z. Brown, Sugiyat Hayesh, Magnes, Jerusalem, 1977, fourth chapter (below: Brown). 3 Renee Descartes, Discourse on the Method, translated by Yosef Or, Magnes, Jerusalem, 1974. 4 The Search for Truth, referenced by Brown, p. 126. 5 Cf. Brown, throughout the fourth chapter. 6 There were those who raised various objections to the argument, but we will not go into detail on that here. 7 Renee Descartes, Meditations on First Philosophy, translated by Yosef Or, Magnes, Jerusalem, 1977. 8 Physicists describe a similar phenomenon — the stroboscope effect. Think of a particle in a dark room which moves in a circle at a very high speed. There is a flashlight which illuminates a particular point along that path. The flashlight turns on and off at a rate we control. If the rate of the flashlight is different than the speed of the particle, sometimes we will notice the particle (when it reaches the precise point that is illuminated) and sometimes we won’t. Our conclusion will be that it moves. What will happen if we set the rate of the flashlight so that it operates at exactly the same rate as the particle spin (or a complete multiple)? In such a situation, we will see the particle at the same point every time. On the assumption that the rates are very high, we will be under the illusion that the particle stands in the same place all the time. 9 In this context, see Rafi Malach’s lecture “From Photons http://www.weizmann.ac.il/neurobiology/labs/malach/sites/neurobiology.labs.malach/files/photonsToFantasies_Hebrew.pdf

to

Fantasies”:

Malach claims there that vision does not photograph reality so much as build a model of it, which benefits us in some ways. This argument greatly supports the preference of the image of reality as opposed to the reliability of our conclusions regarding reality itself. 10 This line of thought is very similar to Schopenhauer’s argument. Kant distinguished between the world in and of itself (the noumenon) and the world is it appears before us (the phenomenon). He argued that we have no way of knowing things in and of themselves, only how they appear before us (or our consciousness). We have no way of knowing the chair in and of itself but only its image which appears in our consciousness. Schopenhauer pointed to the fact that there is one exception to this Kantian distinction: Our spirit. When I look inside myself, I directly encounter myself — that is, Michael Abraham himself — and not just its phenomenon (the image others see when they look at him).

Preface Modern man lives a split existence. On the one hand, he understands that his brain and body are part of the physical world and that both are subject to the laws of physics and the laws of nature in general. These laws certainly determine our bodily moves, but insights from neuroscience indicate that this is also true of mental events which take place within us (such as memory, various emotions, desire, or thought). On the other hand, everyone, regardless of philosophical outlook, feels that he has the option of choosing between the various options he encounters, that is that his next step is not absolutely determined by present circumstances and the laws of nature. How, if at all, can one reconcile these two arguments? Must we choose (!) only one of them and abandon the other? Are we condemned to choose (!) between science and free will? Many researchers in the field of neuroscience say yes: Not only must we choose, they say, but in addition, free will contradicts the scientific approach and the latest findings in the field of neuroscience. That is, they recommend choosing science and rejecting the “illusion” of free will. My main goal in this book is to show that this is not at all a foregone conclusion.

Materialism and Determinism This book touches on two central and very old questions in human thought: The question of materialism (which deals with the relations between the material and the spiritual) and the question of determinism (which deals with freedom of choice or will). Some argue that these two dilemmas are the most important questions in the history of philosophy. As opposed to other philosophical abstractions, these questions would appear to have cardinal consequences for our daily lives, at both the personal and social level. The issue of materialism regards the question of whether man is a different creature than all others, as he possesses a spirit as well as materiality, a fact which can lead to conclusions regarding his moral status and the value of his life as compared to that of all the other creatures in the universe. Is there a problem in cloning human beings, as opposed to other animals or plants? Is the killing of a human being fundamentally different than the killing of an animal? Is the obligation to look after human beings fundamentally different than the obligation to care for vegetation and animals, and maybe even the inanimate (as eco-ethics argues)? The question of determinism can also touch on these questions, as there are those who based the “man is more than beast” principle, the value of man and of his life, on his freedom of will, his ability to choose freely. Determinism also directly applies to more practical questions, such as that of the attribution of responsibility — moral, civic, or criminal — to human beings. If the drill I hold doesn’t work properly, I don’t punish it; at most I fix it. What, then, can we demand of a human being who acts immorally? Should he be judged and condemned or does he only need to be “fixed”? At the legal level, we face the question of whether he can be judged and punished, or perhaps we should “fix” him (though reeducation, like in the Soviet Gulags)? Another example: What can one demand of a child who misbehaves in class? Can he even be “guilty”? Should we punish him or fix him? Should we educate him or just handle him? Or maybe there’s no difference between the two? This book is largely focused on the question of determinism, which is an important question beyond the boundaries of the aforementioned philosophy of science. Materialism will be discussed in the background whenever it ties in with the determinism question. We should stress, however, that the connection between these two positions is not necessarily a given, at least not at the logical level. Prima facie, the materialistic argument — that the entire world, including man, is material — appears to reject the belief in freedom of choice (libertarianism),11 while the contrary argument, that the world is dualistic (containing material and spirit), is seen as enabling free choice. As we will see below, the connection between these two questions does exist, but it is not so simple. The question of determinism is the more important and controversial of the two. Many argue that even if man is an entirely material being, when he has the option of choice — there is room to impose moral responsibility on him, and perhaps even for adopting the belief that his life is of greater value than that of animals, and that it’s important to educate him and not just handle him, and so forth.

The Connection Between Materialism and Determinism in Ancient Greece: Democritus and Epicurus

Already 2,500 years ago in ancient Greece, Democritus (lived in Thrace in 460-370 BC) and Epicurus after him (the head of the philosophical “garden” in Athens, 341-270 BC) built an atomist model of the universe, according to which everything in the world is nothing more than the various compositions and relationships between material atoms. But there was no agreement on determinism between these two ancient thinkers. While Democritus espoused complete determinism, Epicurus argued that the atoms move randomly, thus ostensibly dismissing the Democritean determinism and allowing freedom of choice in his understanding of the world. He was a materialist, as he also believed that the soul is made up of atoms (as are the gods), but he was not a determinist. Ironically, it is the error of Epicurus in confusing randomness with freedom of choice, which we will deal with quite a bit, while the materialist worldview they both share will be less important for our discussion. Epicurus’ error repeats itself in various forms throughout history until today, including among important philosophers and scientists, both determinists and libertarians. Many identify randomness with the soul, seeing it as sufficient foundation for the libertarian approach; we will return to this issue quite often later on.

The Goal of the Book, or: What’s New Here? These issues are very old, as old as philosophy itself. This raises the obvious question: What more can one say and write on these issues which hasn’t already been addressed in thousands of years of philosophy? First, it’s important to understand that even if all this isn’t new, there’s a great deal of confusion on the issue, and it’s important to bring order to terms and arguments. Second, these questions continue to generate heated debate until today, perhaps especially today. One reason for this is the secularization and modernity which came in the wake of the enlightenment. These opened a wide berth for the entry of these ideas into the public sphere. It was difficult to speak of determinism in the religious world (although there are examples of this as well; see the Appendix), and it was even harder to discuss materialism. From the moment religious limitations were removed, the instinctive reticence toward the subjects also disappeared, and the way was now clear to discuss such issues and spread the word. Sometimes they are identified outright, unjustifiably, with secularism. But this is not the only reason for the reawakening of the discussion. For the past few decades, there has been a feeling that the two questions of determinism and materialism are slowly moving from the field of philosophy to that of science. From the seventeenth century and the beginning of modern science, determinism began to slowly move back to center stage. True, humanism bravely fought it in the next two hundred years and pushed it into a corner, but the determinist world view has made a comeback since the end of the nineteenth century in the wake of the theory of evolution of Darwin and his successors, and it has reached its peak in the contemporary field of neuroscience, which has been taking the world by storm in recent times. The “brain sciences” are an impressive integration between fields like physics, chemistry, modern biology, evolution, genetics, mathematics, and psychology. A combination of all these theoretical and experimental tools has led to a wide variety of insights regarding the functioning of the brains of living creatures, especially the brains of human beings. Over the past few decades, a materialist-deterministic consensus has emerged among neuroscientists the world over, to the point that talk of dualism and libertarianism has become almost “illegitimate,” and at the very least quite irregular in the landscape. Many researchers argue that we are finally on the road to discovering final scientific answers to these problems which have so vexed humanity for so long. Many researchers and thinkers even go further and claim that we already have those answers. According to them, neuroscience has already showed that the world is deterministic, and that man has no free choice. Others also add materialism to the basket of scientific findings. According to them, the ideas of spirit and soul, as well as freedom of choice, are illusions (even if useful ones). Some of them even offer evolutionary and scientific explanations for these “welcome illusions.” Michael Gazzaniga is one of the founders and pioneers of modern neuroscience, serving to this day as the head of the Sage Institute and the Institute for Neuroscience and Cognition at the University of California. At the beginning of his book Who’s In Charge?12 (below: Gazzaniga), Gazzaniga describes this common approach, and tries in his own way (incorrect, in my view; see below) to explain why it is incorrect. Despite his irregular approach, it’s pretty clear from his book that his position seems even to him as a sort of provocation or interesting engagement in devil’s advocacy, as in this way he comes out against the “powerful orthodoxy” in the field, as he puts it. Shimon Marom of the Haifa Technion, himself a neuroscientist, points to this phenomenon in his

fascinating article “The Eternal Temptation,”13 which begins as follows: The neurosciences have undergone a profound change in recent years, expressed in textbooks and in the professional and popular press. The discourse of neuroscience has become to introduce terms which traditionally belonged and were unique to the “soft” language of the humanities and the social sciences: consciousness, God, freedom, love, morality, rationality, political positions, and more. By the nature of things, the daily press and the science and knowledge journals meant for the non-professional audience provide a generous amount of room for such exciting articles in this vein. Here, I wish to raise heretic thoughts, a different position, which is shared by quite a few scientists. I wish to shed a light, from the perspective of the average physiologist, on the problematic, hidden aspect regarding the interpretation of modern findings in the neurosciences, a side which the general public is usually less exposed to. I wish to dissipate, even if just a little, the arrogance which neuroscientists have been given to the past few years. The neurosciences today, as reflected in the scientific literature and the titles of lectures at professional conferences, have a new dream: To understand the origin of spirit from matter. Some of my colleagues certainly feel unease upon reading this sentence, mumbling something to themselves about how understanding the way in which spirit derives from material is not a dream they share, and does not even appear to be at all related to the fields of science. But how else can we interpret the titles of chapters in some of the most recent textbooks on neuroscience: “Cellular Mechanisms of Study and the Biological Basis for Individuality,” “Language, Thought, Mood,” “Emotion and Emotional States,” “Consciousness and the Neurobiology of the Twenty First Century,” “Neurobiology of Decision Making”…? At a significant number of research institutes related to neuroscience, there’s a feeling that the broader public discourse, still speaking in terms of freedom of will and choice or on the dualism of matter and spirit, is simply unaware of the latest scientific insights, as not all of them have been made public. Indeed, it seems that when that does happen, many of us are in for a shock. Our fundamental moral yardsticks may suffer a serious blow (whether justified or not). This may be the reason that these insights are leaked at a fairly slow pace to the public, as well as the fact that their portion of popular literature is fairly small. It would appear that a serious, systematic discussion on the issue of determinism, especially in light of the latest scientific findings, has yet to be conducted in the public sphere. Based on my own impressions, many researchers sometimes feel like pioneers, a sort of intellectual underground, which already knows a number of very important and very real things about our existence before we do. There’s a sort of attitude of mockery towards the simpletons outside, those who still don’t know the truth available to those who found or possess this Holy Grail. It would appear that their words are accepted by the public with some interest, but without real recognition of the important consequences of these arguments at the ethical level. Much like other issues, the situation in Israel is a great deal poorer than abroad. Elsewhere, there is a bitter and piercing discussion (if insufficiently so) on these questions from humanistic, legal, and religious perspectives. Conferences are held among philosophers, jurists, psychologists, and neuroscientists, with the aim of ironing out the difficult questions which these findings reveal. But Israel has been very largely quiet while these ideas slowly develop in laboratories and in research institutes, almost without any public discussion of their dramatic consequences for our lives and our values, and even prior to this — without a significant discussion of the ideas and findings themselves. Are these truly scientific arguments and findings, or perhaps they are interpretations of one sort or another —not necessarily correct ones — provided by certain men of science? Is there a reasonable justification for these far-reaching conclusions? One reason for the lack of discussion is the lack of the necessary skill and scientific knowledge to go into this field. People without the basic scientific knowledge in physics or psychology may find it difficult to participate in the discussion, and perhaps also when reading and understanding the findings. The psychological distinctions are sometimes quite nuanced and are not easy to understand and explain. It would seem that this avoidance has created a serious gap in the public discourse, failing to prepare the public for that expected shock, which for now is only leaking out one drip at a time. But the importance of such a discourse also has to do with the researchers themselves. Their philosophical conclusions (as opposed to their scientific ones) are sometimes adopted without any serious scrutiny. By the very nature of things, in such a situation, the picture presented to the public is the

worldview of that particular scientist presenting it, and as noted the approaches sold to the general public tend to look pretty monochromatic. Moreover, the situation in this field is such that philosophers lack the scientific skills and knowledge, while the scientists lack the necessary philosophical skills. As a result, the authority on these issues is being handed over by philosophers to the scientists, despite the clearly philosophical nature of the discussions and resulting arguments. Neuroscientists sometimes identify freedom of choice with randomness, and therefore search for the possibility of such freedom within the area of physics (and of course, unsurprisingly, do not find it there). Few of them have bothered to even define the concept of freedom of choice for themselves, and thus do not really know what they are looking for. The result is a structural defect in the discussion on these important issues. Gazzaniga (chapter six of his book, from p. 187) describes how recent years have seen brain scans even become a part of American courtrooms. Experts come at the behest of lawyers to diagnose if an action done by the defendant was done based on consciousness and responsibility. These scans also serve as part of the evidentiary fabric, as well as for the purposes of arguments from justice. This became even more established in 2002, when the Supreme Court in Atkins v. Virginia anchored the validity of such scans as the basis for the determination that the defendant is not competent and thus does not bear responsibility for his actions, or at least that the punishment imposed on such a person counts as “cruel and unusual” punishment per the eighth amendment to the United States Constitution. The Frye Standard determines the admissibility of scientific findings as evidence in court in many states in the US: “Scientific evidence will be admissible when the relevant technique, information, or method has earned general recognition in the relevant scientific community.” But who determines what this community is? Who establishes when this community has indeed led to these methods gaining general recognition? This is done by the judge, of course. However, judges usually don’t have the ability or the skills to determine whether it is a scientific or philosophical question, and as noted above most scientists don’t have that ability. Thus do these distortions become permanent and even receive legal grounding without any serious public scrutiny. As Gazzaniga himself describes, the experts themselves don’t often clarify that they are using statistical information which is shaky and open to interpretation, and the judges and jurors are certainly not equipped with the tools which would allow them to treat such scientific information with the appropriate scrutiny.14 He brings studies which show that incorrect explanations are more easily accepted when accompanied by exhibits such as brain scans, even if they have nothing to do with the arguments themselves — as he puts it there, the judges are “wowed by science.”15 If we take the neuroscientists’ own lack of legal and philosophical fluency into account, as they do not often understand the meaning of the concept of action without responsibility (a large portion of them, Gazzaniga included, thinks that all our actions are done without choice, so what does responsibility even mean as far as they are concerned?), the problem of course becomes all the greater, and the need to iron out these issues for all sides involved increases.16 This book aims to fill in that void. I cover the main arguments regarding materialism and determinism, at the scientific and the philosophical levels, and arrive at the conclusion (optimistic from my point of view) that the statements that suggest that science has already led us to the final victory of materialism or determinism, are greatly exaggerated. To the best of my judgment, they do not reflect the findings themselves, but rather, at most, an incautious interpretation thereof. It’s important to understand that although I openly consider myself to be a philosophical libertarian, the main goal of this book is not to prove that approach, but rather to show that there is no scientific reason to hold by this view, at least at the moment. Moreover, there are very strong reasons, even if not entirely dispositive, in its favor. My argument is that in such a situation, since the intuitions of most of humanity are libertarian (and we will see this primarily in the first section of the book as well as in chapter sixteen), it’s more reasonable to continue to adhere to those views. Very strong evidence would be required for me to become convinced that my most fundamental intuitive beliefs, and my most basic motivations for action, ethical or otherwise, are a collection of illusions. Beyond the examination of the philosophical-scientific situation in the present, the goal of this book is to make order in the fundamental terms and principles in this field in order to improve our ability to make decisions on these issues, including in light of future findings. We will see that many arguments suffer from conceptual fallacies. Many reach hasty and incautious conclusions, which are sometimes presented as the rock-solid results of scientific research. We thus end up with a situation in which many of the students of neuroscience at various universities hear unending and aggressive preaching in favor of the deterministicmaterialist thesis, even though there is no scientific justification or basis for this triumphalism at present.

Is There a Connection to the Secular-Religious Conflict? The Challenge to LiberalHumanism I already hinted at the connection, at least de facto, between these questions and positions on theological issues. Believing people are usually dualists and usually believe in freedom of choice. By contrast there is a feeling that a secular worldview finds itself increasingly dragged in the direction of materialism and determinism (this seems to be particularly prominent in Israel, which due to its small size is a country with a uniform scientific and philosophical approach). Oftentimes, debates between religious and secular people revolve around these very questions. It’s important to note that at least on the logical level, the consequences for theological positions are not necessary. There is nothing preventing a secular person from being a dualist or from being a libertarian. It is less likely that a religious person will be a materialist or a determinist, although that also exists (see the Appendix for more). But for some reason, despite the lack of a necessary logical connection, many secularists feel an almost religious obligation to adhere to the materialist-determinist position. This is fairly surprising when one understands the far from simple consequences of this position which were already mentioned above. How can a humanist and liberal secularist not become infuriated at approaches which see him as nothing more than a biological robot, a sort of animal himself? How can he speak of moral values, human responsibility, and the image of humanity, to say nothing of rights granted to human beings as such or the sanctity of human life, when he adopts determinist materialism at the same time? In a review of Yuval Noah Harari’s book A Short History of Mankind, who fell quite often into the sort of fallacies which characterize materialist determinism, I found the following fascinating statements:17 Thus, for instance, Harari makes a fascinating comparison between the Hammurabi Laws of the Babylonian Empire in 1776 BC to the American Constitution of 1776 CE, which shows that we’ve all been living in an illusion from the beginning. The two codes claim that they are drafting ‘universal and eternal principles of justice.’ According to Hammurabi, there is inherent inequality among human beings, dictated by the gods, while the American constitution states precisely the opposite but with the same divine sanction, stating that ‘all men were created equal.’ What’s interesting is that these two constitutions are not connected to any objective validity and exist only in the imagination of the faithful. Homo sapiens have no “natural” rights, just as grasshoppers have no such rights. Human beings use what Harari calls an “imagined order,” not because it is correct, but because if we believe in it we can create an order that will protect us. Because of this, Voltaire refused to tell his servant that there is no god, lest the latter murder him in his sleep. As we will see later on in the book, the forced attempts which desperately try to maintain freedom of will within the materialist-determinist picture (such as the approach known as compatibilism) are nothing more than determinism with a fig leaf on top, sometimes just a simple confusion of terms. I do not intend to argue that people need to force their a priori moral agenda on scientific findings or on their philosophical beliefs. Obviously an approach which rejects scientific findings based on their results not being moral, or uncomfortable for the management of a properly functioning society, suffers from lack of integrity and cannot hold water. I do not believe in the pragmatism of political correctness, the idea that that which is can be subjected to that which should be and the real to the worthy. I am attacking the root assumption: How can it be that only believing people feel great distress from these new ideas and their dramatic consequences for morality and our society? Why do we not find any real motivation in the world of general thought (at least, not in Israel) to examine the findings, and search for solutions which will leave freedom of choice in place even within the present scientific picture? How can such dramatic ideas be accepted with an almost stoic tranquility, and how does none of this matter at all to the terms and fundamental ethical approaches we hold dear? We continue to live comfortably within this duality, without forcing ourselves to account for it. Many thinkers have remained indifferent, and perhaps even excited at the theological consequences of the new results, without paying heed to the heavy moral price which it all may have for human society. If religious people have a transcendental anchor for their values, the Divine Command, the secular humanist believes in Man and his moral autonomy, even seeing this as a replacement for religious morality (Image of Man instead of Image of God). If so, he would seem to be even more desperately in need of dualism and all the more so of libertarianism. I do not intend to claim that secular or atheist people are less moral. Certainly not. In my opinion there is

no real difference between these two groups when it comes to their moral conduct. But that only makes the question even more pressing: What is this secular morality based on? As we will see throughout the book, in the determinist-materialist picture of the world, moral conduct is left without any systematic theoretical support. It effectively becomes an irrational tendency left from our religious past, or an evolutionary vestige (many might call it a welcome one). But is there real ethical meaning to moral conduct or is this simply an inherent tendency and nothing else? Such approaches see morality as an anachronism, but continue to use the system of terms and thought it involves (the sanctity of life, the specialness of humanity, moral obligations, the Image of Man, and so on). People don’t see that after the transformation we have undergone and which we are still undergoing, this whole philosophical and terminological structure is built on air. Why does such a correlation exist? Why does a secularist find himself tending towards materialism and determinism, even though as we will see below — such a connection is not at all necessary? One reason is the tendency to oppose religious approaches. If the religious approach says that there is a soul and spirit and free choice, then the secular position finds itself being dragged into the opposite stance, despite the serious cost of such a position. In addition, materialism and determinism are often presented as the necessary results of scientific findings and the scientific method. The arguments and intuitions (which will be presented below) which support libertarianism are put to one side, and sometimes even hidden, so as not to play into the hands of the believers. This is attested to by atheist websites, which are full of arguments (most if not all fallacious) in favor of determinist materialism. Another reason is the common discourse today according to which faith in science (which is “obviously” contrary to religious belief) and the latest scientific findings, requires holding these beliefs. Anyone who wishes to consider himself enlightened and rational must accept this. Thus do the libertarian positions, so intuitive and simple, and in spite of their necessity as the foundation of humanist thought, become the lot of those who believe in God. In my previous book God Plays Dice (Hebrew), I described a similar phenomenon. It also seemed like a struggle between theists and atheists when it came to belief in evolution, with the religious rejecting the findings of the theory of evolution which appeared to contradict their religious faith, and with the secularists uncritically and enthusiastically adopting problematic and entirely not established interpretations of scientific findings, simply because they support the atheist position, and even if they involve a very heavy moral and intellectual cost. In that book I showed that the discourse on evolution is a dialogue of the deaf, in which both sides are blind to each other’s arguments and their significance. Presenting man as solely a biological creature should have given pause also, and perhaps even primarily to those who hold to a secular liberal humanist approach. On the other hand, the discrepancy with scientific findings should cause unease for any intelligent person, even a believing one. I noted there that it is also in the context of evolution that both sides tendentiously hide arguments and findings which do not fit with their ideological paradigm. This is also the case regarding determinism (aside from the concealment, which as far as I know is not in evidence yet, apparently since the issue isn’t a sufficiently “hot” topic). In both cases, believers find themselves facing science, attacking it and trying to refute its scientific conclusions. However, in both cases, this tactic is both unnecessary and incorrect. It is not the findings of science which contradict the approaches they are defending, but only shaky interpretations applied to those findings.

My Basic Approach in This Book Much like the previous book, I start this book with the starting point that reasonable assumptions, and certainly scientific findings, should be accepted regardless of their ideological-theological consequences. There is no room for subordinating what is to what should be, and we must conduct ourselves based on intellectual honesty. It is not just a moral obligation but a cognitive tool of the first order. Dishonest thinking and the dismissal of scientific findings takes us away from the truth. If the findings are valid and reasonable we should adopt them even if they have dramatic moral and/or ideological consequences, simply because they are true. Just as I don’t like those who reject findings about differences between men and women, or between different races, simply because of the moral and ideological dangers inherent in their possible consequences, I would also not expect that this be done on the issue we are discussing here. The question was and remains at the level of interpretation: Are certain philosophical consequences and interpretations a necessary result of the scientific findings? The motivation to examine this question, and to

turn over every stone to try and provide non-standard answers, is what is driven by ideological roots, and there is nothing wrong with this; As long as the ideological motivation which leads us on the search doesn’t slant our judgment or predetermine its results.18 My assumption is that reasonable scientific findings should not be rejected, and if there aren’t very good arguments to do so, we should also not reject reasonable a priori considerations or our fundamental intuitions. But as I will show here, the commonly accepted interpretation of the scientific findings is not always a necessary consequence thereof, and there is therefore room to adapt it to the conclusions of the philosophical-a priori consideration and our initial instincts. Sometimes it is not a zero sum game, and the findings of science can be reconciled with philosophical a priori considerations. Moreover, we will see that scientists are not in a position of authority to establish this interpretation more than anyone else, especially when such interpretations (and not the findings themselves) have such dramatic moral, ideological, and social consequences, and even more so when the interpretations require use of philosophical arguments and considerations, in which the professional authority of the man of science has no relevance. In the previous book, I showed that when it comes to evolution, many scientists deviate from their area of expertise, and often err as a result. Here I will demonstrate this again, this time in the context of physics and neuroscience. In short, my goal in this book is twofold: 1. To cut the connection between the scientific and philosophical levels. The neurosciences do not force us to reach any particular conclusion on the questions of determinism and materialism, and we are free to decide this with philosophical or other tools; 2. To show that the obvious philosophical conclusion, now as it was before the development of neuroscience, is that we have freedom of will and choice. I am placing my summary already at the starting point of our journey as a sign: Do not despair, the rumors of the death of humanism are premature and exaggerated. On the other hand, we also have no need to dismiss scientific research and its findings in order to leave it alive. And you know what? You don’t even need to be a believing man for that…

The Book’s Structure The first two chapters will present the two alternatives — determinism and libertarianism — one against the other. The logic of the discussion throughout the book is of three types: The philosophical discussion includes a priori philosophical arguments in favor of and against determinism (as well as materialism, which we will see is closely related). This part primarily includes chapters three, four, and eleven. I will try to show that in this discussion, libertarianism stands far stronger. The scientific discussion includes discussion of these issues in light of the findings and assumptions of modern empirical science. Most of the chapters are dedicated to this discussion (most of the second and third section). Here I wish to show that the question remains open; at least at our current level of knowledge, modern science has nothing to say on the matter. The diagnostic discussion, the stage of introspection after we see that nothing has been decided empirically. The question is thus left to each and every one of us to decide on their own, based on their judgment. In chapter sixteen, we will propose a number of thought experiments which will allow the reader to form and test their views on the issue. The first part of the book deals with the question of determinism, and its aim is to clearly define the positions as much as possible and place them alongside each other. A great deal of vagueness and many errors in discussions regarding determinism derive from an unclear definition of the various positions, and their removal is the main purpose of the first part. In addition, I also included an a priori discussion (i.e., philosophical, even before going into scientific questions) on the question of determinism. In the second part of the book, I tie the question of determinism to that of materialism, removing a number of failed arguments which connect freedom of will to findings in the fields of physics, and primarily those of chaos theory and quantum theory. At the end of this part, it turns out that the two ideas we face are deterministic materialism vs. interactionist dualism. One must prevail. The same conclusion arises regarding another scientific field, that of complex systems, which we will discuss in the third part. In the third part, the book moves towards working out the differences between these two approaches in light of the latest scientific findings in the field of neuroscience, and those who believe that the matter has already been decided apparently either do not understand what freedom of will is or they do not understand what neuroscience and modern physics say on these issues. The fourth and final part is a summary of the whole book. The sixteenth chapter which starts it is dedicated to a series of thought experiments aimed at helping the reader form their own opinion after having gone through the book. As noted, these are primarily self-diagnostic tools. The final chapter is a

general summary of the book which offers a general outline for deciding the issue of determinism. This is followed by an epilogue which offers a different cogito argument which touches on determinism (the original dealing with materialism), thus closing the circle opened in the prologue. The book ends with an appendix which examines the connection to religious faith, especially Jewish faith, which arises every so often at the margins of the philosophical discussion. The book is interspersed with five asides or “intermezzos,” which expand on traditional Jewish aspects connected to a particular point raised during the discussion. The book is meant for a broad audience, and I therefore added a glossary of terms to make for easier reading. The discussion combines a number of disciplines (mathematics, physics, neuroscience, evolution, psychology, philosophy, Talmud, Jewish thought, and more). I should note that the book is not meant to present the reader with the whole panoply of possibilities and the nuances which arise in the literature regarding the issues under discussion. It is a monograph, which lays out my position before the reader, with the aim of understanding and ironing out the central issues in the various fields which arise here, at least as far as necessary for understanding and justifying my own arguments.

Thanks This book was born out of questions and appeals sent by students learning neuroscience in a number of universities. They found themselves at a loss in the face of the approaches they were presented as well as the stridency and complete confidence displayed when doing so, and more than this — the broad agreement between researchers and lecturers on the issue. Many of them felt that a systematic treatment of the philosophical and ethical issues raised by these findings was lacking, at least as regards the broader, non-expert public. Conversations with them revealed that they were quite often presented with a picture which was baseless and lacking justification, and sometimes even not defined at the conceptual level, yet presented as though they were purely scientific findings. I therefore saw fit to make a contribution and try and clarify the picture and iron out the relationship between its various contexts. My first thanks are therefore to these students. I was already familiar with the philosophical contexts of these issues based on previous study of the philosophical literature, and due to my education in the field of physics I was very well acquainted with the scientific basis needed for discussion of the matter. However, while preparing myself, it became clear to me that without an acquaintance with the current scientific methods and findings in neuroscience, such a discussion cannot be complete and up-to-date. However, as one can see from reading the book, there is no need to be a professional in the field of neuroscience to form a position on the questions involved, but it is nevertheless important to be familiar with the findings and the basic arguments in the field. Quite a bit of my initial knowledge on this subject (primarily as regards the biological and neurological dimensions) is based on the fascinating online lecture series of Hebrew University Professor Idan Segev called “Mashav Moach” [lit. brain wind] and I’d like to take this opportunity to thank him for them. In 2011-12, Hebrew University’s Givat Ram campus opened a course for research students from various disciplines on the subject of “Physics, Brain, and Free Will.” The course was jointly taught by two excellent lecturers: Prof. Haim Sompolinsky, a well-known physicist, the head of the Safra Center for Brain Sciences at Hebrew University, and Prof. Yemima Ben Menachem from Hebrew University’s Philosophy Department. Beyond the lectures themselves, which presented the issue very clearly from various angles, and beyond the sources provided for study and further learning, I found both of them to be patient teachers, answering inquiries and questions, and I therefore allowed myself to express different views than they did and even critique the direction, structure, and assumptions of the course itself. I bothered them quite a bit, and I was usually met with a great deal of patience, the result being that I learned much from both of them. I also made use of the course’s TA Dr. Ariel Furstenberg, whom I also thank. Special thanks also to Ariel for very useful comments sent to me after reading an earlier version of the book. To prevent any misunderstandings, I should note that our conversations showed that both Sompolinsky and Ben Menachem (each for their own reasons) do not agree with the main conclusions of this book. My appreciation for their assistance should therefore not be seen as either their agreement with my opinions nor my agreement with theirs. It is a pleasure for me to thank Ami Berholtz, who accompanied me and the manuscript since it made it to the Yedi’ot publishing house, in Hebrew, as well as the entire editorial team headed by Dov Eichenwald for their professional and dedicated efforts. I will end this preface with special thanks to the wonderful

editor of the book, Matat Eshet, who combined professionalism with mercy and pity (sometimes for me, sometimes for the readers) and improved the book immeasurably. 11 In the sense of the word “liberty.” This word is commonly used in the political context; a libertarian party or outlook espouses the liberty of the individual to do as he wishes without limit (as much as possible). In this sense, libertarianism is a system of values, usually connected to the right in its socioeconomic sense, and it deals with what should be rather than what is. By contrast, throughout this book I will use this term in its philosophical sense: An approach which sees human decisions as deriving freely from his will and not determined by any other circumstances (more on this below). The significance of this libertarianism refers not to the “ought” in the ethical sense but rather a factual description of “what is” in reality. 12 The book is based on his Gifford Lectures series given at Edenborough in 2009. Michael S. Gazzaniga, Who’s in Charge? Free Will and the Science of the Brain, Harper Collins, NY, 2009. 13 Shimon Marom, “The Eternal Temptation” (Hebrew), Odessiya 6, January, 2010. http://odyssey.org.il/209225 14 In my book God Plays Dice (Hebrew; pp. 216-217), I brought an example of a similar fallacy in the field of evolution, where the judge, who lacked scientific know-how, had to decide a scientific question under dispute between scholars (the question of gaps in the evolutionary chain). Regarding the ideological biases of scientists, cf. appendices e-f, ibid. 15 Another clear example of being “wowed by science” is the phenomenon of Munchausen syndrome by proxy, when a British doctor unaware of (not particularly complex) statistical nuances led to the imprisonment of women on the charge of murdering their children, who in fact died from crib death. See my article on the subject, “The Statistical Reliability of Tests for Rare Phenomena — The Fallacy of Representation in Halacha,” (Hebrew) Asia 91-92, 2012. 16 Towards the end of his book (pp. 198-199), Gazzaniga brings a number of reasons why such scans should not be admissible in court at the present time. However, at the same time, he says that this state of affairs is temporary, the result of an information gap which will be filled in the future. As I will show throughout the book, I only partially agree with this. Brain scans may be able to show us a damaged brain which cannot bear responsibility for its actions, but will never give us the full picture regarding human activity. It is to this issue, among others, that this book is dedicated. 17 Iri Rickin, Haaretz, 27.7.2011. 18 The philosophy of science tends to distinguish between the context of discovery and the context of justification. The context of discovery is the manner in which the scientist arrived at the theory he is presenting. This context could be entirely subjective (such as discovery in a dream or according to the numerology of children’s books), and its nature is not relevant for the discussion of the issue itself. That discussion should be only at the level of the context of justification, that is the question of whether the proposed theory meets the standards of empirical tests and reasonableness.

Part One Defining Terms: Freedom, Liberty, and Causality In the first section of the book, we will present the two fundamental pictures of human beings — libertarian and deterministic — opposite each other. Much of the lack of clarity in discussions on these issues is due to debaters on both sides failing to properly define and sharpen their understanding of these approaches before actually dealing with them, a failure which leads to errors in the arguments of each side. We will encounter quite a few of these errors later on. In the first chapter, we will draw an initial outline of the debate on free will and define some of the basic terms of discussion. In the second chapter, we will present both the libertarian and deterministic approaches in further detail. There we will see that a central mistake in this debate is an imprecise division between the possible mechanisms: randomness as opposed to determinism. My goal in this chapter is to demonstrate that free will belongs to neither framework. In the third chapter, we will examine the possible consequences of the principled debate on determinism, where we will see that they are not so common and far from being trivial. In the fourth chapter, we will examine these questions via a priori philosophical arguments (discussion of the empirical science is reserved for Section Three, after the philosophical issues have been fleshed out). The fifth chapter involves a detailed examination of the concepts of causality, which is very important to understand for the purposes of our discussion, as we will see below. The sixth chapter will examine the relationship between causality and free will, and the seventh chapter, which completes this section, will offer a mathematical model to help sharpen the three mechanisms examined here and distinguish between free choice, randomness, and determinism.

Chapter One THE DEBATE ON FREE WILL: AN OUTLINE The difference between a debate and quarrel is that in the former you don’t agree with what’s being said, and in the latter — with who’s saying it. – E.B. Worthington In this chapter, I will present the pair of terms our book will deal with throughout: libertarianism, which espouses freedom of will (or freedom of choice; despite the imprecision, I will not distinguish between the two here), and determinism — which rejects it. We will begin here with just a first impression, but the fundamental distinctions pointed out in this chapter will also be important later on.

The Intuition of Free Will: Libertarianism in Brief Semone Grossman is a Holocaust survivor, the wealthy owner of the GGMC company (Grossman Garage Management Corporation, or as his employees like to call it: Go Get My Car), which owns and operates about fifty parking garages in New York. Grossman described his motivation to enter this field as follows:19 But you understand, the most important thing for me in life is my freedom. That’s why I bought more and more parking lots. If I travelled to the same garage every day, I’d go insane. I wanted to have the option of choice. In the concentration camp, you didn’t have permission to breathe without asking the Nazis. If you wanted to go to the bathroom, you’d throw up your hands and get two minutes to go and then go back to work. What drove me was the desire for complete freedom. I wanted many locations in the city for that. To decide for myself. Here, Grossman is describing the importance of freedom for him, explaining how the lack of freedom was what disturbed him most about the concentration camp. Beyond all this, however, he is implicitly stating a very fundamental assumption: Lacking these artificial and evil constraints, everyone possesses such freedom. It would seem that almost every human being lives with the feeling that he can choose what to do. A man comes to the conclusion that he wants to learn history at a course being held in a particular place, so he gets up and goes there to study. Another decides that he actually doesn’t want to go there and take the course, so he doesn’t go. A man makes smaller or bigger decisions — such as lifting his hand, going to sleep, making a phone call, asking forgiveness, committing murder, giving charity, climbing Mount Everest, becoming religious or secular, and even becoming a determinist or a libertarian — and every one of these decisions is accompanied by a feeling of freedom and autonomy. We need to first define this feeling of freedom. Its most basic foundation is the very existence of options open to individuals to choose from. When someone chooses to take a course in history, and he feels he decided to do so of his own free will, his implicit assumption is that the two options were equally before him. He could have gone and he could have not gone. Only in this case can we say that his decision to go (or not) was truly free. In saying that the two possibilities were open before him, I mean this at least on three levels: (a) the two options exist in the objective world; (b) not only were the two options open in and of themselves, but they were open to him. That is, he could realize either one of them, if he only decided to do so; (c) If there was any external factor which compelled him, and he himself was passive, then even if he himself was unaware of it but thought that he was doing it of his own free will (and also really wanted it), then that is not free choice. He himself needs to be the agent causing the step he takes, him and no-one else.20 Intuitively, it’s clear to us that if any one of these conditions is not in force, the “decision” taken by the individual is not free. In such cases it is incorrect to say that he “chose” to do what he did. To continue the example of the choice to take a university course: If there is no history course in that location, he has no option to go study history there. Therefore, if he did not take the course, it was not due to his free choice. And even if there was a course, but he himself could not make it for whatever reason, then once again his decision not to go was not a result of free choice. Finally, even if there is a course, and he could go or not, but someone hypnotized him (or threatened him) and made him not take the course — we would also not

treat such a decision as “free.” This is true even if he subjectively felt that he decided it himself and in an entirely free manner (as he was hypnotized to think so). This feeling is an illusion. This is the essence of the libertarian approach. Below we will see that the compatibilists question part of it.

Who is the “Other Factor”? It’s very important to discuss the identity of the other factor appearing in condition (c) above. There are two common errors we must first remove on this issue before proceeding. The first error is compatibilism, which defines a situation, in which our internal condition entirely determines our decision, as free choice. Here, the assumption is that internal factors are not considered “others” — that is, only external influences can be considered deterministic constraints. The second error is biological indeterminism, which assumes precisely the opposite: Only internal factors are considered “other,” while the influence of external factors is not considered determinism. We will now examine these two arguments in greater detail.

Compatibilism This approach originated in the theories of Hobbes, Hume, and Mill, who claimed that there is no contradiction between free will and determinism.21 It is very common and popular in our time, as it ostensibly reconciles the contradiction between free will and causal determinism. I will present this approach based on a discussion in the Babylonian Talmud, where we find a very interesting discussion regarding the verse in the book of Exodus (12:36): “The Lord had made the Egyptians favorably disposed toward the people, and they gave them what they asked for; so they used the Egyptians.” God made Israel find favor in the eyes of the Egyptians, and the Egyptians gave them vessels and property, until they ended up bereft of property. In the Talmud (Brachot 9b), a dispute is brought regarding the significance of this action: And they let them have what they asked. R. Ammi says: This teaches that they let them have it against their will. Some say, against the will of the Egyptians, and some say, against the will of the Israelites. Those that say ‘against the will of the Egyptians’ cite the verse: And she that tarrieth at home divideth the spoil. Those that say: ‘against the will of the Israelites’, say it was because of the burden [of carrying it]. Rabbi Ami argues that this lending of property to the People of Israel was done against their will, but it does not specify whose will. The sages of the Talmud were divided in interpreting his words: One opinion believes that this was against the will of the Egyptians, and a second opinion is that it was against the will of the Jews. What is the first opinion, according to which it was done against the will of the Egyptians, based on if it explicitly says in the Bible that the People of Israel found favor in their eyes and that they gave them their property out of good will? It turns out that the reason for this finding of favor was thanks to a kind of divine hypnosis. The verse starts “The Lord had made the Egyptians favorably disposed toward the people,” — that is, he changed the Egyptians’ thinking patterns. He caused them to think they were acting willingly and based on their positive attitude towards Israel, but the truth is they really didn’t like them, and therefore it is considered as if it was done against the will of the Egyptians. The disputing opinion apparently did not see such a situation as being against the will of the doers; after all, that’s what they wanted to do and decided to do. If so, the second opinion is left with the approach that it was done against the will of the People of Israel, who did not want this because now they needed to carry this heavy burden in the desert (“because of the burden”). This dispute very clearly expresses the idea of compatibilism. The compatibilist approach argues that the contradiction we see between determinism and free will derives from a conceptual error. These two ideas, free will and determinism, are in fact compatible. The argument of compatibilism is that free choice means choice done by us and by our will, and without external constraints. But what is that “us”? What is that “will” which makes this decision? Their argument is that even if our internal psychological circumstances “force” us to do something, then that is free choice, as it was done by our will. We do exactly what we “want,” as those internal circumstances are us, they are our “will.” In the terminology we used above, the internal circumstances are not an “other factor.” There is nothing else here which forces us to act. The compatibilist is effectively arguing that there is no freedom here in the accepted intuitive sense (in

the wake of the famous distinction of British thinker Sir Isaiah Berlin, sometimes such a situation is described in terms of “negative freedom,” as opposed to libertarianism which believes in freedom in the strong and positive sense), but there is will (and maybe choice). It also argues that even if the options we choose are dictated in advance by our physical and mental state, we are still the ones who choose, because an internal factor is not “other” or foreign and there is therefore no other factor besides ourselves which influences our decisions. The totality of psychological and physical circumstances which brings us to this choice is what executes our will, and therefore this is our choice. To round out the picture, we will add that the compatibilist does not suffice with the semantic argument — that is that such a mechanism is also worthy of the term “choice.” He goes on to argue that one can impose moral responsibility on a man who acts in such a matter, as he sees this as action based on choice. This is no longer semantics, but a weighty ethical argument. To understand this better, we will present compatibilism opposite the normal libertarian approach. According to the normal approach of free will (which the compatibilist would call “naïve”), our desire is born ex nihilo. There are no previous causes which determine it, neither external nor internal. From this aspect, we are dealing with a highly irregular ability in relation to our known physical world (so long as we are dealing with classical physics; issues of quantum physics will be discussed later in the book). Every physical process ties cause with effect, and therefore physics is nothing more than various chains of cause and effect dependent on one another. They all started with the Big Bang, and have since continued to revolve around each other. The only exception, per the libertarians, is human will. That will was born without any cause bringing it about, and from here on out, a new physical chain starts which joins the other chains of causation in the physical world. For instance, a man decides to raise his hand, and as a result the air moves, a noise is heard, another man hears it and decides to react in some way, and so the chain rolls ever onward.

The Libertarian Picture

By contrast, the compatibilist sees will as also being a physical event. In the chain started by the Big Bang we reach event 1578 which caused 1579, which is what we call the desire of so-and-so. This desire is a psychological state which is nothing more than the consequence of physical factors which preceded it and created it, and it in turn continues the causal chain and creates other physical effects (lifting the hand, &c). According to the compatibilist, desire is a link in the regular physical causal chain, and no more. This chain simply goes through it, like it goes through any other link. Our feeling that we want, says the compatibilist, is not an illusion; it’s actually true: We do want to do such and such. Except this feeling is not spontaneously created (without any cause bringing it about), as the libertarian believes — that is as a result of our autonomous choice — but for other physical reasons which preceded it, like any physical and mental event.

The Compatibilist Picture

The compatibilist argues that our will is not free, as it is not autonomous but rather acted upon by other physical forces. However, he also argues that one can impose moral responsibility on the man who acts in this manner. Thus does the compatibilist change the meaning of the term “freedom” under discussion. If so, it’s pretty clear that compatibilism is nothing more than a synonym for determinism. It contains nothing that goes beyond ordinary determinism, aside from a semantic change in the term “freedom”: What the determinist calls compulsion (by internal factors), the compatibilist calls choice or will. The main problem with this position is that this is an attempt to solve a real philosophical problem by simply changing the terms. In what I will call my quartet from here on out,22 and primarily my first book Two Carts and a Hot Air Balloon (Hebrew), I showed that this analytical way of thinking is often nothing more than wordplay, and one cannot find a real solution to a problem by going this route. At most, one can discover errors in the way the problem was formulated and demonstrate that the problem is illusory to begin with, but one cannot actually solve it in this manner. There is of course another important difference, and that is the ethical argument raised by the compatibilist, according to which such a person can be subject to moral responsibility. The necessary conclusion is that the compatibilist is no more than a determinist who is (justifiably) shocked by the ethical consequences of his conclusions and who is trying to fight them, but as we will see, he is not all that successful at it. One way of understanding why this idea is strange is to think of a kleptomaniac or a drug addict who can’t go clean. According to the compatibilist, such a person, at least if unaware of his disease, is a man who freely chooses to steal or take drugs, as he is only doing what he himself wishes to do without any external constraint. He may be built differently than most people, but everyone has their own uniqueness and physical form. So long as his actions are based on his internal motives, he is considered to have acted freely and is responsible for his actions. Should we impose moral responsibility on the kleptomaniac? The compatibilist will try to say yes, but the justice system tends not to impose criminal responsibility on a compulsive person (we’re talking someone who’s entirely compulsive, without the ability to overcome his urges), and it seems to me that this is the simple ethical and legal intuition of most people. Similarly, in the Talmudic example above, the Egyptians acted due to a mental change of their approach. They would also be considered by the compatibilist to have acted of their own free will and not against it. After all, they — that is, the totality of their psyche at the moment — really and truly wanted it. If so, their decision was truly “free.” This absurdity is very similar to what we see when it comes to kleptomania. In fact, we can in the same way impose moral and legal responsibility on animals (and maybe even on cars), as they are also operated by their nature, and therefore effectively do what they “want.” I would not address all these arguments of compatibilism, which prima facie seem utterly unreasonable, were this approach not so widespread in contemporary philosophy, and were it not for so many important thinkers having poured lakes of ink on the issue. Compatibilism is considered by many to be a third approach in the debate on determinism, and some have even gone so far as to define determinism and libertarianism as two sub-categories of incompatabalism, as though it was the main approach and they are the alternatives. However, as we saw, compatibilism is little more than frustrated determinism and nothing more, and it should certainly not be seen as a third approach.23 It’s important to understand that compatibilism leaves us with two unsolved problems: (1) the decision I make is not really a free decision. True, there is no other factor causing it besides me, but I am not causing it either. In fact, it is not a decision leading to an action, but an action accompanied by an illusory feeling of

a decision. (2) Per this system, the first condition of the three we noted above does not obtain, according to which the two options exist for us. If our overall physical and mental state dictated the decision, then we actually only have one choice before us. Still, per its argument, we bear responsibility for that decision; even a man who chose to take the only path he has, so long as he did it “willingly,” he is considered as having acted freely. The conclusion so far is that contrary to the compatibilist approach, if that same “other factor” which forces us to act is part of us, that is some internal factor, it’s still another factor. Choice exists only when we could have chosen both paths and we freely chose one of them, ex nihilo (i.e., not due to previous causes, internal or external). Only then do we bear moral and legal responsibility for choosing that path. In any event, compatibilism is the first error regarding the identity of the “other factor.” We will now move on to describe the second error, which comes from the entirely opposite direction.

Biological Indeterminism, or: Internal Determinism Surprisingly, the life sciences world has a definition which is the exact opposite of the previous one, and which is also seen as a solution to the problem of the clash between determinism and free choice. According to this approach, determinism is only internal. Thus, if our decision is not entirely dictated by our internal physical and mental state, then we are not dealing with determinism. This is true even if there are external factors which entirely dictate our conduct. In other words, if internal circumstances do not dictate our behavior but rather other ones, external to us, then the picture is not deterministic. To sharpen the picture, I will present the words of Prof. Aharon Ciechanover from the Haifa Technion,24 a biochemist and doctor by profession and winner of the Nobel Prize in Chemistry in 2004. As background to his statement, Ciechanover presents his materialist worldview: As a scientist, I tend to think that everything is chemistry, and that when I pass from this world, I will be here no longer. In the cemetery I will be worm food, and everything I carry with me will disappear. What I wrote and recorded will remain, of course, the films that are left, and the ideas I taught, but I myself will no longer be here, even to interpret them. My spirit will disappear, my creation will melt, my loves will vanish, my emotions and my intelligence — everything that makes me a human will be no longer. This is perhaps the testimony to it all being chemistry. The moment the cells and molecules turn into worm food, there is only the past, which is metaphorical; the ability to create freezes at the moment of death. The mission to continue and develop ideas on the basis of those which exist was passed on at that moment into the hands of those who are alive, into the hands of their chemical reactions. After Ciechanover finished handing assignments to future chemical reactions, the interview went on to address determinism. Ciechanover was asked: One of the longest debates in philosophy and science deals with the question of determinism — that is, to what extent is everything set in advance. The assumption is that if the world is deterministic, then there is no free will, and as soon as we identify the principles — we can know any future action in advance. For instance, when we understand the activity of genes, we will be able to predict all our traits and actions. To this, he replied: The debate on determinism is ancient and the best philosophers and thinkers dealt with it: From Ecclesiastes who said, ‘What was is what will be and what is done is what will be done,’ to Aristotelian Greek philosophy, until today. The uncertainty principle in quantum theory was supposed to remove determinism from this world at least in its physical sense, but that’s not what happened. Who am I, therefore, to state an opinion on the matter. I will only note that one of the fascinating phenomena which has been discovered in the last few decades is the plasticity of the brain — the changes which occur in it during its activity, and its ability to change in accordance with what we undergo and experience. This is great news for the opponents of determinism. The environment we grow up in has a great influence on our formation as people. If you take a child and teach him how to be a man, he will act that way his whole life. If we take his twin brother and raise him in a mussar yeshivah (Jewish religious place of learning with a special focus on improving virtue and moral conduct), he will be an entirely different person than his brother. Culture,

education, and surrounding society — all these have a great influence on human behavior and its shape. We are certainly distant from determinism and the possibility that we will reach the stage where we can decode and predict what is expected in the future. I want to understand the plastic mechanism of the brain, the components which determine and influence our behavior. This is not a deterministic process, even though it is operated by a mechanism which is primarily physical and chemical. We are trying to understand the mechanisms, but even if we understand them, that doesn’t mean determinism. After all, we change unceasingly and adapt ourselves to changes, and therefore, even if this mechanism becomes clear to us, that doesn’t mean that we will have the ability to predict. Ciechanover is arguing here that determinism is a situation in which internal circumstances control us and dictate our actions. By contrast, if what dictates our actions and our decisions (and our future) is not an internal but an external factor, such as education and environment, then he does not consider it to be a deterministic picture. For this reason, the plasticity of the brain (that is, the ability of the brain to change due to external influences) is to him “great news for the opponents of determinism.” It’s very surprising to see a similar hypothesis espoused by one of the great neuroscientists of our time and one of the founders of the field, Michal Gazzaniga. In the sixth and final chapter of his book, Gazzaniga also argues that it’s true that every individual and every brain acts deterministically (he is a physicalist), but free will should be seen in a social context. A criminal also does not act under the eyes of a police officer, he argues, and therefore he is clearly responsible for his actions. This, in spite of the fact that the police officer’s influence on the criminal’s brain is deterministic. The same is true of other social influences, which may change our decisions. Since society influences us, Gazzaniga argues, we are responsible for our actions. This is precisely the meaning of the freedom to choose, he argues. Once again we see a man of science identifying determinism with internal influences, arguing that the moment there is also external influence, even if it is forced on us, and even if our actions are not our own, this is no longer determinism to him. But it’s a mistake to see the plasticity of the brain as demonstrating freedom of choice or as the basis for moral responsibility. A world in which everything is physics and chemistry, even if there are external influences (which are also all physics and chemistry) which can influence our brain, is perfectly deterministic. We may not die as we were born, and our situation in the end is not a necessary result of our situation in the beginning, as the brain is plastic and external circumstances can therefore also change it, but the manner of this change is also entirely deterministic. We don’t decide it, external circumstances do. There is therefore no significance to discussions of morality. At most, determinism lies at the global level: The entire universe necessarily takes a path towards a future which has been pre-determined. Every detail within it is affected by the environment, and therefore observation of that individual is insufficient to predict his future, but this is merely a technical problem. If we observe the whole environment, we should in principle be able to predict the future of every specific element. Thus, the plasticity of the brain is a scientific matter which has no consequence for the ethical discussion of determinism. The argument that education has value since it can affect the brain of the studying child is nothing more than a doubling down on determinism: First, we the educators are determined according to the circumstances in which we operated and in which we operate, as well as by our biochemistry. Subsequently, the education we give the child does not operate via his cognition and decision-making process, but is instead a reshaping of his plastic brain. Still, we should remember that all the while we ourselves are nothing more than the product of previous shaping by our teachers and environment. All of it was of course sculpted by the greatest sculptor of them all: Not Leonardo da Vinci, but the Big Bang. From there on out, everything was the result of entirely deterministic processes which we and our predecessors underwent. Thus, philosophically, such a worldview is entirely deterministic, as the circumstances we are in bear part of the “blame” for the decisions we made — this in addition to our internal physics and chemistry. Everything bears a part if this blame, except for us-ourselves (it would seem that such a concept as “usourselves” does not exist with Ciechanover). If so, it is still clear that there is no room for imposing moral responsibility on the consequences deriving from these decisions. Why then do Ciechanover and Gazzaniga see news here for opponents of determinism? It’s hard to say. This is just one example of prominent scientists demonstrating their lack of philosophical understanding. Ciechanover’s error derives from the use of the same term in two different senses, which are used in two different contexts. There is a great deal of interest in the life sciences in the question of whether or not our internal situation entirely determines our decisions, as well as how much external influences can change

our phenotype and our brain. The field which investigates influences on the phenotype not caused by changes of the genotype is called epigenetics. The parallel field in neuroscience, which investigates changes in the brain due to environmental influences, is the study of the plasticity of the brain. Therefore, in the biochemical context and in neuroscience, they tend to describe a situation in which only the genotype influences the phenotype as “determinism.” If external influences have a part in this influence, it is called “indeterminism.” The problem is that Ciechanover, like many others, mistakenly uses this scientific term, which is defined and serves to understand scientific phenomena in biochemistry and neuroscience in a useful manner, and applies it to the philosophical and ethical realm. In that context, when external influences affect us, that is a clear case of determinism. There is another error by Ciechanover in that interview, which will also come up later in the book, and I will therefore discuss it here only in brief. He consistently confuses the ability to predict matters with determinism. For instance he was asked: The conclusion from what you said is still a sort of determinism. When we know all the details of the mechanism and how it operates, we will know what will happen in the future. To this he replies: I think that would be impossible. If we know everything going on around us, would we really be able to pull out a computerized prediction? I can’t say, but I don’t presently see a situation in which we could really know everything, primarily because of the reverse pyramid I described [that every additional piece of information raises new questions. M.A.]… Ciechanover here identifies determinism with predictability, which sounds reasonable. But in addition (as in the previous quote) he also identifies the inability to predict matters (and in this case simply because of lack of knowledge) with indeterminism, and this is already a conceptual and philosophical error. The term determinism does serve to describe predictability in the scientific context (primarily in physics). However, this is also inaccurate, as sometimes even deterministic systems are unpredictable (see, e.g., chapter nine, where we will discuss chaos). Thus, the fact that we cannot predict something does not mean that the system is indeterministic. In the philosophical context, a deterministic system is one whose future state is entirely determined by its present one, even if we can’t know or predict what that future is. Accordingly, indeterminism in the philosophical sense is a situation where not only can’t we predict the future as observers, but the present system does not predetermine that future. It’s not just our lack of knowledge, fundamental or coincidental, but the simple impossibility of the thing in reality itself. Let’s imagine a man whose decisions are determined towards a single value by his physical-biological state in the present, but which cannot be predicted by anyone else, and for our purposes we will assume that the world at large won’t be able to, either. Is such an individual acting out of free choice? Of course not. Our knowledge of reality is simply incomplete, and it does not in this case fit reality itself. After all, it is not reasonable to impose moral responsibility on someone for his actions based on the amount of scientific knowledge in the world at that moment in time. Knowledge can serve as an explanation for errors of judgment done by him, but not the very fact of moral responsibility or non-responsibility for his actions. For us to attribute free will to an action, the indeterminacy of the future action in and of itself is required. It’s not enough that no-one can predict it. Now we will examine the situation from the other end. Does an action done randomly express free will? That is, an action which is not at all determined by the circumstances of the present moment. It just happens, for no reason at all. The “decision” came up into my consciousness without any prior judgment, thought, or pondering on the matter. It’s just there and that’s it. Here, too, there is a lack of ability to predict, as well as true indeterminacy, including regarding reality itself. Still, it is clear that this picture also does not reflect free choice or will. We will not impose moral responsibility on someone who acts in such a way, as free will assumes something else: That we with our decisions and judgment now determine the future situation, and not just that the action is not determined by the present. Thus, it is clear that free will contains the assumption that we cannot predict the future, but we cannot forget that there are other necessary elements here without which we cannot speak of free will. The conclusion is that to arrive at a philosophically non-deterministic picture, it is not enough that we have no prediction capacity. This is a very common error among scientists who deal with these questions, as the accepted scientific terminology does indeed identify lack of predictability with indeterminism. In science,

this term is beneficial and correct. In philosophy, this use of the term is no more than a mistake deriving from lack of philosophical skill.

Interim Summary So far, we have reviewed two opposite errors about determinism. The first is compatibilism, which defines an action driven by internal causes as an action for which the individual has responsibility. Per this approach, determinism is only in effect when the action is the result of the influence of external causes. By contrast, various scientists, due to the mistaken scientific terminology, define determinism in precisely the opposite manner: internal influence. They do not consider the influence of external factors to be deterministic. Bottom line, there is no real difference between those who treat external influences as the redeemer of our free will (the plasticity of the brain), and those who see internal forces as the basis for our liberty (compatibilism). No-one disputes that the two factors affect our behavior (and we will explain the significance of that next chapter), but at the same time, those who understand what free will is also understand that these two factors combined are what constitute deterministic behavior, and they certainly do not save us from that conundrum. Libertarianism must include another component, one which does not dictate our behavior in a predetermined manner. Something beyond both internal and external factors — what we know as will. According to most libertarian approaches, will is activated in a non-physical framework, which leads us to the existence of the spirit or soul. One could call the entity which activates will I, but that’s not enough. After all, why isn’t I one of the internal factors, or perhaps the totality thereof? What does it have that they don’t? We will examine this later in the book. The connection between dualism (a belief in the existence of both the spiritual and material) and libertarianism, which is not obvious, will also be discussed below. For now, the conclusion is that free will means that we can in principle and at this moment arrive at different decisions, and they are all possible. None of them are determined in a unidirectional manner by “another factor” — not based on the present internal biochemical circumstances and not by any external factors. On the other hand, the libertarian certainly believes that our decisions are not made in an entirely arbitrary manner, as if that is the case, then we are not dealing with determinism but also not with free choice. The concept of judgment is essential to the definition of the freedom of will and choice, and we will deal with this in greater depth below.

Additional Limits on Freedom in the Libertarian Approach Even the libertarian who believes that our decisions are our own and not determined by foreign causes, does not mean to say that everything is in our hands or that there is nothing we can’t do. Even in the libertarian lexicon, free choice does not mean omnipotence. A man can decide to climb Mount Everest and fail. Does that mean that he has no free will or that he does not choose to do his actions? Another man decides to murder his friend because of an irrepressible, momentary urge. We will not treat such actions (or the harm caused by inaction) as deriving from free choice, but on the other hand, there is no doubt that the libertarian will admit that there are actions like this. Their existence does not undermine his fundamental approach. He claims that human beings usually make decisions and carry them out freely, but he does not claim that this is what always happens. Another category of decisions which are not based on free choice involves situations where a man decides to climb the Everest and doesn’t even try to do so. When the reason for this avoidance derives from internal causes, and not external and objective ones, we reach the problem known in philosophy as “weakness of the will.” Such situations are familiar to many of us. For instance, someone decides to go on a diet, and he is in principle capable of seeing it through, but in practice does not even try to do so, or he tries and doesn’t succeed (apparently he did not try hard enough). What accompanies him is a feeling of missed opportunity and failure. If he was entirely incapable of doing it, then there would obviously be no failure. We are talking about situations in which he wants to and to the best of his understanding can actually achieve it in principle, yet he still fails in practice to realize his vision. The philosophical literature deals quite a bit with the analysis of these situations, and opinions on the matter are split. Some believe that we are dealing with a fundamental inability (he didn’t do it because he really couldn’t). Others believe that the person didn’t actually want it at all, or didn’t decide it (he just thinks he decided, but that’s an illusion). A third approach, which seems more appropriate for our own intuition but which is also harder to explain and justify, argues that this was a real decision which he could have carried out, yet he still failed because of the weakness of his will. We will return to the question of the weakness of the will in the next chapter.

So, the libertarian also agrees that there are quite a few examples of situations in which the individual does not truly have free choice. In such situations, he is forced to do what he does, whether by external circumstances or by his internal nature. Does the existence of such situations refute libertarianism? Certainly not. According to the libertarian approach, these situations are possible, but they are exceptions. There may be situations in which a man does not freely choose, and there may be various reasons for this, some of which will be detailed below. But the argument of the libertarian is that there are quite a few situations in which individuals can make decisions of their own free will. He does not claim that people always freely choose, nor does he argue that he is always right in assuming that he chose freely even if he feels he did. The libertarian’s argument is weaker: sometimes (and perhaps usually), there are situations in which people choose freely.

An Opposite Look at the Intuition of Free Will: Determinism on One Foot What’s new in this approach? If every one of us feels that they decide and choose freely, why do we even need a definition of libertarianism? Does anyone dispute the minor claim made at the end of the previous section — that sometimes people choose freely? It turns out that quite a few people dispute it, and the position they hold is called determinism. Determinists argue that all actions of an individual are forced on him, either by his internal nature or by external circumstances. Even in cases where an individual feels that they decided and chose freely, the determinist will argue that this is an illusion. Determinists like to use the following common example: If a rock thrown into the air had consciousness, it may be that it felt that it decided to fly in the direction it was thrown in. Does that mean its feeling is true? Is the fact that it feels that its flight is the result of its will and its choice is sufficiently strong proof that this is indeed the case? If so, the fact that we have a feeling of free will, says the determinist, is not proof that we actually do possess such freedom. He would argue that this feeling is also something forced on us. Ostensibly, the argument of the determinist sounds reasonable, even trivial. The feeling of freedom is not necessarily proof of its existence. In a state of fata morgana, an individual walking through a desert can feel that there is a spring up ahead, or an oasis, and be wrong. People fear the dark and that they are being followed by hostile forces (once called demons) which will harm them. Does that mean that every dark place has such forces? It would seem not. Sometimes Yaakov feels that Rachel loves him, but he’s wrong. Moreover, sometimes he feels that he loves her, and here, too, he may be wrong. If so, it is clear that people sometimes err in their feelings and beliefs, and it is not right to assume that every human feeling necessarily reflects reality as it is. But the determinist does not suffice with undermining the credibility of the feeling; if he did, there would be no real difference between determinism and simple skepticism. The determinist’s argument is stronger than that: Not only might the feeling of freedom be an illusion, it is indeed an illusion. In contrast to the minor libertarian position presented above, according to which there are situations in which the individual chooses freely (although not in all of them), the determinist argues that no such situation exists. He rejects the existence of the human ability we are used to calling will, or at least free will. By his lights, any feeling of freedom is an illusion; in each and every situation in which someone acts or does not act, he is driven by either internal or external forces. Thus, there is a certain asymmetry between the two sides of the debate. The libertarian argues that sometimes a person can act freely (though there are times when he doesn’t). The determinist, by contrast, argues that all of our actions are forced on us, without any exceptions.

The Origin of the Illusion of Freedom Here, of course, we once again face the fundamental question which started the discussion: The argument that feelings do not necessarily reflect reality is obvious and hard to dispute. But the next step taken by the determinist, the assumption that the specific feelings of freedom are always mistaken, is an argument that requires additional justification and grounding. Why does he arrive at this far-reaching conclusion? After all, he agrees that in other areas, we have feelings which do reflect what is going on in the world. When we see a wall in front of us, the basic assumption of both the determinist and the libertarian is that we are looking at an actual wall. When we feel pain, even the determinist agrees that usually (although not always) this is an expression of a physical event which occurred to us (a wound, a blow, and so forth). So why is it that he considers the feeling of freedom to be the exception? Why is it here that he chooses to state that it is always illusory, and that it necessarily does not reflect true reality?

Moreover, pain can also be illusory (like phantom feelings). Sometimes it deceives us. Will we cease to believe what our own eyes see or our nerves feel? Of course not. Why, then, should we pay feelings of freedom any less heed? I once heard a possible deterministic explanation for the formation of this “illusion.”25 According to this proposal, the feeling that our will is autonomous and free (and not influenced by other factors) is a simple result of how it is influenced. Since the factors which influence will are outside of us, our consciousness takes no note of them. It is only aware of the chain of causation which lies inside of it. Thus, our consciousness is only exposed to the part of the causal chain which begins the moment will is created within consciousness. The previous stages, which awaken this will (see above in the description of the compatibilist picture), are outside of us, and therefore our consciousness cannot notice them. In other words, will is the beginning of the psychological process of our action, and it therefore appears to be free, that is created ex nihilo, as though nothing came before it. This proposal appears very problematic. There are quite a few things which occur by us due to some external cause, yet it is still clear to us that they don’t happen by themselves but due to that cause. For instance, when we are hit, we feel pain. Are we experiencing pain as though it emerged ex nihilo? Of course not. We are aware that the cause of this feeling is external: the blow. Sometimes we are saddened because of some event which happened outside us. Are we thus feeling an illusion that the sadness occurred spontaneously and autonomously? Here, too, the answer is in the negative. The determinist argues that when we experience a provocation, our need to respond is a compelled consequence of that provocation (in our given psychological state). According to him, our feeling that the rage and the will to respond is our decision is an illusion. But why assume this? We feel the external cause which caused this process quite well. The provocation has not disappeared from our eyes, yet still it is clear to us in many cases that the decision to react was our own. Why doubt this? In all these cases, we can certainly attribute our reaction to an external cause. We distinguish it clearly, even though it is outside us. Therefore, it is not clear why the context of will should be any different or create the illusion as though there is no external factor and that it awakens ex nihilo. This problem is solved on its own, indeed does not even arise, if we adopt the approach that there really is no other factor involved when our will awakens. If libertarianism — which argues that will really does emerge without cause — is correct, then it is obvious that this is precisely what we are experiencing. It’s important for me to stress that no libertarian would dispute that we are influenced by our environment, but his argument is that we are speaking of influence and not determination by an external factor (more on this next chapter). So, what are the arguments which bring thinkers and scientists to hold to a deterministic position, which is entirely at odds with our most fundamental experiences? This isn’t just a psychological question. We aren’t dealing here with the motives of people but their arguments. Are these arguments necessarily true, or perhaps they are only likely, or even not that? In light of the arguments we will present and examine here, we must also turn around and ask the other side: What are the arguments which lead other thinkers and scientists to hold a libertarian position? We will also want to find out for ourselves the consequences of holding to each one of these positions (the question of moral responsibility, and so on)? This well-worn question is far more complex and delicate than it might first appear. We will nevertheless discuss it later in the book.

A Note on the Attitude to Chronology Before we go on to the next chapter, we should make a brief comment on the attitude of each one of these attitudes to the timeline. The determinist treats the past in the same way he treats the future. Both are out of his hands; the only question is where I am located on that line from past to future. Everything is set in advance, and there is in effect one long line of causation from the Big Bang through the present to the future. Human will is nothing more than a mental result which is also one link of that chain. It is caused by the previous links, and participates in forming the following links. By contrast, the libertarian distinguishes strongly between the future and the past. The past is set and unchangeable. The future is still in our hands. It depends on what we decide and what we do. One sharp expression of this deterministic conception of time is a consolation letter sent by Einstein, who was a clear-cut determinist, to the family of Besso, his good friend from the Swiss Polytechnic Institute, after the latter passed away. Einstein wrote there:26 Now he has departed from this strange world a little ahead of me. That signifies nothing. For those of us who believe in physics, the distinction between past, present and future is only a stubbornly persistent illusion.

What this means is that Besso’s death on that particular date is a true fact, and it was always true. Nothing changed when Besso dies on date X, as the statement that Besso died on date X was true even ten or a thousand years before, and would be correct for a thousand years afterwards. This is a look at the timeline as though it is static, with nothing really changing anything. The whole question is which of the moments we focus on or where we are located. I will note that in terms of the theory of relativity, the basic picture we observe is the “world line” — a whole graph of location and time (that is, the path) of the body. In the theory of relativity, we do not discuss the location of the body at any given moment, but rather its world line, that is its space-time totality. The dynamic approach to time sees the specific moment and place as the basic event, while the theory of relativity sees the world line as the basic element. It is there always, and we focus on this or that moment along its length. It is not that this moment “arrived,” or was “born.” It was always there. It is we who are presently focusing on it. Against this is the approach which assumed the flow of time (creative time, or “duration time”) one of whose strongest defenders was the French philosopher Henri Bergson. In his book Duration and Simultaneity,27 he argues with Einstein about the meaning of the theory of relativity. Many accuse him of not having the appropriate background in mathematics and physics, and the book is therefore considered by many to be riddled with errors. It’s true that Bergson had no chance against Einstein in this debate, but his conception of time is not necessarily mistaken. To the contrary, many philosophers (including yours truly) believe that it was he who had the upper hand philosophically, even if the theory of relativity uses a different picture of time. Bergson supported a non-static picture of time, and in this sense (and not just that) we see a reflection of the libertarian position. 19 Interview with Semone Grossman, “The Freedom to Park in the City” (Hebrew), Ynet, 12.12.2010. http://www.ynet.co.il/articles/0,7340,L3997032,00.html. 20 See below regarding the identity of this “other factor.” 21 Philosophers Daniel Dennet and Harry Frankfurt are primarily responsible for properly conceptualizing and establishing this argument. 22 This is in reference to the quartet of philosophical books I wrote, which deal with the tension between analytical and synthetic thinking in various fields (the first book deals with philosophy, the second with science, the third with understanding the human spirit, and the fourth deals with law and Jewish halacha). All were published by Beit El and Mida Tovah. 23 We should note that there are more complex presentations of the compatibilist position, such as that of Harry Frankfurt, which distinguishes between “first and second order volition.” See his controversial article: Harry Frankfurt, “Alternate Possibilities and Moral Responsibility,” Journal of Philosophy 66, 1969, pp. 829-839. However, these presentations did not escape the determinism at their foundation, so I will not deal with them here. An accessible and easy to understand critique of compatibilism can be seen in Richard Taylor, Metaphysics, Prentice Hall, 1963, where he calls this approach “soft determinism.” 24 Taken from an interview with him for the Hebrew journal Odessiya, issue 9, October, 2010. 25 Prof. Sompolinsky brought it up during the course I mentioned in the preface. 26 Freeman Dyson, Science and the Search for God: Disturbing the Universe, New York: Lantern Books, 2003, p. 24. 27 Henri Bergson, Duration and Simultaneity, Trans. L. Jacobson, Indianapolis: Bobbs-Merill, 1965.

Chapter Two A PHENOMENOLOGY OF LIBERTARIANISM Lev Tolstoy was right when he dreamed of being put in prison. At a certain moment the giant began to dry up. He actually needed prison as a drought needs a shower of rain... And I say without hesitation: ‘Bless you, prison, for having been in my life!’ (Solzehinitsyn, The Gulag Archipelago) Philosophical and scientific discussions of the question of free will often result in quite a bit of confusion due to the lack of attention paid to the full phenomenology of the thesis they are discussing — i.e., due to their inattention to part of the libertarian picture. In the first chapter, we saw one such typical mistake: The identification between the lack of ability to predict and free will itself. We also hinted at another error, this time from the other side: The confusion between randomness and free will. Others make use of scientific arguments — from the fields of psychology, genetics, and neuroscience — to “prove” determinism and attack libertarianism. To the best of my understanding, all these arguments are based on errors, and to understand this there is no need to have a complete command of the scientific fields in question. For us, it is sufficient to note what libertarianism claims and what it does not. Such errors are quite common in discussions of free will, even when they involve scientists or philosophers, and they lead to a great many misunderstandings. I therefore decided to devote a chapter in this book to describe the libertarian picture itself and present the different phenomena and their relationships according to libertarianism. I believe that those who read this chapter in depth will see, despite the apparent simplicity of its arguments, that it solves almost all of the usual problems which arise when discussing free will. It’s important for me to stress that this chapter’s goal is not to argue against determinism or in favor of it. My only aim right now is to present the libertarian picture in a complete and consistent manner. It may be that parts of this chapter will seem too detailed and a little exhausting to the reader, but I hope their necessity will become clearer later on.

Three Different Mechanisms for the Relationship Between Events28 A significant portion of the problems which arise in our field of inquiry do so because of the omission of one of the three different mechanisms for relating between events. To present these in a more understandable manner, I will present three examples of choices in the political sense. The citizens vote in general, and every one of them by virtue of their decision acts as an individual, creating some political result. The three mechanisms I will describe here regarding political elections determine a relationship between the circumstances before the decision and its results.

A. Elections in Syria. Syria has entirely free elections every few years, in which a voter enters the polling booth, entirely freely chooses the only option there and puts it in the box. According to the best traditions of Syrian democracy, the candidate with the most votes wins — Assad. Assad the father, Assad the son or maybe the Holy Ghost, so long as they got the last name right. What, indeed, is the problem with Syrian democracy (aside from massacres of civilians, denial of human and civilian rights, and small stuff like that)? The problem is that the voter is indeed entirely free (no one tells him who to vote for), but he has one and only one choice. In the previous chapter, we saw that one cannot speak of free choice in such a case. Thus, the absolute freedom in which the voter operates avails him nothing and cannot be truly realized. Elections in Syria are the equivalent of determinism, or compatibilism, as in spite of the voter’s illusion of choice (to the extent it even exists), the result is predetermined by circumstances. It’s pretty clear that in such a case, the voter is also not responsible for the results of these elections, as he in fact had no influence over them.



B. Elections in Switzerland. In Switzerland, by contrast, the situation is entirely different. There, the voter enters the polling booth, and entirely freely takes one of a number of pieces of paper available to him according to his preferences. He then places it in the box and the party which gains the most votes wins. Ostensibly, this is truly democracy at its best. What could be more democratic? Still, there is a serious problem with Swiss democracy. The problem is that our metaphorical Switzerland has no real difficulties or problems which it faces. In such a situation, it is clear that the policy and people running the Swiss system won’t really change anything. There is no price which the Swiss citizen is set to pay for choosing wrongly and also no benefit in choosing correctly. In Switzerland, a state without problems, one could flip a coin instead of holding elections. Since the results don’t matter and don’t really affect citizens (aside from the candidate winning getting a Prime Minister’s salary), there is no point in choosing a candidate based on the citizenry’s judgment. Elections as they actually happen lack any meaning, and they are therefore nothing more than a convoluted and inefficient way to do a coin toss. Elections in Switzerland, then, are an expression of the mechanism of randomness.

C. Elections in Israel. In Israel, the citizen goes into the voting booth and chooses one of a number of alternatives. The piece of paper representing his choice is then put in the polling box, and the winner is the side that wins the most votes. Israel has quite a few problems to solve, and the solution of many of them is unknown to us. We can only choose what to do in such a situation, and thus determine the benefit we will derive or the price we will pay. A bad choice means a heavy cost for us all, while the right choice means benefit or profit for us all.29 Such elections can be seen as free in the fullest sense of the word. Elections in Israel, then, are the equivalent of free individual human choice. The person considers what to do, decides which step to take, and then acts based on that decision. He is therefore responsible for the decisions of his actions (even though they are, as noted, largely dictated by circumstances external to him). The benefit or harm deriving from his choices is his and he thus bears the responsibility for it. Just as elections in Israel are not elections in Syria or Switzerland, human free choice is neither determinism nor randomness. It is a third type of mechanism, different than both.

Between Randomness and Choice As I already noted, these three descriptions are equivalent to the three ways in which an individual human being conducts himself. Elections in Syria are equivalent to deterministic conduct, in which the result of human actions is predetermined by the individual’s circumstances. Swiss elections are equivalent to random conduct (a coin toss), in which there is no connection between an individual’s “choice” and what actually happens to him in the future. Meanwhile, elections in Israel are equivalent to free choice in its fullest sense. In such a situation, the individual decides what he wishes to do, and this decision determines what will happen to him. He therefore bears the responsibility (both the cost and the benefit) of the consequences of his decision. This analogy clarifies an important point: In the previous chapter, we noted the common confusion between randomness and choice. In terms of the previous analogy, this would be to confuse Swiss elections and Israeli elections. True, with both mechanisms, the present circumstances do not determine the future results — i.e., we have no ability to predict the results in either case (aside from the Syrian one) — but these are certainly not the same mechanism. With a random mechanism, the future is determined regardless of the present circumstances, and the decision is arbitrary, as though we just had a coin toss. Meanwhile, when it comes to the mechanism of free choice, we are speaking of a decision based on the judgment of the choosing individual. He decides what is right in his eyes, and he acts based on this decision. This judgment accompanies the choice and does not accompany the coin toss, and therefore his will — based on his judgment — is the difference between these two mechanisms. Think of the person who goes and chooses a path according to his judgment. He wants to arrive at a certain destination. He knows his abilities, his limits, and the different routes leading to his goal, and in light of all these considerations, he makes his move. Another person who stands in the same spot, wants to make it to the same place and has the same abilities and limitations, but the decision he makes is based on a coin toss. In other words, he just walks in an entirely random manner, as though his muscles are driven by reflexes towards random directions. Obviously, this is not the same form of conduct. The former reflects choice born of judgment, and the

second represents randomness. We impose responsibility on the former and not on the latter. Why do so many people confuse these two mechanisms? Why is it so common to identify randomness with free choice? Simply because both mechanisms are not deterministic. In both cases, the present circumstances do not unequivocally determine the sequel. If we spoke of a stream instead of human beings, the water would flow according to the contours of the ground itself. This is deterministic conduct, in which the results are determined in a unidirectional manner according to the circumstances. Later in the book, we will see that libertarianism contradicts the principle of causality (it argues that sometimes things happen without a cause). Due to this difficulty, many of the attempts to explain libertarianism in the physical framework peg their argument to various gaps in the physical picture (primarily quantum or chaotic randomness), where physics is unable to predict what will happen in the future based on the present. This seemingly allows us to bring free will into the physicalist world picture. But as we saw in the previous chapter, the inability to predict does not necessarily mean free will. A situation of randomness is also unpredictable, but as we saw, randomness is not the same as free choice. We can argue this from another angle: Even when we reject determinism, that does not mean we are espousing free choice; our model may be that of randomness. On the other hand, if we wish to demonstrate free choice, we cannot rely on a random picture. Randomness can at most serve to attack determinism, but it cannot be an argument in favor of free choice. We will see this in further detail in the tenth chapter, where we will deal with the quantum theory. Our interim conclusion is, then, that there are in effect three different mechanisms: determinism, randomness, and free choice. We are used to referring only to the first two, primarily because it would seem that ostensibly only these two are familiar and understandable to us. Determinism is based on the relationship of cause and effect which we are fully familiar with from the world around us. Randomness is also understandable to us, as the negation of cause and effect. We understand a situation in which there is a causal relationship between event A and event B, as well as a situation in which such a relationship does not exist. By contrast, free will stands opposite both. It contains a dimension of a relationship, but not a causal one. The decision and subsequent action are not necessary the result of the circumstances in which they are taken, but this is also certainly not the random scenario of a coin toss. However, before we move on to examine the mechanisms of choice, we will also demonstrate that randomness and causal determinism, as surprising as it sounds, are also far from being entirely understandable and recognizable.

Is There True Randomness in the World? Randomness seems very familiar and clear; a coin toss, a throw of the dice, a spin of the roulette wheel, a random computer generator (a program which generates random numbers), and so on. But we ought to realize that we don’t actually recognize such a mechanism in nature. Computer people and anyone involved in computer simulations of natural phenomena know very well that all the common mechanisms for creating random processes are based on entirely deterministic processes. We have no way of generating a truly random series of numbers, that is, a series in which the numbers would be generated for no reason and based on no rules, as any mechanism which generates them is described by some rules and is therefore not random.30 Tossing a coin or dice, for instance, is always portrayed as a clear example of randomness. However, when we toss a coin, its path (and therefore also the result, the side it will fall on) is entirely determined by the rules of Newtonian mechanics. If we know the direction and power of the initial speed, the orientation of the cube when it starts its path, the shape and weight of the cube, the air density, the direction and force of the wind, and so on, the result will be predetermined and predictable ahead of time. It’s merely a question of calculation. Even if we can’t calculate the result due to its complexity, this does not mean that the result is not predetermined by the circumstances (we noted above the difference between unpredictability and objective non-determination). Therefore, tossing dice is an entirely deterministic process. The same is true of a coin toss. So why do we use probabilistic tools to analyze the results of a coin or die toss? The reason is simple — probabilistic calculation in this case is a replacement for deterministic calculation which is very complex and complicated, and this due to three factors: The number and complexity of the data which needs to be considered, the complexity of the equations and the strong dependence on the starting conditions (a small change in the starting angle may entirely change the outcome; this is what is known in mathematics and physics as a chaos effect, and we will discuss it below in chapter nine). Thus, instead of calculating the path of the die with every throw, we can suffice with the determination that if we do it enough times, the distribution will be equal (equal odds for each side). So if the throw of a die or the toss of a coin is not a random process, where can we find real randomness

in the world around us? According to the known laws of physics, the answer is nowhere, at least not at the macroscopic level. The conclusion is that the use of probabilistic and statistical tools of calculation does not mean that we are truly dealing with randomness. Only in quantum theory, which we will deal with later on, can we apparently find essential randomness (and this is also not entirely agreed upon among physicists). This means that the picture according to which two of the three mechanisms presented above (randomness, determinism, and free will) are familiar, known, and understood, is not accurate. Randomness is not familiar, not known, and not understood. To the contrary, when it appears in quantum physics, it leads to a great deal of ruckus and confusion among scientists. Even in our daily lives, an event which occurs for no apparent reason seems not understandable. It would seem therefore that the only mechanism which is truly familiar and understandable to us is the causal mechanism — determinism. However, in the fifth chapter, we will present David Hume’s discussion of causality. Hume argued that we have no direct empirical testimony of a causal link between events. When we see a man kick a ball and the ball flying afterwards, we have no direct sensual testimony to the kick being the cause of the ball’s flight. All we can say for sure is that the ball flew immediately after the kick. This would imply that the causal mechanism is not the result of empirical experience but of our way of thinking. We thus complete the confusion on this matter, as the two mechanisms which were seemingly understandable and familiar to us (causal and random) suddenly seem unfamiliar and not understood. At the end of chapter six, we will arrive at the surprising conclusion that it is free choice, so embarrassing and confusing, which is the most familiar and understandable among these three mechanisms.

The Problem of Libertarian Psychology In the previous chapter, we saw how compatibilism identifies the “me” with the totality of my psychology at any given moment, and it therefore sees the determination of our decisions by internal circumstances inside us as free choice. But real libertarianism rejects this. As far as it is concerned, free will means that our decision is determined autonomously by our will, not by any foreign circumstances, external or internal. Does the libertarian mean to deny the psychological influences which exist on human decisions? Does he deny that a nervous person will decide differently than a calm one, or that a miser will react differently than a spendthrift? In what sense can we interpret his claim that our internal circumstances in the present do not determine our decision? Does he believe that my character does not influence my decisions? That sounds entirely unreasonable. We must seek out an answer for this question based on the libertarian assumption that internal circumstances (i.e., our psychology) are an “other factor” — that is, the assumption that psychological predetermination is nothing more than determinism, and therefore libertarianism rejects this as well. How can we understand this, while at the same time not deny psychology? Or perhaps libertarianism does indeed force us to entirely forgo psychology? This question is acute for the discussion later in the book, and I will therefore dwell on it a bit. I will present a general libertarian paradigm for the relationship between freedom of will regarding any foreign factor — be it genetics, psychology, neurology, education, environment, or any other internal or external influence.

The Parable of the Hiker Think of a hiker in an area with a complex and varied topographical landscape. At every point, he must decide which direction to take. He can climb the mountain or slide into the valley. Sometimes, he decides to take the ridge, and sometimes he goes up the hill or continues on the plain. There are difficult paths and easy ones. Now let us think of a hiker on the ridge point: He has two mountains he can ascend to from north and south, and two ways to descend into the valley from east and west. What to do? If he was a small ball in the same spot, he would go down the hill in one of the two directions. If he was an accumulation of water, he would also find himself flowing downward. However, as he is a human being, he can also decide to climb one of the mountains — i.e., take the path of greater resistance. He can of course also decide to take the easy way down. The difference between the human and the ball or the water is that for them, the topography necessarily dictates the direction of advance. They always strive for the lowest point (in the language of physics: the place with the minimum of potential energy), and their conduct is therefore deterministic — entirely determined by circumstances. By contrast, the human being does not need to follow topography; he can choose where he wishes to go, even if this costs him energy and involves greater difficulty. In contrast to the ball or the water, the individual human is not necessarily driven by external physical forces. These only

determine how much effort he will have to exert to carry out his choice. This parable helps us to better understand the role of the constraints created by human psychology and genetics, as well as the society in which he operates. This entire system is not the person himself, but rather a periphery which surrounds him and influences him. It is a system of constraints (internal and external) in which he operates. The libertarian believes that these constraints do not dictate his choice but are only a causal system which set costs and benefits for every action he takes. Man is not a ball or a water flow, and he therefore has the possibility of deciding, this way or that, and to choose whether to yield to pressure and circumstances (and slide into the valley) or act against the pressure (climb the mountain). It’s important to stress that it is not always the case that the valley attracting us is the “evil inclination,” that is, it is incorrect that being attracted to it means desiring the “wrong” actions. In many cases, humans are attracted to do the good, as it is very comfortable and appropriate given his education, leads to society’s esteem, and also brings him satisfaction. Thus, the various impulses and influences, the inclinations, the social pressure, the genetic structure, education, physical fitness, mental capabilities, tolerance, patience, energy, social sanctions — all these and more are part of the topographical map in which the individual is journeying. A “valley” in this “map” represents an action which is easy for the individual to do, or that he is attracted to doing (he gains “energy” from doing it). By contrast, a “mountain” is an action which is difficult for him to do; it requires great efforts and strength. What is the meaning of the influence of all these topographical pressures and impulses on the decisions of the individual? To understand this, we need to think now of two people with similar characteristics in terms of the relevant parameters (interests, physical fitness, tolerance, and so on). We will choose a point in the ridge in which the beauty (in the eyes of the hiker) of the valleys is similar to that of the mountains, but the ascent to the mountains is very steep and the descent into the valleys is more convenient. We will place each of our individuals at that point, and let him decide which direction he wishes to continue his hike. The libertarian believes that despite the similarity between the two people, each one of them can decide to go up or down and there is nothing necessary about both making the same decision. For one, reducing difficulty is what matters and he will therefore slide down to the valley, while the other will perhaps prefer the challenge in climbing the cliff and will decide to ascend the mountain. What happens if we do such an experiment with a thousand people? It’s likely that there will be more people who will decide to go down into the valley than climb the mountain. After all, it’s the same beauty with less effort. Ostensibly one could therefore conclude that the topographical landscape necessarily determines the decisions of the individuals being tested, as we saw that effort determined the breakdown of the tracks. However, this is a mistake. It may very well be that every individual made an independent and autonomous decision, but he also took into account the constraints, costs, and benefits of each direction among available alternatives. The decision tended to go downward because the constraints influenced his judgment and therefore also his decision. But that does not meant they dictated it: As I already noted, there will also be those who prefer the challenge of the climb. As I already noted, my aim is not to prove libertarianism here. Seemingly, the very fact that there will be people who still decide to climb proves that the topographical landscape does not establish the outcome in a deterministic manner. But this is of course not proof, as it can certainly be that what determined the minority decision was some internal factor (a personality which seeks out physical challenges, for instance). Therefore, no such experiment can prove libertarianism. I will also note that even if all the people in the experiment moved in the same direction, it would still not necessarily demonstrate the validity of determinism. My goal here is not to prove but to demonstrate, in order to clarify what I speak of when I discuss influence which is not deterministic. The libertarian also sees psychological influences as influences which set costs and benefits, but in the end, the decision is by force of the free will of the choosing individual. The topographical landscape of the costs and constraints in which every individual makes their own decisions does indeed determine the tendency towards the results, that is their distribution, but not necessarily the results themselves. Bottom line, libertarianism does indeed argue that circumstances do not determine the result, but it certainly agrees that they influence it.

A Note on the Question of Materialism In the picture we presented above, there is a collection of pressures, tendencies, and impulses which attract or push the individual in different directions at different levels of intensity. If we necessarily acted based on them, we would have to calculate all of them and automatically go in the most convenient

direction, that of minimum energy (the valley). But the libertarian picture also includes a stage of judgment and decision, in which the individual must decide if he will go with these pressures or against them, or perhaps somewhere in the middle. The circumstances do not determine his conduct, only the costs he must bring into account when calculating his moves. This description implies another aspect of the libertarian approach to the individual: There is a totality of psychological constraints which constitutes a topographical landscape within which the individual operates, but who is the individual? Who is that one who acts within this complex system of pressures and circumstances? The individual himself is something else, more abstract, which operates within this landscape. The individual is the choosing “object” which chooses among the possibilities, and the psychological topography within which it acts only tries to sway it in different directions. By contrast, the determinist does not accept the distinction between these two dimensions of the decision process. According to him, the topographical landscape is man himself, as we saw in the discussion of compatibilism. The internal factors are not “other” factors to him, but are the individual himself. Therefore, he sees these influences as being the decisions as well. In the basic sense, he sees the individual as the ball or the water, whose path (choice) is entirely determined by the parameters in which he operates. The determinist effectively sees the topographical layout of the individual’s mind as the individual itself, not just his surroundings as the libertarian would argue. An obvious interpretation of this dispute is that we have here a debate about materialism. The libertarian bases himself on dualism — i.e., seeing the individual as being composed of spirit and matter. Biology and physics determine the topographical landscape (which may also include psychological tendencies), and the spirit, freed of the shackles of physical causality, carries out its free decision within this psychological topography. The determinist does not accept the existence of the spirit and sees the human being as a collection of material subject to the rules of psychophysical causality, and therefore he sees the biological topography as determining his decisions and his behavior. Let me repeat that this interpretation is not necessarily correct. I already noted in chapter one that at least logically, there can be a libertarian position which is also materialist, as well as a determinist position which is dualist. In the following chapters, we will return to the connection between the question of dualism and that of determinism.

Education in the Libertarian Picture In the libertarian picture, education — much like genetic, psychological, and sociological influences — primarily shapes the environment in which the individual operates, not the individual himself. Various conditionings create mountains and valleys and determine that actions in certain directions will be either easier or more difficult, but the decision will forever be that of the individual himself of his own free will. But this is not because education determined his decision at any given moment, but rather that education created a mountain in the direction of violence, and therefore the odds that the individual will take that path are lower. In a society without education for restraint, there is no mountain in that direction, and it is therefore clear that more people will react with violence. But again, that does not mean that education determines the response, only that it influences it. It determines the environment in which the decision on the response is taken, not the decision itself. Do we have the ability to use educational means to also influence judgment and decision making themselves, the second stage in the process described above? Can it be that education not only establishes the parameters around us including a psychological mountain making violence more difficult, but will also influence us to adopt in given parameters a decision which is the opposite of ethical considerations? Here we are speaking of education and not conditioning, but I believe this is possible. At least humanistic education and psychology believe so. To the best of my understanding, one cannot truly test this matter empirically, as an experiment using statistical methods, one which examines how many people from each group reacted in a certain way, will give us tendencies and not influences; if the group which underwent education for tolerance reveals less violent characteristics, we will conclude that this education is effective. But can we truly test where this effectiveness came from? That is, did the education reshape our environment (erected a mountain in the direction of violence), or did it reshape the choosing individual (that is without creating a mountain in the direction of violence, it created a stronger strength within him to resist the test and not react violently)? Can education create an individual who chooses differently, and not just shape a different environment? We return to the question of the drill brought up in the preface. Is man a tool which if not operating properly should be fixed, or is he a choosing creature who when erring should be educated? Is education nothing more than shaping our environment in a manner ensuring we behave properly, like the repair of a

drill, or does education act on our autonomous decision making? As noted, there seems to be no empirical way to decide this question, and it therefore remains open. Every individual must form an opinion on the matter based on their logic (or education…) and judgment.

How Does Psychological Statistics Work? Think of a psychological experiment, in which we examine the influence of education on violent behavior. Let’s assume that at the end of this experiment we will end up with some kind of graph showing the rate of violence as a function of investment in education. We will therefore conclude that there is clear dependence, and that the level of violence in the group is a direct function of the amount of education invested in that group. Does this not point to deterministic influence? Does it not show that education is a form of programming (or repair, like a drill)? My argument is that this is certainly not the case. Let us define our axes quantitatively. The independent axis measures the amount of education. For purposes of demonstration, let us assume that we measure it in terms of the number of years in which five hours a week were devoted to education against violence. The dependent axis represents the level of violence. Here, too, we must define a quantitative level, and we will do so with the percentage of children who hit another child and caused him some sort of injury. Now let us assume that we found some graph which describes the relationship between the amount of education and the level of violent conduct:

If the measurement is reliable, and if we indeed find that there is a valid quantitative law here (a rare situation in psychology), this means that in every group we take, the findings will be similar. The Law of Large Numbers states that the larger the sample, the closer the results will approach the expected average. For instance, if we throw a die 12 times, it’s unlikely that the distribution of the results will be exactly twice (odds of 1:6) for each side. However, if we throw a die 12 billion times, the results will draw far closer to 2 billion for each side. This is the Law of Large Numbers, which states that if we repeat the experiment enough times, the results will move ever closer to the results predicted by the average. The Law of Large Numbers deals with random processes. How can it be that we reach average laws which truly work on large numbers of samples, even when we are dealing with human behavior? After all, they do not reflect random processes. How can it be that in any sufficiently large group, the amount of violence after three years of education will be 20%, for instance? Essentially, what we are asking here is this: How can it be that one can operate in the field of psychology with the tools of statistics and probability? The ostensibly obvious conclusion from the existence of psychological laws is that our behavior is random.31 The number of quantitative laws in psychology is negligible. But let’s assume for a moment that we did arrive at a whole set of quantitative laws in psychology, a system which allows us to predict with great precision the average of every group experiment we carry out. Does that mean that individuals have no free choice? Certainly not, as these laws speak of average behavior, and they are correct for any experiment dealing with a sufficiently large group of test subjects. However, this is not enough to say that one can truly use them to predict the behavior of any individual human being. The deviations from the average (i.e., people acting unpredictably) demonstrate as much. Many claim that the fact that we have no system of laws which perfectly precisely predicts the behavior of an individual human being derives from the great complexity of the human psyche. This means that in principle we certainly could reach such a level of predictive ability, if we continue with a multi-year, systematic study and map out all the parts of the human soul (and perhaps we are speaking of the human brain; see the last chapters).

We may not reach that point, but this is simply because research will fail at this mapping due to the great complexity of the task. The libertarian, by contrast, will argue that this task is impossible in principle; it is a lost cause from the start, by definition. According to him, the problem is not technical but fundamental, and that we can never reach a “perfect” psychology. If the psychological system with all its tendencies and influences on us at any given moment do not determine our conduct at all (only influencing it, as we saw above), then psychology is doomed to failure or inherent imprecision from the outset. Above, I distinguished between the lack of predictability and the objective indeterminism. Here I must note that the problem I am dealing with is not the problem of predictability. The inability to predict the behavior of an individual human being is just an indication of the objective inability to determine it. Libertarianism argues not only that the perfect psychologist cannot predict the behavior of the individual person, but that circumstances (the psychological topography and others) do not at all determine this conduct, not even at the objective level. We are not speaking here of inability or lack of information. Circumstances do not determine conduct, they only influence it. As we saw, this can be reconciled with the agreement of the libertarian that a system of laws which is correct on average for any sufficiently large group is a possible task, at least in principle. The significance of these average laws is a reflection of the environment which influences a human being of a particular sort. On average, we behave according to our internal and external influences, and therefore in a large group, it may well be that the decisions of the average person will be a determined function of psychological and other functions which operate on him. This average reflects the topographical landscape which influences us, and therefore it has on average a clear statistical relationship with the decisions we make as a group. But every individual human being has another level of freedom and that is his free choice, the stage at which he decides whether or not he wishes to yield to this topographical environment or act in another direction. Therefore in the libertarian picture, psychology, like any tool, will never entirely predict the conduct of an individual human being. So, quantitative psychological research can at most strive to fully describe the internal topographical landscape (both the inborn landscape and that formed by external influences) of every one of us. Even if psychological research is perfected at any point in time, this will be expressed at two levels: A. A system of average laws which will be true for large groups. The libertarian argues that there can never be a psychology which fully determines the effective conduct of the individual human being. B. The libertarian agrees to the possibility of drawing a complete psychological map of the individual human being standing before us. But the conduct of that individual will always contain an additional non-deterministic component, which will prevent perfect prediction. One can speak of what is likely to happen or the average when it comes to the behavior of the individual human being, but perfect prediction will never be possible.

On Incitement and Tipping the Scales The libertarian picture described here can help us derive and understand mechanisms of public influence. When an individual incites a particular public to carry out a crime, say murder, every single individual still has free will to act based on their own understanding. Thus, there is seemingly no room to punish the inciter for their action. Jewish law puts it as follows: “There is no [non-accountable] messenger for sin” (Kiddushin 42b) — that is, if Reuven sends Shimon to commit sin, it is not Reuven who is punished but Shimon.32 The argument presented by the Talmud is: “The words of the master and the words of the student, whom do you listen to?” (Cf. also Sanhedrin 29a and Bava Kama 56a). If God (the master in this scenario) orders you not to do it, and another person (the student) incites you to commit the sin, who must you obey? This consideration removes criminal liability from the inciter, as the messenger has free choice, and if he decided to murder, that is his responsibility and he must bear the consequences for his actions. But regular courts of law also impose responsibility on the inciter. Why? Ostensibly, the one who commits the act does so of his own free will. The inciter did not hypnotize him or deterministically change his way of thinking; at most he convinced him. One can see incitement as a reshaping of the psychological environment of the individual (primarily if it was not done via his cognition but rather through rhetorical and perhaps demagogical influence, of the sort that appeals directly to his subconscious). Such shaping leads to a situation in which there are greater odds that he will do the deed (there is a valley in the

direction of that action, and it is easier for the one being incited to descend to it), and therefore the inciter bears part of the guilt. We can now understand why incitement is seen as more serious when done towards a public. In such a situation, tipping the psychological scales of everyone towards doing the action (in the wake of the incitement, abstention is a taller mountain and action is a steeper valley) will increase the odds that the act will be carried out. If there is a public of a thousand people, all of whom are in a balanced psychological and moral state, it may be that none of them will commit murder. However, if someone comes and tilts that statistical scale and turns the topographical landscape of everyone into an unbalanced one, then every one of the thousand now has a slight tendency towards murder, say at a rate of 1:1,000. Regarding an individual person, the odds he will carry out a murder as a result of the incitement is still very small. However, a group containing a thousand individuals, the incitement will greatly increase the odds of finding at least one person who will effectively carry out the murder. This tipping of the relative scales is the fault of the inciter.

First Intermezzo: The Determinism of the Collective A sharp expression of this approach can be seen in the debate between Maimonides and the Raavad regarding the punishments imposed by God on the Egyptians who enslaved the Israelites in Egypt (Hilchot Teshuvah 6:5): [A question may still arise, for] behold, it is written in the Torah (Gen, 15:13): “They shall enslave them and oppress them,” [seemingly implying that] He decreed that the Egyptians would commit evil. “Similarly, it is written (Deut. 31:16): “And this nation will arise and stray after the alien gods of the land,” [seemingly implying that] He decreed that Israel would serve idols. If so, why did He punish them? Maimonides is asking here why God punished the Egyptians if he decreed in advance that they would enslave the People of Israel. The same is true regarding the idolatry committed by Israel for which they were punished — he asks how could they be punished for this if it was predetermined?33 He responds: Because He did not decree that a particular person would be the one who strayed. Rather, each and every one of those who strayed to idol-worship [could have chosen] not to serve idols if he did not desire to serve them. The Creator merely informed [Moses] of the pattern of the world. To what can this be compared? To someone who says, there will be righteous and wicked people in this nation. [Thus,] a wicked person cannot say that because God told Moses that there will be wicked people in Israel, it is decreed that he will be wicked. A similar concept applies regarding the statement (Deut. 15:11): “The poor will never cease to exist in the land. Similarly, in regard to the Egyptians, each and every one of the Egyptians who caused hardship and difficulty for Israel had the choice to refrain from harming them, if he so desired, for there was no decree on a particular person. Rather, [God merely] informed [Abraham] that, in the future, his descendants would be enslaved in a land which did not belong to them. We have already explained that it is beyond the potential of man to know how God knows what will be in the future. Maimonides claims that there was indeed a collective decree on the entire group, but that does not mean that every individual therein did not have free will. His argument is that even if there was a collective decree, every individual could act freely, and he therefore bears moral responsibility and deserves punishment for his actions. How is this possible? Based on our model, we can say that God tilts the psychological scales of the Egyptians slightly in favor of the desire to enslave Israel (creating a “valley” in that direction in the psychological topography of all Egyptians), so there will certainly be a significant number of Egyptians who will effectively do so. However, the fact that there is a tilt does not mean that each and every one of the Egyptians did not act out of free will, and therefore each one is responsible for their actions.

The Raavad, Maimonides’ great disputant, comments on this in his critical comments there: And after all it is written in the Torah ‘They shall enslave them and oppress them.’ Said Avraham [the Raavad]: These are long words lacking seasoning, and on the life of my head almost I say that these are words of youth, the Creator told the strayers why did you stray and I did not mention you by name so that you could say I decreed for you, and the strayers will respond to him and on whom does your decree fall, on those who did not stray, hence your decree was not fulfilled. The Raavad is arguing that Maimonides’ words are “words of youth” (i.e., unbalanced and unconsidered, like that of a novice student). After all, any Egyptian who is accused can claim it was decreed he act thusly. Is God supposed to “defend himself” on the grounds that he didn’t mention anyone specifically by name? Let me put it this way: Did God decree that the Egyptians enslave Israel, and let us even assume that all the Egyptians save one decided freely not to enslave them, did the remaining Egyptian have free choice? Seemingly, the Raavad’s argument was correct. Can one impose moral responsibility on an individual human being in such a case? However, in light of what we saw above, the answer is yes. Each and every Egyptian chose freely, and it is only the general distribution which was determined by the psychological scales tipped by God. This solution does not allow any one of the criminals to escape responsibility for their actions, as each one acted freely. We see here the possibility of determining in advance (in an almost deterministic manner, at least de facto) a situation within the libertarian world. There is a very high likelihood that a tipping of the scales of the general public will lead to the desired request, despite the freedom of choice of each of the individuals. This is precisely the argument of Maimonides here, and its roots are in the two-tiered picture we described above. God determined the psychological topographical landscape of the entire Egyptian people, and since we are dealing with a large group, the average conduct in practice is derived from the relation between these theoretical scales.

*** In recent years, there has been a lot of talk of the discovery of genes responsible for diseases, various physical traits (such as left-handedness), as well as many psychological characteristics. We are inundated daily with reports on genes responsible for belief in God, being cheap, hypochondria, violence, good heartedness, depression, adventurousness, charity, morality, and so on. Many conclude from all these studies that our feeling of freedom is nothing more than an illusion, and that it is our genetic structure which unequivocally determines our conduct. In fact, genetics have no definitive consequences at any stage. To the best of my knowledge, there is no trait which can be clearly attributed to one clear section of the genome, nor any complex configuration thereof, and one cannot clearly predict human character and certainly not human conduct. How do determinists deal with deviations from these results? Sometimes they attribute it to the fact that we have still not yet uncovered the entire genetic structure which determines the behavior in question (there is almost no trait attributable to a single, local section of the genome). Other explanations attribute it to various environmental influences (see the previous chapter and Ciechanover’s talk about the plasticity of the brain), or: not everything is about the genome. Still, a common assumption among researchers of the genome is that there is a complete explanation for all conduct and that human will is not a part of any of it. Many of them express a compatibilist-deterministic position, according to which will is operated by all these influences and is not an independent source of influence itself. This is of course a possible hypothesis, but to the best of my knowledge, there is not a single study whose findings can lead to such a conclusion. In fact, as of today scientific research can say nothing about this hypothesis, and it is all stated on the responsibility of the argument’s advocate. In light of the picture drawn above, it’s important to distinguish between two levels of reference: The psychological environment in which man operates (i.e., the system of pressures and impulses, tendencies, pains, and other psychological influences), and the decision of his will. As we saw, the libertarian argues that conduct is always the sum of these two components. A man who makes a decision examines the totality of the influences (his present topographical environment and the forces and tendencies acting on it), and he afterwards makes a decision (to join with the psychological tendencies and go down into the “valley,” or act against them and climb the mountain). The psychological environment of the individual human being is formed from three main factors: (1) inborn genetics; (2) environmental influences (from birth until the present); and (3) his past decisions (which also contribute to the formation of his internal environment). But all of these are only responsible for

shaping the environment of the individual (his character or sum of all his traits) and not his effective decision. This is adopted in the second stage, which takes into account all these influences, but which ultimately decides whether or not to yield to them. Thus, much like we saw above regarding psychology, genetic research can also at most only explain the inborn portion of the individual’s psychological environment. In addition, there are acquired influences (which Ciechanover mistakenly saw as being a deviation from the deterministic picture, as we saw in the previous chapter), but all these are still not yet free choice. And then there is the individual’s decision which is done from within and based on his psychological environment. As we saw, the libertarian argues that the environment does not determine this but only influences it. The libertarian believes that the relationship between genetics and psychological traits are relationships which are determined in a manner similar to that of psychological statistics which we discussed above. In other words, if we take a large group of people with the same genetic structure (theoretically), we will on average end up with conduct which reflects this genetic influence. A group with a violence gene will on average behave more violently than a group in which this gene is weaker. Again, the averages of conduct of groups reflect the rigid environment (character) of every one of the individuals (the topographical landscape in which they operate). However, as we saw, this rigid environment, even if it is entirely fixed, does not necessarily determine the conduct of the individual. To sum up, the psychology and genetics of the individual form a map in which he operates, describing his tendencies and traits, his character, but not his effective conduct. An individual can be nervous, spendthrift, friendly, and so on, and this character will create certain behavioral tendencies within him, but the decision how to act in practice will never be entirely determined by character. The picture we described thus far shows that at least as of now, the findings of genetics are not relevant to the question of libertarianism.

The Problem of Weakness of the Will: The Decision to Choose To round out the phenomenological description of libertarianism, we must add one more piece to the puzzle, and that is the decision to decide. I will explain by delving deeper into the question of the weakness of the will. As I explained previously, weakness of the will is a term we use to describe failure. Situations of weak wills appear in every field in which an individual makes a decision and wishes to carry it out, but is unsuccessful. An individual decides to do a diet and fails, usually describing his experience as a failure deriving from a weak will. In the religious field, an individual speaks of the weakness of the will after he has sinned. He wants to observe all the commandments, but he doesn’t manage. He fails and surrenders to his inclinations because his will was week. Even when it comes to obeying the law, we can speak of crime deriving from weakness of the will. An individual believes that it is proper and necessary to obey the law, but fails and surrenders to either his inclinations or his interests. The same is true of athletic or ethical failure. A description of failure as a weakness of the will is fairly problematic in the philosophical-psychological sense, and it is therefore no surprise that the question of weakness of the will fills a number of shelves in the philosophy section in libraries. Here I chose to make use of one of the sharpest and most popular formulations of the problem formed by Donald Davidson, one of the most important analytical philosophers of our era.34 We speak of two actions, x and y, and the individual must decide which one he wishes to do (either because he cannot do both — such as go to both Tel Aviv and Jerusalem at a particular time, or because these are opposite decisions — such as to kill or not kill so-and-so). In such a situation, we can raise three arguments, each of which sounds very reasonable: A. Freedom of Will Argument: When a man thinks action x is more worthy than action y — then he wants to do x more than y. B. Freedom of Action Argument: When an individual wants to do x more than y, and he is free to do both and acts consciously and is not coerced — then he does x and doesn’t do y. C. Weakness of the Will Argument: An individual sometimes acts consciously and freely yet still does y, even though he wants to do x more, even though he knows he can do either one of

these actions. His feeling in this case is that he was weak and therefore failed. Argument (a) regarding free will seems obvious. If this is what I think should be done, why should I not want to do it? In this context, it’s important to clarify that we are not talking about worthiness in the ethical or religious sense; for our purposes, “worthy” means what arises from the sum of my own values, wishes and beliefs. For instance, an individual works as a hit man. A hit is ordered, and he’s willing to do it for a proper price. When he murders a target, does he want to do it? Of course. Is this an ethical act in his eyes? Not necessarily. It may well be that he also understands that this is not a moral action, and he is not necessarily an individual uncommitted to morals in general. So why does he nevertheless commit murder? It seems that his will for money overcomes his will to be moral. For our purposes here, it doesn’t really matter. We will say here that he considers it worthy to murder, with the term “worthy” not being an ethical claim, but only in the sense that this is a positive in terms of his world of considerations and constraints. Bottom line, this is what he considers most worthy (i.e., x, the act of murder, is more worthy than y, avoiding committing murder). If it was not worthy in his eyes (i.e., he would think that the value prohibiting murder outweighs the value or interest in making a lot of money), so he doesn’t want to do it, and apparently would not do it. Therefore the argument of freedom of the will seems obvious: If it is worthy in my eyes, I wished to do it, and if I wished it so it must therefore have been what I consider worthy. Argument (b) regarding the freedom of action also appears obvious. If I want to do something and I can do it, then I will. True, here too, questions can arise. Maybe I want to but there’s something interrupting me, and therefore I can’t actually carry it out? For instance, an individual wants to be healthy and therefore thinks it worthy to do a diet. It therefore arises that he wants to diet (argument a — freedom of the will). However, this is difficult for him, because he really wants to eat the tasty chocolate in front of him. It’s hard for him to restrain himself. It therefore ostensibly seems that in this situation it may be that the individual will not do what he wants. But this is a mistake. The individual’s desire in this case includes two different components: The desire to be healthy and thin and the desire to eat tasty things. Thus, he must weigh these two desires and reach a bottom line. If he eats the chocolate — then this seems to mean that the desire to eat something tasty has overcome his desire to be thin and healthy. In other words, it’s a mistake to think that the individual did not do what he wanted to do in this case; he wanted to eat the chocolate and did. The will, just as it should, was obeyed after making all his considerations. Thus, freedom of action is an entirely reasonable assumption. Argument (c) regarding weakness of the will, as already noted, derives from our intuition. Every individual feels that sometimes he fails when doing some action or avoiding doing so. He does so not because he wanted to, and not because he couldn’t, but because he wasn’t strong enough. He feels that he acted against what he really considered worthy (after weighing all the considerations), and therefore also against what he wanted. So, argument (c) also sounds prima facie reasonable. This is where we encounter the fundamental problem. The argument of weakness of the will, which is itself reasonable, does not fit with the previous two arguments, which seem no less reasonable. If I want what I think worthy, then I do what I think worthy. How, then, can one accept the argument according to which there are situations in which I did something I do not consider worthy? One can think of two principled ways to come at these paradoxes: Find a loophole which demonstrates that the contradiction between the arguments isn’t really there, or forgo one of the arguments despite its reasonableness in our eyes. What can be the escape route in the case of weakness of the will? Seemingly, there is only one option to get out of this entanglement with all three arguments intact: There was something that hampered me, I didn’t fight my strong inclination, and therefore I failed. In other words, I really thought that it was not worthy to eat chocolate (because health is more important than momentary pleasure). Therefore my will really was not to eat it. So why did I eat it? Because my inclination was so strong I couldn’t help myself. This was not a desire to eat something tasty, as I described previously, but an inclination. The inclination, or urge, is something unconnected to my will, something external which overcame my will. In the above-described picture, the urge or inclination is somehow part of the topographical landscape in which the individual operates. As we know, this landscape applies pressures and urges on the individual which sway him to do or not do all sorts of things. We saw that in the end, the libertarian believes that the individual decides whether or not to surrender to these pressures, and this is an act of choice. Thus, the

obvious solution to the problem of weakness of the will is that the topographical landscape around me applied such heavy pressure that I couldn’t resist (the mountain was too high, and I couldn’t climb it despite my desire), and I therefore acted in a manner which was against my will. In principle, a situation of the sort so described is of course possible. We saw already that there are situations in which the individual cannot carry out his will because of urges which are too great (an urge which cannot be overcome). Jewish law also recognizes this, when it says that there are times when “the inclination overcame him,” that is the inclination was so strong that it could not be overcome, and this is a situation of compulsion which exempts the criminal from punishment. Even regular law courts recognize situations in which the individual acted based on an urge which he could not control. The criminal in these cases is not considered responsible for his actions, and there is therefore no room to punish him (though there may be room to defend against him and treat him). But this is the very reason that this solution to the paradox of weakness of the will throws out the baby with the bathwater; it does not help us solve the problem. If we do accept this picture as a solution to the problem of weakness of the will, then the argument of weakness of the will is indeed an excellent argument. If an individual reports to me that his failure derived from weak will in this sense, that is his urges compelled him and he could not fight them, then I really do not see him as responsible for his actions. Even the court would exempt him in this case. But the weakness of the will is a situation in which the individual sees his actions as a failure, and not as something which is out of his hands (onnus or compulsion, in Jewish legal terminology). The weakness of the will is not an excuse meant to lead to my being found innocent, but to the contrary — to recognize my guilt and accept that I need to strengthen my will and not fail in the future. For instance, a defendant is brought before the court and accused of murder. He is found guilty, but claims in his defense that the victim provoked him and greatly angered him. What does he mean to say by this? Does he mean to say that he was not responsible for his actions? Was this urge uncontrollable? If so, then he is exempt from punishment (and may be sent into treatment). However, in most cases this is not what we are dealing with. The judge may see this as a reason to be lenient with his sentence, but certainly not as grounds to exempt him from responsibility. Why not? Because weakness of the will does not mean lack of responsibility. I failed, but I am still responsible for my actions. I was not strong enough, but I certainly could have been. There was no action I was compelled to do. The argument of weakness of the will, then, is aimed at expressing error and weakness which must be overcome, not compulsion which exempts the individual of responsibility. Therefore, the problem remains in place. How can we accept the possibility of an action due to weakness of the will, while at the same time adopting the hypothesis of freedom of the will and the freedom to act?

The Choice to Choose It’s very hard to offer a complete solution to the problem of weakness of the will; I will only try to provide an outline which such a solution should contain. An individual who feels that he acted against his will effectively tells us that he was dragged into doing what he was doing and that there were forces which overcame him. But as we noted, when one speaks of weakness of the will, we are not talking about insurmountable forces of the sort he could not resist. We are speaking of forces he could have resisted, yet still didn’t. This means that beyond the choice, which determines what is worthy in my eyes and what I want, there is another element in this picture — just how much I want what I want; how much energy I am willing to invest to realize my decision and my will. We are not talking about measuring the power of each one of the wills against the other wills. We are dealing here with the final will, the bottom line. This brings us back to the libertarian picture described above: The issue of weakness of the will deals with measuring the power of the will against the height of the obstacles presented by the topographical landscape of the pressures and the urges in which he operated. Let us return to our previous example. A man decided that he preferred health over taste, and it is therefore worthy in his eyes not to eat the chocolate. Therefore he wants not to eat the chocolate. Still, he fails and eats it. He attributes this to the weakness of his will. Now we understand that this means that the urge which drove him to eat the chocolate was so strong that his will (composed of his desire for health minus the desire for a pleasing taste) couldn’t resist the urge. The topographical landscape before him presented him with a mountain so steep, that climbing it (that is, not eating the chocolate) would have required strong will power. If so, it’s not enough to decide that this is the decision he considers worthy; he must want it hard enough to climb the mountain (that is, go against his urges). Since he didn’t activate enough willpower, he failed. Could he have succeeded? Certainly. But for that he needed to use more

willpower. That is, when an individual chooses he operates on two levels, and each one of them is important for the success of his action: (1) He decides what is worthy and therefore what he wants, and (2) he decides how much energy (mental and physical) to exert to realize his decision. The second level is but the choice to choose (how much power to put in realizing his choose). Essentially, the individual needs to decide whether to choose, that is act based on his choices, or choose not to choose, that is be weak and let the urges (the topography) direct him.35

Decisions and Mechanical Actions So far, we have dealt with the free decisions of human beings, as well as the mechanism which serves to make those decisions. In the previous chapter, we pointed to the libertarian picture also allowing for exceptional cases, in which an individual “decides” deterministically and not based on free choice. In these cases, his topographical landscape acts in his stead. It’s important to stress that this doesn’t mean that every action he takes is like that. After all, libertarianism argues that every individual decides freely on the arguments he has before him, but the libertarian also agrees that there may be extreme cases in which he deviates from this regular conduct, and there the urges are what drive him. We will present two such exceptions here: A. An Irresistible Urge. There are situations in which an individual is pushed to do an action based on forces he cannot resist. He has an irresistible urge to murder or steal, and therefore does so. These are cases in which the topographical map’s psychological mountain is too steep, and the individual cannot overcome it. This is a sort of inbuilt weakness of the will, in the face of powerful urges which compel the result. In such situations, we lack the basic requirement for free will which we noted above, in which the individual has more than one option he can realistically take. Such human actions are not the result of his free choice, or of his will (in the libertarian sense), and therefore any court would exempt him from punishment (although they would still try to protect society from him). In such situations, the libertarian would also agree that what is acting here is the individual’s psychological topography, and not his free choice. B. Actions Done Without Judgment. These are actions where the individual could in principle act otherwise, but he didn’t use his judgment. Why not? For two reasons, which divide these situations into two sub-categories: (1) Actions in which the individual did not use his judgment. There are cases where an individual did an action requiring judgment, but where he acted instinctively — i.e., did not use consideration or thought. For instance, an individual faces the decision of whether to invest in A or B. Usually, such a decision requires consideration, an evaluation of the various options, and ultimately an informed decision. But there are certainly cases in which the individual does so hastily and without any real thought. Daniel Kahneman and Amos Twersky, in a series of works which earned Kahneman the Nobel Prize in 2002, showed that people make incorrect decisions even in situations where systematic judgment is not complicated. They just don’t use it. One can treat such a situation as the choice not to choose, but sometimes it seems that this is not a choice but rather conduct based on inattention to the fact that this is something which requires judgment. There may be a degree of negligence here, as the individual should have been paying attention. There is therefore room for seeing such a situation as a form of weakness of the will, or a choice not to choose. (2) Actions where it simply isn’t relevant to use judgment. For instance, in the Libet experiment (discussed in depth in chapter fourteen), an individual was placed next to a table with a button on it, and was asked to press the button whenever he wants. They examined how his decision was made whether and when to push and when he made it. Let us state right now that the decision whether to press the button at any given moment is certainly a candidate for a decision not requiring judgment. There is the appearance of hesitation and ambivalence, but it would seem that it is not possible to think of any considerations when pushing the button at any given moment. How does such a decision get made? It just does. It turns out that if the individual does not decide based

on judgment, the moment he pushes the button is based on neural processes in his brain. Something in his psychological topography awakens in him the impulse to push, and then he pushes. This is not a case where it’s a question of his choice whether to surrender to the pressures or these topographical urges or to not. He just places himself in their hands. In this case, too, the libertarian will agree that the action is the result of the individual’s topographical landscape and no more. It’s interesting to note that there is also a statement in the Talmud which sharply presents the approach that not all our actions are subject to free choice. The Talmud (Megilah 25a) states: “Said Rabbi Hanina: All is in Heaven, except for the fear of Heaven, as it is said (Deut. 10:12) ‘And now, Israel, what is your God asking of you but to fear [God].” That is, everything that happens in the world is in Heaven’s hands. It follows the laws of nature, and is not given over to our control. The only things for which freedom of choice and will are relevant are actions which have ethical dimensions (related to fear of Heaven). An action like pushing a button, then, is not the result of judgment. What led to the decision to push? Something physical or entirely coincidental (what we might call “in Heaven’s hands”). What decisions are based on our free will? Only those with ethical significance, that is those which require judgment.36

Second Intermezzo: The Halacha of Rabbi Ilai We find an interesting rule in the Talmud regarding the sorts of irregular cases described above, in which even the libertarian agrees that a human action was done without free choice (Mo’ed Katan 17a): For R. Il’ai says, If one sees that his [evil] inclination is gaining sway over him, let him go away where he is not known; let him put on black clothes, don a black wrap and do the black deed that his heart desires rather than profane the name of Heaven openly. This saying instructs the individual who sees he cannot overcome his inclination to wear black, go somewhere distant, and commit his sin there. Some of the later Jewish authorities argue that Rav Ilai aims to argue for damage control (See Rashi’s commentary, starting verse “Ma”). In a distant location, the degree of desecration of God’s name in the sin lessens (no-one knows him), and therefore if he’s already going to sin, it is best he do so there. Others argue that if he goes somewhere far away, and especially if he wears black, perhaps he could overcome his inclination and not sin (See commentary of Rashi in the name of Rav Hai Ga’on, as well as the opinion of Rabeinu Hananel presented in the Tosefot, starting verse “Im Ro’eh,” and Tosefot, starting verse “Vaya’aseh,” Kidushin 40a, and more). The fundamental difficulty which arises regarding this saying appears in the words of the Rif37 (9a in his section) and the Rosh’s commentary: And it is not like Rav Ilai but rather even though his inclination is overcoming him he must settle his mind as we accept that everything is in the hands of Heaven except for fear of Heaven. The Rif assumes that there cannot be a situation in which the individual cannot overcome his inclination, and he therefore does not rule like Rav Ilai. The reason: “Everything is in Heaven except for the fear of Heaven.” Above, we saw that this statement expresses the assumption that actions related to ethical matters are given to our decision and are not in the hands of Heaven (or nature and physiology). This is precisely the principle we pointed to in the second irregular case noted above. By the way, this is apparently the basis for the method which interprets Rav Ilai’s words as a tactic to avoid sin (those who adhere to it are not willing to acknowledge situations in which people do not have the option of choice). What, then, does Rav Ilai himself believe? We should note that there are those who rule like him (Cf. Tosefot “Vaya’aseh,” Hagigah 16a). Some commentators understood Rav Ilai to be a determinist. We’re not necessarily dealing with an assessment that his will was too weak, a case of weakness of the will, as in such cases Jewish law would order him to strengthen his will (that is, choose to choose) and overcome his inclination. It is not reasonable to assume that Jewish law would be prepared to give an individual legitimacy to sin. This is the reason that some of the commentators reject Rav Ilai. But even he himself does not necessarily dispute this principle. From the irregular cases we discussed above, we could argue that what Rav Ilai is talking about is an irresistible urge, of the sort even libertarians accept exist. In other words, this is a mountain he cannot climb. In such a situation, says Rav Ilai, the individual must do damage control, as he has no real possibility of overcoming his urges. But this does not mean that Rav Ilai thinks that an individual can never choose in any situation and is truly a determinist. To the contrary, he himself seems to be speaking only in those irregular cases in which he is dealing with an irresistible urge.

There is another interesting Talmudic saying which points in a similar direction. In the Talmudic tractate of Berachot 8a, a contradiction is presented between two braytas:38 Said Rabbi Hiyya son of Ami in the name of Ulla: A man should always live in the place of his master, as so long as Shimi Ben Gera lived, Solomon did not marry the daughter of Pharoah. And yet another brayta says: He must not live! — There is no difficulty: Here when he was subservient to him, there when he is not subservient to him. One brayta instructs the individual to live in the residence of his master, and the other recommends living far away. The gemara reconciles the contradiction on the grounds that the first brayta deals with a situation in which he is subservient (obeys) to his master, while the second brayta deals with a situation where he is not subservient to him. If he does a sin and his master corrects him, if he is subservient to his master he will mend his ways. However, if he is not subservient to his master — he will not obey him and then we have a double problem (deliberate sinning is more serious, aside from the problem of contempt for the rabbi). Here, too, we have a similar problem. If the individual is indeed not subject to his master and is not willing to mend his ways when his master calls him out on his error, then Jewish law should have rebuked him and told him to overcome his inclination, i.e., strengthen his weak will. Instead, Jewish law recognizes the existence of this situation, and instructs him to live far away. This is ostensibly an ostrich-like policy, and it seems unreasonable. Thus, we likely have here a situation where the individual evaluates that he cannot be subject to his master (an irresistible urge), and not one of a state of weakness of the will.

On People and Bees In an article published by three Israeli researchers in the journal Nature,39 the authors compared the decision making of bees and human beings. They arrived at the conclusion that people, like bees, choose the option which brings them the highest amount of successes, even if they don’t actually maximize their profit. The experiment presented test subjects with two options: Press button D (dangerous) which brought them a profit of four units but only in 80% of cases, or press button S (safe) which always brought profit, but at a lower rate of three units. Most of them chose the second option, as it gave them 100% chance of success, even though the profit margin (3) was lower than the first option (3.2). A similar experiment was conducted with bees with two sorts of sugar solutions at different levels of concentration which produced similar results. The bees had to choose an option that provided them a solution with a high concentration of sugar in eighty percent of the cases, or the option that would give them a solution with a lower concentration in one hundred percent of the cases. The bees, just like the people, chose the second option. They also preferred the safe option over the more profitable one.40 This is one example of many where it comes to there being a substantial similarity in conduct between animals and humans (in this case, decision-making). What are the conclusions to be had from these results? The formulation of the conclusion as presented by the researchers was that human decisionmaking is similar to that of bees. But is the decision-making mechanism similar? Do human beings also make decisions based on a mechanical calculation, or do judgment and even choice drive them, neither of which appears to exist for bees? As we will see later in the book, neuroscientists tend to think that the human decision-making mechanism is also built like a mechanical calculator (at its root, it is a neural process and not a decision at the mental level. See below chapter ten and following). So, according to them, there is almost complete similarity between humans and bees — and for some reason this mechanical mechanism failed in decision-making. This is certainly a possible approach, but it is certainly not a necessary conclusion from the experiment. To understand why, let us think of a situation in which we explained the logic of the profitability to the test subjects. I assume that many of the subjects would change their decisions afterwards. Bees have no such option (maybe they can be trained, but not taught). The determinist will also see this as a deterministic mechanism, as the “explanation” is nothing more than a reprogramming of the brain. However, the libertarian will argue that this experiment did not actually examine the nature of human decision-making, only their instincts. As we saw throughout this chapter, everyone operates within their topographical landscape which applies various pressures and urges on them. Within these parameters, the individual must decide whether to follow these urges or not. In our case, we are dealing with the urges applied by our instinct, which tell us to go for the safe option. However, even those who feel these urges can listen to the consideration of profitability, be convinced and choose to act against the instinctive urge. The libertarian argues that the possibility of convincing the individual expresses his option of choosing to act against his environment. We should note that even if this action involved using the mechanism of judgment and choice, we might still end up with similar results for

humans and bees. These similar results are only at the level of the average human being or bee when measured in large groups, but not in reference to the decision of an individual creature. As we saw, the libertarian agrees that the law of large numbers also works for both random events and choices made by free will.41 So, the human instinct (his topographical environment) is indeed similar to bees. However, the human is different from the bee in that he can overcome his instinct, and decide against it. This may be the result of study and thought — that is, of judgment. Therefore the comparison of human beings to bees, which is often translated into determinism or realism, is of course not a necessary conclusion from these experiments. Experiments which point to a similarity between the conduct of human beings and animals are at most describing instincts — that is, the topographical environment in which the human being operates, not the person himself. The second stage of the decision, in which the individual decides whether to follow his instinct or oppose it, is what separates man from beast. Similarly, one can point to the impressive collection of fallacies in probabilistic and statistical thinking and in decision-making pointed to be Kahneman and Twersky. All these fallacies describe an incorrect instinctive manner of decision-making, but Kahneman and Twersky are themselves human beings. So, how do they break free of these foolish decision patterns? This is done via the controlling mechanism of higher thinking, which takes over instinct and can correct it. In all these cases, after the error is explained to the test subjects, they have the ability to change their behavior, and they will probably do it. These are type42 B143 irregular actions of the sort described above, where judgment is certainly a factor, but where quite often human beings don’t use it, instead acting instinctively (and incorrectly, in these cases). In chapter fourteen, we will continue to discuss this issue, and we will see that even training does not ensure we will witness non-instinctive behavior. It may be that this training only refines our instincts. We will conclude that there is in effect an almost impassable obstacle before any systematic scientific study of the question of determinism.

Between Freedom and Liberty We will end this chapter with an important distinction which illuminates the picture described here from another angle. This distinction is also relevant for understanding the findings in neuroscience which we will discuss in chapters fourteen and fifteen. Many err in thinking that freedom of choice means acting without constraints. These approaches see any constraint as contradicting freedom of choice. For instance, many ask the following in the theological context: If God gave us freedom of choice, then why does he command us what to do and what not to do? This would seem to limit our freedom of choice. Moreover, why does he punish those who sin or who do not fulfill a commandment? After all, such a punishment ostensibly contradicts freedom of choice. But there is an error here: The fact that there are costs and that I believe in their existence does not contradict my freedom of choice. To the contrary; the fact that there are costs and benefits is what grants meaning to my freedom to choose. Does the existence of moral rules contradict free choice? Does true free choice only exist where there are no rules which state what is right and what is wrong? The opposite is the case: If there is no problem in murdering or stealing, then the choice to act or not to act has no ethical meaning. The significance of this choice derives solely from the fact that it was done within a framework not determined by the individual himself (but rather by social morality). If I decide on my actions and I am also the one who decides how they are to be judged, then such judgment has no meaning. This is the Swiss freedom mentioned above regarding elections. Only metrics which serve as constraints which are not dependent on me can grant my choice ethical meaning. In the following two chapters, we will expand our discussion on morality. To sharpen this distinction, let us elaborate on our terminology. I am distinguishing here between freedom and liberty.44 I will note here that this distinction is related to the famous distinction of Sir Isaiah Berlin, in his lecture which later became an article “Two Concepts of Liberty.” He distinguishes there between negative freedom (freedom from) and positive liberty (freedom to). For the purposes of our discussion, let us define freedom as the lack of any limitations, with liberty being autonomous action within limitations. When an individual acts within a vacuum, he is indeed free. However, when he chooses freely in a place where there are problems (costs and benefits), then he is at liberty. This is a state of liberty, not freedom. Sartre put it very well44: In the torture chamber the prisoner (who has in mind the norm ‘I must not talk’) looks all around him and tries to foresee the tortures in order to discover psycho-physiological means of resisting them. There is no place for the norm. So talking becomes, abruptly, a possibility of the self-aware subject.

This is not simply the triumph of external determining factors: it’s the internal decision to let oneself be determined by external factors –by anti-norms and anti-values. Sartre is describing the dilemma of the prisoner: Should he be at liberty and act autonomously, or should he surrender to circumstances? This liberty exists precisely in circumstances which deny him absolute freedom. Even though these two terms, freedom and liberty, are somewhat interchangeable in our discussion, they are in fact practically opposites. As we saw in the examples of political elections, liberty can only exist where there are limits imposed on us (like in Israel), that is where there is no freedom. When there is freedom (like in Switzerland) — liberty is not possible, as our autonomy is not expressed in an ethically significant manner. Thus, if we speak of terms with an ethical connotation, it is more correct to speak of liberty rather than freedom. In the ethical context, liberty is not the choice to determine that it’s wrong to commit murder. This value is forced on me. My choice is whether to be committed to this value — that is, not murder in practice. When I am the one who also determines the values, and not just my commitment to them or lack thereof, then that is freedom and not liberty. In contrast to the intuition many of us have, the choice of values themselves is a meaningless action. One man chooses the value of not murdering and another chooses the opposite. If there is no objective and obligatory ethical standard outside of individual choice, then how can I judge either one? I judge the first as a good person and the other as wicked only because the value of life is not dependent on individual choice; it is forced on all of us. Thus, the only thing that can be judged ethically is our effective commitment to values, as this is an expression of liberty and not freedom. Why do so many think that freedom is a value? It seems to me that the reason is that the denial of freedom is a moral crime and injustice. When I unjustly imprison someone, I am a criminal. It would seem that people learn of the value of freedom from here. But this is a mistake. This lesson can only teach us that freedom is an asset, but not necessarily a value. When I unjustly deny someone their money, then I am also committing a moral injustice. Does that mean that ownership of money is a value? Of course not. Money is an asset, yet still its denial from someone without cause is a moral injustice. What about liberty? In contrast to freedom, liberty is indeed a value and not just an asset, when it is defined as the ability of the individual to act autonomously and choose his path and conduct within given constraints; not to give himself over to the influences of the environment (i.e., not to surrender to weakness of the will). Such conduct is a very important value. What of the denial of someone’s liberty. In principle, it is almost impossible to deny liberty. At most, one can place limits on it, and thus reduce his freedom. But the fact of someone being at liberty is almost entirely dependent on him, and no-one else. The question which determines his level of liberty is what he will do with it in a situation filled with limitations. There are people who in the most terrible of circumstances — in the Nazi death camps (like Viktor Frankl) or the Soviet Gulags (like Anatoly Sharansky) — who wonderfully expressed their liberty. Precisely because they were in a situation almost devoid of freedom, their liberty could reach a more exalted level. From another angle, we can say that liberty expresses choice with ethical significance (Israel), while freedom represents randomness (Switzerland). Obviously we are not speaking of randomness in the physical sense, as the Swiss voter doesn’t actually toss a coin, but he does do an action with no meaning (and without judgment), and in this sense his action is like a random one. According to this terminology, it is more accurate to say that libertarianism espouses the liberty of the will and not the freedom of the will. Here we once again see the mixture between choice and randomness, which we dealt with above.

Summary: A Phenomenology of the Libertarian and Deterministic Pictures This chapter dealt with the phenomenology of the libertarian picture. Phenomenology is the description of how things are in practice, without getting into theoretical explanations, proofs, arguments in favor and against, or empirical confirmations.45 Here we just dealt with how it’s supposed to work, drawing a map of the phenomena, distinguishing between different kinds of actions and various situations. We saw that there are three kinds of influence on people: genetics, psychology, and the environment. Libertarianism argues that none of these is the same thing as the individual human being himself, but rather constitutes a topographical landscape in which the individual operates. Their significance is the various pressures and urges in favor and against different actions applied to the human being. These pressures create different levels of difficulty for decisions and courses of action (valleys, plains, and mountains at various levels of steepness). Still, the libertarian believes that the individual’s choice is his and his alone, and not that of his environment, external or internal. The individual possess will, and free

choice is solely a decision of the will, not an automatic surrender to impulses. We saw that this free decision-making contains two components: the decision as to what is worthy and desirable to do (is it worthy to yield to pressures or not? Should I climb the mountain or slide into the valley?), and a decision as to how much and at what level of intensity to fight (i.e., the choice to choose; how much energy to use to climb the mountain). The second decision effectively influences our success in implementing the first decision. If we are talking about yielding to the pull of the environment and sliding into the valley, which is easy to do, here there is no need to decide to use energy of course (or even decide to choose). Even without choice, that is even if the individual gives himself over to his psychological environment, in the end that’s what he will effectively do. The significance of the decision in the second stage is primarily at the point where he makes the decision in the first stage to climb the mountain and not surrender to the pressures of the environment. Here, the amount of energy the individual decides to invest as opposed to the steepness of the mountain he wishes to climb is a very important question. If he doesn’t use enough energy, he will fail, and then we are of course dealing with weakness of the will. We also saw that the libertarian also does not believe that every human decision is the result of decision and choice. There are cases in which the individual acts automatically by force of his psychological environment. Such situations can appear where the mountain is too steep (an irresistible urge), or in situations in which the individual chooses not to choose, either based on weakness or inattention (and then it is weakness of the will and he bears ethical and criminal responsibility), or in situations which do not require judgment, that is situations in which we are not dealing with a value dilemma but an action (or decision) which lacks that dimension. The phenomenology of determinism was also described here, albeit very briefly, as it is simpler. According to this approach, everything is dictated by causes and there is no freedom at all. It identifies the topography we spoke of with the individual himself. Therefore the landscape is what determines the decision, and there is therefore no reason to assume that there is anything beyond this materialpsychological environment. The problem of weakness of the will does not exist in the deterministic picture, as after all: Everything is established in advance. There is no room for feeling failure or weakness, and even if they do appear they have no meaning. Even these feelings are themselves predetermined. The determinist sees weakness of the will as an illusion, in the same way he sees free will itself as a mirage (we will encounter many more such “mirages” within the deterministic picture later on). In the next chapter, we will examine the consequences of holding to these two pictures, and afterwards enter into the a priori and a posteriori considerations which will help us decide which one of them is correct. 28 For a clear presentation of this distinction, the reader is directed to read the fifth chapter of Richard Taylor’s Metaphysics. 29 I tend to think that Israeli elections are also not truly a matter of free choice, as in our case, the candidates are largely the same. Therefore even in the actual country of Israel, one could just as easily hold a lottery instead of elections. We are thus back to the illusion of freedom. But here, just for demonstration’s sake, we are dealing with a metaphorical Israel, in which there are essential differences between the candidates, the election of any of one them leading to different circumstances on the ground. 30 There are a number of attempts nowadays to try and create a truly random generator on the basis of quantum theory. However, there is no way to do so based on normal methods of programming. The field of creating random generators requires a great deal of expertise, and it primarily revolves around the ability to create series which are not random but which have characteristics which approach randomness as much as possible. 31 Let us repeat that we saw above that there is no real randomness in the macroscopic world. Use of probabilistic tools does not attest to randomness. Here, we encounter a similar phenomenon. 32 Incitement to idolatry is an exception to this rule, as Jewish law does specifically impose a punishment on the inciter in this case. 33 Here, too, there is an apparent mixing up of predictability and objective determination. What the Torah says is that God knows that this will happen, but it does not state that he decreed it on them. Maimonides apparently understood that the statement to Abraham was a decree on the Egyptians and not a prophecy of the future. 34 “How is Weakness of the Will Possible?,” in: Donald Davidson, Actions and Events, Oxford: Clarendon Press, 1980. 35 This is not a complete solution to the problem, as one could now raise a similar question regarding the decision to choose. If someone wanted to use sufficient energy, then why didn’t he? And if he didn’t want to, then it’s no wonder he failed. Here it’s easier to find a solution, but I will not continue this discussion to avoid infinite regress. 36 We distinguished here between decisions of no consequence, such as pressing a button, and those with ethical dimensions. However, there are intermediate cases: Decisions lacking an ethical dimension (fear of Heaven), but which nevertheless involve judgment. For instance, the decision whether to go eat at restaurant A or B. This is a dilemma unrelated to ethical decisions, yet our hunch still tells us that we also require a free decision by our will, which is also the result of judgment. I therefore preferred the term “decisions requiring judgment” to “ethical decisions.” 37 R. Yitzchak Alfasi, the greatest Jewish religious jurist in the eleventh century. 38 Tana’ic sources not included in the Mishnah. 39 “Perpetual Accuracy and Conflicting Effects of Certainty on Risk-Taking Behaviour,” Nature, 453, 917-920 (12 June, 2008). 40 There is an assumption here that the percentage of sugar is directly related to pleasure from the solution, and that as far as the bees are concerned the pleasures from the various tastings amount to some general enjoyment which is at a higher level. Moreover, there is also an assumption that there is no value in the certainty of the profit but only the overall amount. These are all problematic assumptions but not important for our purposes. 41 This argument will obviously not convince the determinists. As we saw, they will explain that the convincing also has a deterministic effect

on the individual, and he therefore chooses differently. But that’s not my goal here. My aim is only to point out the libertarian option for interpreting such experiments. 42 Decisions and Mechanical Actions 43 I know very well that these terms are used differently than here, and sometimes indeed used in the reverse. I am not making a semantic claim; the question which term is interpreted in what way isn’t really important to me. What’s important to me is the very distinction between the terms, and the terminology we use to do so is unimportant. We will therefore address the terms of freedom and liberty as they appear here, as definitions of the relevant terms, solely for the sake of the convenience of differentiation. 44 Jean-Paul Sartre, Michel Rybalka (editor), Michel Contat (editor), Richard C. McCleary (translation), The Writings of Jean-Paul Sartre Volume 2: Selected Prose, Evanston: Northwestern University Press, 1985, p. 246. 45 Scientists distinguish between phenomenological theories, which focus on describing phenomena, and essential theories, which offer an explanation for phenomena (and usually include theoretical entities, such as particles, forces, and fields).

Chapter Three ON MORALITY AND FATALISM We live in a world where morality is good for preachers. We don’t live based on it. Even the Jews have ceased to live based on it. We will not prove to the Yishuv [Jewish community in the Land of Israel] — if we cannot prove that there is a political disaster — moral words alone will not suffice. (David Ben Gurion, Herev Hayonah) “Freedom is a fate which man cannot repeal, man has no choice but to be free.” (Jean-Paul Sartre) “Freedom is a hard habit to break.” (Robert Heinlein, Citizen of the Galaxy)

In the previous chapter, we presented the two pictures opposite each other: The determinist and the libertarian. In this chapter we will examine the consequences of the two positions. The immediate consequence in this context is the question of moral evaluation, judgment, and responsibility. Another important consequence is the question of fatalism.

Kantian Morality and Taoist Morality American logician Raymond Smullyan also indulges in popular logical writing, including logical stand-up acts. In his book The Tao is Silent,46 in a chapter entitled “Is God Taoist,” he brings an entertaining dialogue between man and God, in which man asks God to remove his free will:47 Mortal: And therefore, O God, I pray thee, if thou hast one ounce of mercy for this thy suffering creature, absolve me of having to have free will! God: You reject the greatest gift I have given thee? Mortal: How can you call that which was forced on me a gift? I have free will, but not of my own choice. I have never freely chosen to have free will. I have to have free will, whether I like it or not! God: Why would you wish not to have free will? Mortal: Because free will means moral responsibility, and moral responsibility is more than I can bear! Man in this dialogue can no longer deal with the moral responsibility imposed on him by free will, and he therefore wishes to be rid of it. That gift we received (against our will) also gives us the possibility to be masters of our own fate, but it also imposes a substantial commitment and responsibility for our actions. Moral (and legal or religious) responsibility is the product of the assumption that we have the possibility of controlling our actions — i.e., that we have free will. This is the main consequence of the discussion of freedom of the will. Later in the dialogue, God tricks man in a fairly roundabout way, in a series of proposals meant to figure out what he really wants (all the while asking man every so often if he really wants the possibility of wanting taken from him, as then he can’t even want that). He asks him what he thinks of being left with free will but being freed of the associated moral responsibility. It turns out that man is terrified of this possibility, as he fears that this will lead him to commit great sins. Then God offers him wholesale forgiveness in advance for all his sins and an assured place in Heaven. Man is of course even more terrified by this option, as this will certainly lead him to sin, as all the obstacles to sin (law and sanctions) are removed. It turns out to him and to us that he actually does not want to sin, even without the consequences of punishment (which of course makes us seriously question his initial request). Then God asks him: Why does he sin if he doesn’t want to? Ostensibly, if he does this than this is precisely what he

wants, and without freedom of will, he can’t do it. This is the question of weakness of the will which we discussed in the previous chapter. The entire dialogue aims for a far-reaching argument: In effect, free will isn’t necessary to ensure moral conduct. Smullyan is trying to tell us that we are stuck in the Kantian framework, according to which moral conduct is in essence the conduct of overcoming — overcoming of inclinations and tendencies, and which is founded in obeying the law, the imperative (the categorical imperative, in Kantian terms) or the sanction which accompanies them. By contrast, Smullyan tells us, God himself is actually a Taoist — he believes that we should act naturally, and then the whole world will be much better. Without threats and without imperatives, without punishments and without sanctions, there is no need to overcome anything. We need to simply listen to our heart and “go with the flow.” According to Kant, by contrast, the moral act is considered moral if and only if it is done consciously and based on a decision to act based on the values of the good and the moral. Just doing good, Kant tells us, is not a moral act. Animals do the same, and maybe even inanimate objects. Sheep do not murder each other, and also don’t forcibly take anything from them. They also don’t libel each other. Would we call this moral? Probably not, because the sheep are simply following their nature rather than making a choice. The moral virtue of man is his autonomy and his ability to decide what he will do and how he will act. This alone is what can define him as a moral (or immoral) actor. Let’s sharpen the differences between these moral approaches some more. We could view the human ability to be autonomous and make decisions freely and independently (without necessarily being dependent on the given psychological circumstances) as solely a tool, a tool whose purpose is to ensure that we do not surrender to our evil inclination and do bad things. According to this view, the goal is the character and content of the actions we will do (and primarily their results), and our autonomy is nothing more than a tool meant to ensure that we overcome our inclination and do what’s right. If we are not autonomous, we are given over to our inclinations, which from a Kantian-religious perspective, very pessimistic at its heart, are always evil inclinations. That’s how Smullyan presents it in his dialogue as shown above. But Kantian ethics argues that this picture is incorrect. Autonomy is not just a tool whose goal is to ensure we act properly and prevent us from surrendering to the evil inclination. To Kant, autonomy is a value in and of itself, and it is man’s moral goal in itself. A hypnotized human being, even if he only does good deeds, helps all the needy and doesn’t harm a fly, is not seen as a moral person. In some sense he isn’t even a human being, as he doesn’t autonomously choose his actions, just like the aforementioned sheep. Even if the tendency to good deeds is embedded in the person and he does so because he feels good about it, that doesn’t amount to moral behavior. Only an individual who decides on his own how to act, and is not automatically driven by his internal tendencies, can be a moral individual. Thus, moral judgment according to Kant is determined not only based on the character and content of our actions, but more importantly by the manner in which they are done — with our decision to do them and the extent to which these actions express our autonomy through our motivations and mode of action. In other words, morality is not just a definition of proper action, but is also an expression of commitment to some imperative and conduct by force of commitment thereto. Smullyan’s Taoism proposes an entirely different position. What is expected of the person is to do good, specifically to do good deeds, and no more — not to cause suffering, not to harm others, to be a good person. It doesn’t really matter if he’s hypnotized or not, if there are imperatives or not, if he chooses nor not, if his free will rises above his inherent tendencies or not — what matters is only the result, what he did in practice. Smullyan sees autonomy as at most a tool for dealing with inclinations, but certainly not an independent value. The goal is simply that the world be better and more comfortable. That’s it. Moreover, Smullyan’s Taoism assumes something else: Man is by nature good, and therefore if he acts naturally it will be good. Contrary to Kant’s pessimistic approach, which sees Man as wicked at heart, Smullyan argues that there is no need to overcome inclinations, as they are essentially good. He therefore argues that there is no need for imperatives, threats, or punishments, and perhaps not even free will and autonomy as moral tools. It’s important to understand that the debate revolves primarily around the first point. Kant does dispute the second point as well, arguing (apparently in the wake of common Judeo-Christian approaches) that “man’s inclination is wicked from youth,” but even if he agreed to the optimistic Taoist approach that man is essentially good, he would still dispute the question of moral estimation. He ethically evaluates man’s actions according to the degree to which they are expressions of autonomy. Even if he was essentially good and acted exclusively according to his good inclinations and was a model human being, that still does not

necessarily mean that he should be seen as good in the ethical sense. To merit such esteem, he must act autonomously and do good because it is good, and not because he feels like it or because they are his inborn tendencies. Man is not a sheep and is not supposed to be a sheep. Smullyan goes one step further. In his book, he constantly repeats the message that all the evil in the world is in fact a consequence of the artificial and frozen Kantian approach with its fundamental pessimism. This is an approach which convinces all of us that we are wicked by nature, and therefore we do moral acts based on fear, almost compulsion. By contrast, if we were more optimistic and assumed our nature was good, we would do so naturally, without constantly suspecting ourselves of foreign motives (because motives do not matter). The world would then be a much better place. But Smullyan is begging the question with this argument. According to him, the goodness in the world is determined according to the mount of good that is done in it. He therefore measures both approaches according to this metric (the extent to which they lead to the doing of good). He doesn’t notice that Kant disputes the validity of that same moral metric. Even if he is right (and I am not at all sure he is), and the world would look better if it were based on natural Taoist conduct, this would still not be a more moral world in Kant’s eyes. At most, it would be a more comfortable and enjoyable one. We can easily see that these two opinions are a reflection of the pictures we described in the previous chapter. Kantian morality argues that human choice is a decision not dependent on his natural tendencies. Man’s uniqueness lies in his ability to detach himself from them and decide against them when need be. This does not mean that Kant thinks that natural inclinations are always bad. Absolutely not. We certainly have a strong tendency to do good, as an individual who helps out his fellow man feels very good after doing so. But a human being is supposed to consider his decisions and not blindly follow inclinations. If they are good — he follows them (sliding into the valley). If not — then he acts against them (climbing the mountain). What matters is man’s control and autonomy. By contrast, Smullyan offers the alternative of natural conduct. For him, a deterministic world in which man makes decisions according to his tendencies (this is precisely what compatibilism is), is a perfect world.48

The “Categorical Imperative”, Hillel the Elder, and Arbinka Immanuel Kant was a German philosopher from the eighteenth century. Many see him as the greatest philosopher of the modern age and one of the greatest philosophers of all time. Three of his books are devoted to deciphering the elements of ethics. Kant, per his consistent method, searches for a rational basis for moral obligation, as well as a clear definition of that obligation. His ultimate conclusion is that the foundation of morality is the categorical imperative — i.e., absolute and unconditional — as opposed to the hypothetical, imperative. The content of this imperative, according to one of the formulations, is “do your actions only on that practical rule which, in accepting it, you would wish to see become a general one.”49 People usually understand that the meaning of the Kantian imperative is that if I don’t act correctly then others will also not act correctly, and then the world will look bad. It is therefore forbidden to steal, as if I steal, so will others. Hillel the Elder said in the Babylonian Tractate of Shabbat (Shabbat 31a): “What is hateful to you, do not do to your fellow man.” This is how a parent teaches their child: “Would you like it if someone did that to you?” meaning that if I do it to him, he will also do it to me. This is ostensibly what Kant meant, requiring that we act in such a way that everyone else would. But that is definitely not what Kant meant regarding his imperative. That interpretation expresses a utilitarian bent, and Kant believes that utility and consequences have nothing to do with morality. Kant argues that the question of whether I want an action to become a general rule is a criterion for defining the action as moral. Even when my actions do not influence other people’s behavior or affect the world at all, I must still act in the manner I would like everyone to act. The morality of my act is not dependent on my expectations regarding the resulting future, specifically not what I think others will do. Take the obligation to pay income tax, for instance. We will use Israeli satirist Efraim Kishon’s literary alter ego of Arbinka for a demonstration. If Kishon decided whether to hide a thousand shekel from the tax authorities, he would certainly have a dialogue between himself (= his good inclination) and Arbinka, who is known for his criminal tendencies (sadly, I do not possess Kishon’s literary and comical gifts): Arbinka tempts him: “If you didn’t write this book, the state would not look different, and in such a situation no-one would demand you pay that thousand shekel. So you must treat it as though you didn’t do the work at all. Does it bother anyone that you have another 5,000 shekel? Besides, the state will not even notice such a sum. What’s a thousand shekel when it comes to the state treasury? There is no service the country provides which will not be given to anyone because a thousand shekel is missing. So who’s it harming?” Kishon doesn’t miss a beat, responding in a strident, moralizing tone: “Yes, but if everyone did that, how

would the state treasury look I would after all not want it to become a general rule!” But Arbinka offers a very simple solution: “If so, you must hide your taxes in such a way that no-one will learn from you and no-one will know about it. In such a situation, the state treasury will be safe.” But Kishon, the innocent immigrant from Hungary, a loyal and honest citizen, does not give up: “But everyone can say that to themselves: ‘I will do it without letting anyone know,’ and once again we end up with an empty treasury.” Arbinka keeps it up: “Not true, after all — they’ll all do what they want to do, anyway. Even if there are people who hide taxes, that will not change if you decide to hide your taxes or not. Your hidden actions have no effect on the state treasury, as you will certainly reveal this to no-one.” Kishon can still debate with Arbinka forever, as even if it’s true, of course, every individual can still tell themselves that they will continue to hide taxes. On the other hand, Arbinka is certainly correct (Arbinka is always right) in arguing that what Kishon decides will have no effect on the decisions of others, nor will it have an effect on the treasury. So why not hide taxes? The same is true when it comes to voting in elections. A man can come to the conclusion that there is no point to go out and vote, as one vote has no effect (the odds that any given party will be dependent on a single vote are close to zero. To the best of my knowledge, it has never happened, and can almost not happen). Therefore, the individual person has no real reason to go out and vote. But what happens if everyone follow suit? Things will be bad indeed. But their decision is not dependent on his. He won’t vote and won’t tell anyone, so whatever others do will be done in any event. It’s not dependent on his choice, and once again the individual’s vote has no significance. So is there really an obligation to go out and vote? If so, why? Kishon’s good inclination is a Taoist, and he therefore makes use of utilitarian-consequentialist ideas. He tried hard to convince Arbinka that the action he is proposing has problematic consequences, and therefore the action is immoral. In the example we provided here, we can see that this argument also fails in practice (because this action doesn’t have actual negative consequences). However, the Kantian is repulsed by such arguments. In his eyes, the action is not moral because even Arbinka himself would not want it to become a general rule. The categorical imperative is a criterion which examines actions irrespective of the future. Kant is not arguing that there is really any fear that Kishon’s actions will become a general rule and lead to the emptying of the state treasury, but only that if it became a general rule, then things wouldn’t be good. If that is the case, then such an action should be treated as immoral and be avoided. Kant does not measure morality based on the consequence but on the manner in which the action is carried out — i.e., the degree of the person’s autonomy (as he puts it: action based on respect for the rule or the imperative). In his opinion, one should pay tax or vote not for fear of any consequences, but this is an unconditional obligation. It is a moral duty because if the custom did not become a general rule (a hypothetical one, even though this won’t happen in practice), the world would be the worse for it. It seems that anyone who feels that voting in elections or properly paying taxes despite Arbinka’s correct arguments, is important, cannot avoid recognizing the categorical imperative as a necessary basis for moral conduct. There is no other basis which successfully replies to Arbinka’s correct arguments and establishes the existence of such obligations as moral conduct.

Moral Responsibility So far, we have dealt with the moral judgment of people and the moral evaluation of actions. But the important question for our purposes is the question of moral responsibility. The dialogue we brought at the beginning of the chapter starts with a request which assumes Kantian moral theory that the moral (and legal) responsibility imposed on people is conditional on his acting freely. Put differently, free will is the basis for moral responsibility. If there is any other factor, external or internal, which determines the individual’s decision and conduct, then there is no room for imposing moral or legal responsibility on him. It is for this reason that we don’t impose moral responsibility on animals, inanimate objects, or natural disasters. After all, their “actions” are not the result of judgment and free choice. Smullyan may perhaps be offering us a possible outline for good behavior, but he did not offer a basis for moral judgment, condemnation of immorality, or the moral responsibility of the individual for his actions. According to him, even when a person acts immorally, there is no room for taking him to task. The character of so-and-so is good and that of such-and-such is bad, and everyone acts according to their character. In other words, moral condemnation and judgment assume the existence of an objective imperative according to which everyone is judged, as well as the freedom of the man being judged to decide whether to follow the imperative or not. In a deterministic picture, there does not seem to be any room for imposing moral responsibility on

someone. In this picture, human beings are like animals or inanimate objects. They may have reason, consciousness, emotions, and so on, which do not seem to exist among animals in the same manner, but moral responsibility can only be imposed on those who choose freely. All the traits enumerated above are not relevant to this question. Even if animals had consciousness, so long as they are compelled to act the way they do, we still cannot impose moral responsibility on them. At the same time, if the individual acted as he acted simply because it was in his nature and that the circumstances surrounding him led him to it, then there is no possibility of judging him for his actions. Thus, moral responsibility can only exist in the libertarian world.

How Can We Judge? A reflection of the problems presented above is the argument often raised against the determinist: How would you judge a man for his actions? How can we appoint judges to punish criminals for their actions if they are compelled to do what they did? Does this world view picture really allow for the justification of punishment? The determinist can respond at three levels. First, he could say that such judgment is indeed unnecessary and there is no need to have use of it. The fact that society does so is an unjustified anachronism. But more sophisticated determinists have a stronger argument, which leaves our justice system in place. At the second level, the argument is that the judges and society are also subject to these deterministic constraints. Much like the criminal is compelled to commit crimes, the judge is also compelled to do what he must (and society is for its part compelled to appoint judges). After all, the judge is also a flesh and blood human being, and he is therefore also driven by those same “other” factors which drive everyone. His actions are also a result of causal determinism. At the third level, the determinist will argue that punishment is indeed justified, but here too, the justification is consequentialist. If we punish someone, he may not “deserve” it, but it will achieve the desired results. The odds that this person will not commit crime anymore are now higher, and that is now part of his calculus when considering committing a crime in the future. This of course leads us back to Smullyan’s Taoism, which pegs morality to practical consequences. Logical positivism is one of the most prominent supporters of the third-level argument. Its main focus is the denial of metaphysics and the clarification of the philosophical, logical, and scientific discourse. As part of that estrangement from the metaphysical and the non-factual, it sees values as also a sort of mysticism. It therefore bases punishment and moral condemnation on facts, results, and benefits. In the main book of Moritz Shlick50 on this subject, The Problems of Ethics, he proposes viewing punishment as a shaping factor. Shlick argues that punishment is not imposed on the criminal because he “deserves” it, but to change him (remember the drill fixing example). The change was not formed by the individual’s decision; it is rather the automatic result of the punishment (and perhaps also the fact that the punishment appears on the law books, deterring potential criminals). A famous Roman story will demonstrate the point. It tells of a master about to whip his slave for an infraction. The slave tells the master: “I am not guilty. It was written in the stars that I should sin.” The master immediately replies: “Indeed. And it was also written there that I should whip you.” The determinist and the compatibilist shouldn’t see this story as a joke; this is precisely how they see reality. So, the third-level argument for defending punishment in a deterministic world sees punishment not as an educational sanction but as an act of conditioning meant to serve as an additional factor shaping the human psyche. The goal of law and punishment is to create a different environment, and now punishment is among the factors which dictate human action, leading more people to act in accordance with the law.51 This is the meaning given by the determinist (or the compatibilist) to moral responsibility. By its lights, the punishment of a criminal is much the same as fixing a drill or a car. Moreover, according to the determinist, it is worthwhile to punish a person even if he was not morally guilty for his crime, if only to “fix” him (this was known as “re-education” in the Soviet Gulags, and this was also an outgrowth of determinism and dialectical materialism). But it’s important to understand that such a justification of punishment and judgment can indeed justify the practical actions of judgment and imposition of punishment in the practical sense, but this picture clearly has no room for concepts like moral responsibility and certainly not moral condemnation and praise. The determinist effectively argues that the feeling of responsibility and the experience of imposing responsibility on any individual is effectively based on an illusion. Of course, these illusions are positive and beneficial, as they shape us in a way which makes it difficult to commit crimes, but they have no real justification. The criminal is not “deserving” of punishment or condemnation, as he has no criminal or moral

responsibility for his actions. Moral responsibility is but another illusion in a series of illusions discerned by the determinist in the human experience.

Interim Summary: The Question of Moral Responsibility The conclusion is that there is in fact no practical consequence for the question of moral responsibility in the debate between determinism and libertarianism. The two approaches support judgment and punishment, but they see these actions in different ways. The difference is not in the practical imposition of morality but the theory at its base: Is it proper to feel condemnation towards an individual who behaved immorally, or praise towards someone who acted morally? Libertarianism sees these actions as the imposition of moral responsibility and punishment as an educational tool and deterrent in the commonly accepted sense. They educate and influence, and these contribute to the creation of a better world in the Kantian sense of the categorical imperative, not necessarily in the consequentialist-taoist sense. By contrast, determinism sees these actions as part of the universe’s causal change. They are done due to causes which force all people to act in this manner and treat actions and people in various ways. Moreover, as we saw above, the determinist even finds a way to justify these coerced actions, as they shape our world in a way which will lead to more good, in the Taoist sense, of course, not the Kantian. This justification may not really be important, as even if it wasn’t justified, we could not act or think otherwise in the deterministic world. We will return to deal with other angles of the question of norms and moral condemnation and judgment in chapters fourteen and fifteen, where we will examine it from the perspective of neuroscience.

Fatalism Another consequence of the debate between libertarianism and determinism lies in the field of fatalism. Fatalism is a position arguing that anything that happens had to happen and we have no influence over it. Fatalism appears quite a bit in literature, and one of its clearest expressions is the story of The Appointment in Samarra by John O’Hara: Death: There was a merchant in Bagdad who sent his servant to market to buy provisions and in a little while the servant came back, white and trembling, and said, ‘Master, just now when I was in the marketplace I was jostled by a woman in the crowd and when I turned I saw it was Death that jostled me. She looked at me and made a threatening gesture, now, lend me your horse, and I will ride away from this city and avoid my fate. I will go to Samarra and there Death will not find me.’ The merchant lent him his horse, and the servant mounted it, and he dug his spurs in its flanks and as fast as the horse could gallop he went. Then the merchant went down to the marketplace and he saw me standing in the crowd and he came to me and said, ‘Why did you make a threating gesture to my servant when you saw him this morning?’ ‘That was not a threatening gesture, I said, it was only a start of surprise. I was astonished to see him in Bagdad, for I had an appointment with him tonight in Samarra.’ This story is a demonstration of fatalism, according to which man cannot escape his fate. Fatalism is usually tied to death. As this is an event no-one can escape.52 If so, fatalism would seem to be nothing more than a synonym for determinism. But fatalism can exist in a non-deterministic world, as well. A man can tell himself that even if we have a degree of choice — in the end, “what needs to happen will happen” (every bullet has a name on it). The fatalist argues that in the long term, all the things which are important in our lives are dictated to us and are inescapable. Another difference between fatalism and determinism is that fatalism is a psychological position and not a philosophical worldview. Thus, on the one hand, a libertarian position is no guarantee against a fatalist mindset. On the other hand, a deterministic position could be accompanied with happiness and joy and the feeling of complete autonomy, as though everything is up to me. So does the individual live his life, even though he knows at the intellectual level that everything is in fact predetermined. He makes decisions, ponders how to act and makes long term plans — all this as part of the imaginary life of the determinist, as described above. Most of the determinists I know are not mired in a fatalistic funk. Luckily for them, they seem to be fated to be happy and cheerful, at least in part. Moreover, sometimes fatalism is even comforting. When something terrible happens to someone, he tends to console himself that this would have happened even if he had chosen another path and that it couldn’t be avoided. It appears that this is how the Druze find comfort in the loss of their loved ones. Here

fatalism leads to making peace with the situation, and sometimes even unusual bravery (after all, every bullet has an address, and my actions will not change this). Sophocles’ play Oedipus demonstrates this very well.53 Oedipus may indeed be a clear representative of fatalism, but as Dorothea Crock (winner of the Israel Prize for the Humanities, 1973) noted in her book The Elements of Tragedy (Hebrew), it’s interesting that even though fate imposed this on him, Oedipus takes responsibility for his actions, and punishes himself (gouging his eyes out). She sees this as a parable for our own lives, where even though everything is predetermined, we live as though we bear the responsibility, and impose moral and legal blame on people who sin. But why must we take responsibility for our fate if everything is predetermined? After all, there is no room for moral responsibility in the deterministic picture. If I know that neither I nor anyone else has responsibility for my actions, how can I have true feelings of moral responsibility, or condemn or praise others for moral conduct? The demand is simply absurd. Can or should a man gouge his own eyes out for an action for which he was not even responsible? Is such conduct worthy of praise? Such behavior should be seen by the determinist as the instinctive action of a man who lost his sanity due to grief. It has no moral significance beyond this. Moreover, the acceptance of responsibility is also forced on the individual, and it is therefore also unworthy of praise. To the best of my understanding, all those who act and speak this way do not really believe in determinism. It’s just that they themselves have not found an adequate response to the arguments of the determinist. They therefore console themselves with literary fantasies of taking moral responsibility in a world in which this concept has no meaning. My book is meant, among other things, for such people — people who clearly see determinism as folly, but who see the arguments for determinism as convincing. My main aim is to offer them a respectable way out, an alternative worldview which fits our basic intuitions much better, and show them that this picture does not contradict any scientific finding. Despite what I said here, it would seem to be correct that fatalism as a psychological state of mind is one of the possible consequences of the debate between determinism and libertarianism. Ostensibly, the determinist should be a fatalist, treating whatever happens to him with equanimity, in the past and future, and as we saw also moral outrages. He should see it as he sees fish in the aquarium. Still, he is nevertheless compelled to feel otherwise despite the injustice, and the determinist therefore sees no difficulty involved. Another possible consequence of fatalistic determinism we would expect is indifference and lack of initiative. We would expect the determinist not to work too hard to achieve any goal. After all, his assumption is that it will happen or not regardless of his desires and actions, and there is therefore no point in investing any effort. A demonstration of this consequence is what is known as the “idle argument.” Say that I have a test tomorrow. A smart person with unlimited predicting capabilities could write my expected grade on a notepad. Let us assume that there is a complete match between what the note says and the grade I will actually get. If so, then there is no point in studying for the test, as my grade has already been established.54 It is true that if I don’t study, my grade will be appropriately lower, except that the decision whether or not to study is not really in my hands. Any judgment is wrong, and therefore human beings should not be worried by such questions. Are determinists truly always fatalists, good-natured bums who never reel from their blows (or to the contrary — mired in a funk)? As I wrote, certainly not. In response to the question why this is, they are expected to answer with the calm deterministic answer: Apparently we’re just not built for it. After all, their reference to events is also an event in the world, and as such determined by independent factors. There is therefore no point in asking them why they are not fatalists. Thus, fatalism is also not a practical consequence of the debate between the two positions.

Interim Summary All the consequences of the debate are effectively written off by the determinist in the same manner: Yes, there’s no justification for moral condemnation and no justification for my not being a fatalist, and there really is no significance to moral responsibility, but that’s just who I am. One cannot dispute the facts. This means that the feelings of the determinist are built in the same way as anyone else, but he does not see them as a reflection of any objective truth, but simply the necessary and unavoidable consequence of the various factors which shape it. The gap between them and reality, when it exists, is a fact and needs no explanation. The conclusion is that the difference between the two positions is not at the practical but at the theoretical level. It lies in the question of how much we should see feelings of this sort or another as

reflecting any sort of truth, and whether we should simply see our own intuitions as illusions. The determinist sees our (and his) most fundamental intuitions as illusions, he treats his own feelings and intuitions as something which is the process of a blind, undirected process, and he therefore sees no problem in them not reflecting the truth. He also has a feeling of freedom, he also acts as though he can decide morally and as though human beings have moral responsibility (he might even gouge out his own eyes after a sin he committed), but to him this is no more than an illusion. Why is his deterministic conclusion not an illusion itself? We will expand on this point in the next chapter. It’s important for me to clarify that all the considerations we presented here are no proof against determinism. These are the consequences of the deterministic position. While it’s true that our intuition tends toward the libertarian position, the position of the determinist remains consistent, even if it clashes with intuition.

Third Intermezzo: Rav Hisdai Crescas and Theological Determinism An interesting demonstration of this point can be seen with the approach of Rav Hisdai Crescas on the question of freedom of choice. In the tenth chapter of his book Or Hashem, he presents an original approach in Jewish religious thought. According to him, the world is determinist (also based on a religioustheological consideration, because God in his omniscience knows everything in advance), and as such, everything we do in the world is dictated in advance. What is given to us is only the decision how to address what is happening. If we treat sins as bad then we chose good, and if we treat them positively (desire them), then we chose bad.55 It seems to me that there is a fundamental conceptual error here, which is easier to notice in the world of neuroscience. Our reference to events and things is also a sort of event in the world (as we will see later on, there is also room for seeing it as a physical event). If so, God’s omnipotence should predict this event as well; he should know in advance what is happening to us. Thus, we still don’t have a solution to the question of the choice the Torah gives us (“And you shall choose life”). Is seems that Crescas erred here. He sees our reference to events as something which is outside the world of events. But that is itself an event, mental or physical, like any other event. Therefore the same problems that apply to any other event apply here. If God knows everything in advance, He must also know this. If his knowledge contradicts my ability to choose, then it also prevents me from independently forming a position on matters. This is the equivalent of the determinist’s answer to all the questions, as we saw above: My reference to events is also an event, and it is therefore predetermined by various deterministic factors. There is therefore no need to explain or justify them.

So Who’s More Moral? In Gazzaniga’s book (p. 115), there are a number of experiments mentioned which show that belief in free will improves our morality. For example, psychologists Vohs and Schooler conducted an experiment in which they gave a group of students a section from the book of Francis Crick The Astonishing Hypothesis which has a deterministic character, and to another group they gave a section from another book expressing a libertarian point of view. The two groups of students then took a computerized test, and they were told that there is a malfunction in the computer and that answers appear automatically. To prevent this, they were asked to push one of the buttons on the computer during the test, which of course made it very easy for them to cheat. The group which read Crick’s book didn’t do as asked and many of them cheated on the test, while the second group acted better. In a further study by psychologists Baumeister, Masicampo, and Dewall, it was also found that the group which tended more towards determinism also tended towards more aggressive behavior and as a rule avoided helping others. These experiments ostensibly point to libertarian society being more moral and determinists acting less moral. However, as I already noted, to make a moral judgment here would be a bit hasty. One can perhaps say that one of the groups acted in a manner which was less beneficial or sympathetic, but the degree of morality in human conduct is also a function of influences acting on people and their motivations. Second, it seems to me that these results don’t prove themselves in actual life. Determinist-materialist societies (such as atheist societies) can certainly be moral, while there are libertarian societies (say, religious ones) which act less morally. Beyond this, we are speaking here of consciousness and not a world view. We already saw that even people who are clearly determinists in their world view live with a libertarian approach. It may be that their conduct will be moral because of their welcome approach, even if their philosophical views do not align with that.56 So, the determinist position might lead to apathy. As we saw, this is not a necessary result, and there are quite a few determinists who are not fatalistic or idle, but it’s certainly reasonable to expect a deterministic

approach to lead many others to become indifferent or fatalists. In addition, the same fear exists when it comes to immoral behavior. As I already noted, I don’t think that people with deterministic views act morally worse than others in practice, but there is a fear that at least some people might be affected by it, especially if it becomes buttressed by scientific evidence and becomes more certain and convincing. There will be those who will not bother to fight their evil inclination because they see it as forced on them and there is therefore no reason (or possibility) to do something else. It is as if they are saying: “That’s who I am, what can I do?” However, even if these arguments are entirely correct (and I think that’s how things actually are), it’s entirely irrelevant to our discussion. My argument here isn’t that atheist society is more evil (i.e., has more murderers and thieves), but that it is less rational and consistent. The fact that it maintains rules of morality and uses Kantian ethical terminology (of accusation and condemnation, against the “wicked” and not just the “harmful”), is not an explanation for its own approach. It is therefore clear that this is an expression of one of two things: Hidden belief, or an irrational illusion which survived from the era in which this society was still religious. Those who want to have a consistent approach in these societies should agree that their moral judgments are mere illusions. In addition, he will probably not have an answer to the question why they were created, as they don’t really help his survival. In other words, the very existence of moral norms and moral judgments even in these societies is itself an argument against their point of view. 46 Raymond Smullyan, The Tao is Silent, Harper, 2013. 47 It’s highly recommended to read the entire dialogue there. Even though I think it contains a number of errors, it teaches us a great deal about our subjects, from different angles. I should also note that this dialogue was included as one of the articles in the wonderful collection of Hofstadter and Dennet: Douglas R. Hofstadter and Daniel C. Dennet, The Mind’s I: Fantasies and Reflections on Self & Soul, Bantam Books, 1982. The book contains quite a number of articles and stories written by thinkers, scientists, and various authors dealing with body and soul questions from different angles and in different ways. This entire book is highly recommended for anyone interested in these questions. 48 I do not know if Smullyan also thinks that this is actually how the world works and that man has no free choice. That is a slightly different question, as Smullyan speaks of moral evaluation — the ethical question — and not what lies within the human heart, the metaphysical question. 49 Immanuel Kant, Groundwork of the Metaphysics of Morals, Edited and translated by Allen W. Wood, New Haven and London: Yale University Press, 2002.. 50 A physicist and philosopher, student of the famous physicist Max Planck, who founded the Vienna Circle, and specifically the intellectual movement known as logical positivism. 51 How was a world created in which punishments are imposed to achieve a more protected world? The determinist believes that we don’t choose that either, and that this is a necessary consequence of our environment and the laws of nature (apparently due to evolutionary processes). 52 A more modern version of the story can be found in Richard Taylor’s Metaphysics in the chapter dealing with fatalism. He tells the story of Osmo, which is very similar to the appointment in Samarra (though here a plane replaces the horse). My friend, Justice Menachem Finkelstein, directed my attention to a far older version of the story in the Babylonian Talmud, in Tractate Sukkah 53a. 53 Oedipus was the sun of Laius, King of Thebes. The Oracle of Delphi prophesied that he would kill his father and marry his mother, Jocasta. Laius heard of this and wished to kill his son, who was saved by the King of Corinth, who adopted him as a son. When Oedipus learns of the fate the Oracle prophesied for him, he escapes the King of Corinth, convinced he is his father. On the road, he kills Laius, without knowing that he is his real father, and after he solves the riddle of the Sphinx which threatens Thebes, he marries Jocasta his mother (without knowing this fact). At the end of the play, Oedipus learns the true identity of his mother and father, gouges his eyes out, and goes into exile. 54 Newcomb’s paradox deals with precisely the opposite problem. See in the next chapter on logical determinism. 55 Crescas also justifies his determinism based on a causal view of the world and not just due to theological argument. Here, I address only the second argument, as my aim is to contrast it with the causal argument. 56 This is a possible explanation for the question of why evolution would create such a false consciousness. The remaining question is of course why evolution created such creatures, for whom false consciousness would have survival value. Would it not have been simpler to simply create creatures who act in an entirely survival-based manner without fooling themselves?! This is the question of the final explanation (in Dawkins’ terms), and it is discussed in my book God Plays Dice (Hebrew), p. 261 and onward. See also the next chapter, in the discussion on Plantinga’s argument.

Chapter Four DETERMINISM: A PRIORI CONSIDERATIONS Experience is a flashlight which merely illuminates the path we already took. – Confucius “Why would you want to join the navy when you can be a pirate?” – Steve Jobs57

The goal of this book is to examine the question of determinism and free will in light of the findings of modern science. By the very nature of things, a scientific discussion is a posteriori, based as it is on observation and facts. However, before we go into the new and more updated formulas of the dilemma, we must discuss the issue at its a priori level, the philosophical-conceptual level which is prior to any experiment or scientific observation. The first part of the present chapter is dedicated to examining the a priori arguments in favor of determinism, and the second part is dedicated to a priori arguments against it. 57 This chapter was written around the time Jobs passed away.

A. A Priori Arguments For Determinism The Dilemma Argument Dilemma arguments in logic are ones which draw a map of two possibilities, A and B, proving that some conclusion is correct whether we choose A or B. For instance: There is no point in giving tests, as if the student is diligent and studies properly, he will do so without a test. If he is lazy and doesn’t study, however, he will still be the same way even if we have tests. What’s the fallacy in this argument? Obviously, it’s the assumption that we have only two polar opposites as options: Either pathologically lazy students or incorrigibly diligent students. The argument ignores a consequential number of students who do not belong to either of these two extreme categories. These students will learn little without a test, and will certainly learn more if there is a test. When it comes to them, there is certainly logic in having tests. Dilemma arguments, even though many of them appear prima facie convincing, therefore always hide some dichotomous assumption which requires further examination. In many cases, the fallacy lies in that hidden assumption, which is usually not presented in the argument itself (as in the example we brought, where the assumption is that there are only two types of students). For our purposes, the most common argument against libertarianism is the following dilemma argument: Let’s assume that I am now at moment t and face two options: Do X or do Y (Y can also be not doing X). Libertarianism argues that the two options are open to me, and I freely choose one of them. Let us assume that I chose Y. In principle, God could rewind this “movie” backward, return me and the world to moment t, and place me again at the same fork in the road, and there may be situations in which I will decide to do X, instead. But the state of the world up to moment t is identical in both situations. What, therefore led me to choose X some of the time and Y in others? If there was something different in the world or in my psyche, then there are some circumstances which led to my choice, which means the choice was not truly free. If the situation in both cases is indeed identical, yet two different results emerge from the same situation itself, then my decision was done without a cause and was therefore random. I thus face a dilemma: whether there was a previous reason or there wasn’t, my decision was not the result of free choice — we are either dealing here with randomness or a deterministic state. This is the dilemma argument as presented by French philosopher Peter Van Inwagen, among others.58 Where is the fallacy in this argument? As expected from an argument formulated as a dilemma, the fallacy lies in its assumed dichotomy, which sees only two options: causal or random. The assumption of Van Inwagen is that if there is no difference in the situation at moment t, and the individual can nevertheless make different choices at different opportunities, then we are necessarily dealing with randomness. But this is a mistake, or at the very least begging the question. As we already saw in chapter two, while it’s true that such a thought experiment can lead us to the conclusion that the process is not deterministic (i.e., not necessarily determined by the conditions present at moment t), but according to the libertarian’s way of thinking, indeterminism is divided into two categories: Choice and randomness. Thus, while we showed that there is no determinism here, there is nothing in situation t which necessarily determines the resulting choice, but this situation could still reflect free choice and not randomness. Van Inwagen effectively denies the possibility of choice on an a priori basis, and based on this reasoning proves that if determinism is not in force here then we are clearly dealing with randomness. This is of course begging the question, and it is no wonder that he arrived at the conclusion that it cannot be that libertarianism is true. As an aside we will note that in truth, Van Inwagen’s argument is a little more complex than that. He speaks of a statistical experiment of a choice repeated a thousand times in the same situation, in which we end up with a statistical distribution of choices of X or Y. The fact that nothing in the previous situation determines the result, and that a statistical breakdown of the different choices emerged, point in his opinion to randomness. However, in chapter two we not only pointed to the lack of a cause in a situation at moment t not pointing to randomness, we also saw that even a statistical distribution of results, even if it repeats itself on average with every thousand experiments we do (say, we always get 300 choices of X and 700 choices of

Y), still does not necessarily point to a random process. As we saw there, even a collection of free choices usually breaks down statistically like a collection of random events. We saw there that according to libertarianism, all of psychological and educational research is based on this assumption. In other words, the argument of Van Inwagen begs the question. He assumes that the only possibility which is not deterministic is randomness (i.e., he assumes that there is no free choice), and thus concludes that there is no free choice. But the libertarian who espouses free choice will not agree to the dichotomous assumption itself; for him, there is also a third possibility, beyond randomness and deterministic causation — free choice. His statistical argument assumes that in a libertarian world, there is no possibility of using statistical tools to describe human behavior, and he is wrong here as well, as we saw in chapter two. I find it odd that despite these blatant fallacies, this dilemma argument repeats itself as an ace in the hole in almost every discussion on the issues of libertarianism.

Logical Determinism There are some, albeit not many examples of arguments which derive factual conclusions from conceptual analysis. All these arguments lead to a lot of unease, but the best of them really get the philosophical creative juices flowing, as Bertrand Russell said (in reference to the ontological argument, as we will discuss shortly): It’s far easier to smell the problematic nature of the arguments than to actually put your finger on them. The most well-known argument of this sort is the ontological argument for the existence for God, which was first proposed by St. Anselm in the twelfth century, and has been used in various versions by many philosophers59. This argument tries to base the claim that God exists on an analysis of the concept “God,” and effectively derives His existence from His definition as being perfect. It seems to me that the most prominent such argument, after Anselm’s, is Decartes’ cogito (see our discussion of it in the Prologue, where we explain why it is a conceptual analysis not based on facts). Another conceptual analysis, which I will presently discuss, is the argument of logical determinism, whose goal is to prove determinism. It is also a proof of a factual claim about the world which is based on a conceptual analysis. This argument has come down to us through Aristotle, and others have made various contributions to it throughout history.60 Aristotle based himself on the assumption that the truth value of a statement cannot be time-dependent. Let us use an example he used himself. Let us assume that tomorrow on Tuesday (August 23,2011), a naval battle will take place between two great fleets in the heart of the miniscule Yeruham Lake in the Israeli Negev, which is a small pond in the middle of nowhere. In such a situation, the argument “On the date of August 23, 2011 a great naval battle will take place in Yeruham Lake” is a correct argument. We say that the truth value of this statement is “true.” In the event that a naval battle does not take place, then the truth value of this argument is “false.” If such a naval battle did take place, then it is clear to everyone that from Tuesday onward, the truth value of this statement is “true.” But what was the truth value of this statement on Monday, a day before this dramatic event left an indelible mark on modern history? One can argue, with a great deal of logic, that the truth value of this argument would be “true” then, as well. Why? Because its content fits the situation it describes (even though we don’t know it yet, of course). What about two thousand and five hundred years before that (in Aristotle’s own time)? Let’s say that Aristotle tells his student Alexander the Great: “Know, Alexander my dear pupil, that on the date August 23, 2011, a great naval battle will take place in the Yeruham Lake in the Negev in Israel.” Did he make a true statement? Of course he did. Again, the content of the statement fits the state of affairs in the world itself, and this is the accepted definition for a true statement. Thus, the truth value of this argument was “true” thousands of years before the epic events it described occurred. It was always true. What meaning does the timeline nevertheless have in this context? It has meaning only as regards the question of whether this information is in our possession or not. Until the date in question, no-one in the world can know, at least not with certainty, that the truth value of this argument is “true,” even though this is the case. This analysis is of course applicable to any truth claim. We were not limiting ourselves here to naval battles or Yeruham Lake. Thus, the conclusion is that the truth value of any statement is fixed and eternal, and is effectively not dependent on time. What changes with time is only the state of our knowledge on the matter. But that is an argument about us, not the truth value of the statement itself. The fact that Aristotle, like his contemporaries before the stated date, could not know the truth value of this statement is unimportant. What matters for our purposes is that already then, this statement had a defined and exclusive truth value, even if no-one knew or could have known it. The next stage of the argument is that if this argument was true two thousand and five hundred years

ago, then it cannot be that the battle will not happen on the day. If it does not happen then the argument is not true today, and was also never true. And this contradicts our assumption. There is no need to say that even if the argument was never true we would not be spared the deterministic conclusion; in such a case it cannot be that such a battle will take place, as its existence would contradict the falsity of the argument, which was correct long before that. From here, the logical determinist concludes that in any case, regardless of the truth value of the claim made by Aristotle to Alexander, the event in our time is predetermined and unchangeable. It cannot be otherwise, and it is the case when it comes to any other factual claim. The fact that I ate a tomato now is a fact that has been true since creation, as the argument “Michael Abraham ate a tomato on date X” was also true already in the days of Aristotle, and could therefore not but occur on the day it was to happen. Beyond the disagreement with the surprising conclusion of logical determinism, there is an additional problem here: How can it be that we derive a factual conclusion about the world (i.e., the conclusion that the world is deterministic) based on a conceptual analysis? Ever since the ontological argument was made, and primarily after Kant’s arguments against it, this difficulty has haunted all conceptual arguments about the world. It has also led to a number of challenges to logical determinism. Here we will present two common kinds of challenges: non-binary logic and constructivism. Three-valued or non-binary logic was developed by the Polish logician Łukasiewicz. This argument argues against the law of the excluded middle. In other words, if it is not false that a naval battle will occur tomorrow, then it must necessarily be true. Against this, the proposal has been made that we adopt a different logic regarding future events, according to which we are not willing to even make the banal statement that “tomorrow a naval battle will or will not happen.” A second argument against logical determinism is constructivism, which comes out against the assumption that there is a given group of true claims, even if no-one ever determined them. Constructivism argues that a claim is true if and only if we have determined its truth constructively, or at least if we know about it. This approach effectively rejects the idea that the truth value of claims is not time-dependent. I see no need to go too much into detail in either direction here, as both suffer from at least two fundamental problems: (a) it does not understand the problem; (b) it offers an utterly unreasonable solution for it. The first solution is no solution at all. I’m not looking for a different logic which will allow me to not even present the problem; I’m looking to see why there is no problem based on my normal logic. Just as an example, if I wish to discuss Łukasiewicz’s argument itself, I will of course do so based on my normal (and binary) logic. On the other hand, the second type of challenge offers an utterly unreasonable definition. Of course the truth value of a statement is not time-dependent. So long as the content of the statement fits the situation it describes, it is a true statement. One can offer a different definition for every concept which bothers us, but as I already noted above, to solve a real problem, it is not enough to simply change definitions.

On Logic and Time61 It seems to me that the main fallacy in the argument of logical determinism lies in the complex relation between logic and time. The argument of logical determinism explicitly assumes that there is a disconnect between logic and time, but there is an implicit assumption that such a relationship nevertheless exists. On the one hand, the argument assumes that even if the truth value of the claim is determined tomorrow, it exists today as well. This is the detachment between logic on the one hand and physics and time on the other. On the other hand, this argument also assumes that if the truth value of the claim is “true” today, then this also dictates the event itself tomorrow, which is to say that logic does indeed dictate physics. This is precisely the point where the logical determinist errs. The physical event is what dictates the truth value of the claim. In other words, the proper direction is from physics to logic, not the other way around. The content of the claim determines its truth value, not the truth value determining the content. If a naval battle does take place today, this establishes the truth value of the claim, but it is “true” from creation until now. What is disturbing in this picture is the influence which appears non-causal, or causation working backwards in time. If the physical event is now, how can it be that its results (attaching a “true” value to the claim describing it) take place already thousands of years before? How can it be that an event which takes place today determines the truth of a claim made millennia ago? The answer is that causal influence which moves backwards in time is apparently not possible; that is the limitation imposed on us by the laws of physics. But here we are not dealing with causal influence related to physics, which represents the relationship between two events in the world, but in a logical (or meta-logical) connection. When it comes to such a connection, there is nothing preventing it from moving from future to past.

Let’s try and put it another way: The event today determines the truth value of the claim, and this truth value is non-temporal and atemporal — it is true and fixed at all times. The truth value of the claim was always “true,” but until now we (and everyone else) didn’t know it. Now we do. Yet at the same time, that does not mean that there cannot be a situation in which a naval battle does not take place and that the battle is predetermined (as determinism argues). Such a situation is certainly possible, as libertarianism argues. If the naval battle does not occur, then the truth value of the claim will retroactively turn out to be “false,” and this will also have been the situation always even though we didn’t know it yet. The relationship between the events and the truth value of the claim describing it is not a causal one, as we are not speaking of a relationship between events in the world but between an event and a logical situation. Such a relationship can certainly work against the timeline. Logical determinism is therefore a result of a categorical confusion, and there is subsequently no need to seek out explanations or excuses. The truth value of a claim is not a kind of reality, but a logical definition we created, and we must therefore avoid applying categories related to facts (like the temporal direction of causes) to it. We will encounter a very similar error, which also derives from the confusion of logic and time, in the next chapter when discussing the paradox of the initiation of movement and causality.

The Theological Argument: Divine Knowledge and Free Choice Religious approaches stridently oppose determinism (see more on this in the appendix). The common religious approach is that God expects man to choose good and do good (“And you shall choose life so that you and your seed may live” – Deut. 30:19). There are deterministic approaches in Christianity (Calvinism, and Protestantism in general, are special kinds of determinism), and there are also some esoteric approaches along these lines in Judaism, but we can say that the mainstream of religious thought in general is libertarian. Sometimes it seems that many atheists see religion as the source of libertarian thought, and they therefore expect scientific findings to extricate us from these “prejudices” and “superstitions.” Against this background, it’s worth noting that one of the most ancient considerations in favor of determinism is actually a religious one. A common assumption in monotheistic religions is that God is allpowerful, and therefore is also all-knowing. If He is all-knowing, then He should know the future, as well. But if he already knows the future, then that means that the future is fixed. An event cannot take place which is contrary to divine knowledge, as if such an event happens then that means God is not all-knowing. For instance, God knows that a naval battle will take place tomorrow. Thus, it cannot be that a naval battle will not take place, as if it does not, then God’s knowledge would be incorrect, contradicting the assumption that he is all-powerful and all-knowing. These arguments lead to a religious approach which ostensibly supports determinism. A slightly different version of the religious argument is based on everything occurring in the world being done by God. It cannot be that something will be done here not by His hand, and certainly not against His will. This view of the world dependent on God leads to a fatalistic-deterministic approach. For instance, many Jewish thinkers wonder: If all of man’s sustenance throughout the year is determined already from Rosh Hashanah, the Jewish new year (see Tractate Beitzah 16a), then why should he go to work? Will it change anything? The argument here is that the acts of man are determined by fate, or a supreme power, rather than his own choice. The kabbalistic concept of tsimtsum or reduction, according to the common interpretation, is meant to solve this problem. According to this, God ostensibly reduced himself, thus leaving room for the individual to decide and act according to his choices. It’s important to understand that this argument is fundamentally different than logical determinism. It can therefore not be solved by detaching logic from time, as we did in discussing logical determinism. God’s knowledge of any event is not the truth value of a claim describing the event, but rather information which exists right now in some location (metaphorically speaking, “in God’s mind”), and He can in principle pass it on to humanity. In other words, this knowledge can in principle be accessible (this in contrast to the truth value of a future claim, which as we saw is fundamentally inaccessible before the event actually occurs). If such knowledge exists and is credible, then it cannot be that the unfolding of events in reality itself will not align with it.62 Later in the chapter, we will see proof of this from the Newcomb paradox. Let me add another angle which will clarify the uniqueness of the religious argument as opposed to the causal one. At the end of the previous chapter, I presented the position of Rav Hisdai Crescas, in his book Or Hashem [The Light of God], which proposed a deterministic thesis based on this argument. In his opinion, the world is deterministic and everything which occurs therein has already been set. What is given

to us is only the decision how to address what happened (or what we do). If we treat sins negatively then we chose good, and if we treat them positively (we wanted them), then we chose bad. But Crescas comes to deal also with the theological difficulty, not just the one related to physics. If his determinism derives from God knowing everything in advance, and is not based on the assumption of physical causality, then it is an argument with different consequences. When the basis of the argument is theological, there is no room for dividing between physical events and mental events. After all, God knows what is going to happen because He is all-powerful, whether the event is physical or mental, so how can He not know an individual’s future attitude to the event? By contrast, the argument of causality can be true only when it comes to physical events, as mental events are not necessarily subject to physical causation (we will expand on this later in the book). And in truth, further examination of Crescas’ argument reveals that he presents it in the context of causal-physical determinism, not theological determinism. In my book Two Carts and a Hot Air Balloon (Hebrew), I explained there that the simplest solution is that God does indeed not know. His very desire to allow us free choice forces him to forgo his knowledge of what will happen in the future. This means that what God knows is only the topographical map within which I operate, meaning the totality of circumstances and influences affecting me, and based on this what I am likely to do on average and reasonably. But as we saw in the second chapter, beyond this totality there is also individual choice, and God does not and cannot know what this is (as one cannot know information which does not exist).63 This argument is ancient, and dozens if not hundreds of thinkers have struggled with it, both religious and not. The important point for our purposes is that the mainstream of religious thought sees this argument as either a contradiction or a challenge. It considers it obviously untrue; it’s just not always clear why and where the mistake lies. Thus, religious approaches at most present this argument as a challenge, but they certainly do not deliberately intend to follow it down the deterministic path. 58 Peter Van Inwagen, “Free Will Remains a Mystery,” in R. Kane, The Oxford Handbook of Free Will, Oxford University Press, 2002, pp. 158177. 59 Norman Malcolm (1960), “Anselm’s Ontological Argument,” Philosophical Review, 69 (1): 41-62. 60 See a survey of the subject at Stanford University’s Virtual Encyclopedia (SEP) under “fatalism,” from 2010. See also Richard Taylor (January 1962), “Fatalism,” The Philosophical Review, Duke University Press, 71 (1): 56-66. 61 See, for instance, Michael Abraham, Israel Belfer, Dov Gabbay, and Uri Schild, Temporal Logic in the Talmud, College Publications, 2011. 62 See more in the first section, ibid. 63 See ibid. on the question of whether God can be “subject” to logical constraints.

B. A Priori Arguments Against Determinism What is judgment? The libertarian sees the individual as acting based on judgment. Per his approach, when we face a number of options, we consider each one of them, try to predict their consequences, and then make a decision what to support or how to act. The result of judgment can for instance be a socioeconomic position (becoming a capitalist or communist, preferring equality or freedom), but it is not the blind derivative of the circumstances but rather the result of judgment. Another individual examining the same circumstances and equipped with the same knowledge may reach an entirely different position than mine, because his judgment (or his value system) is different. This is an expression of the approach of the libertarian that the present situation does not determine the result, and that value decisions are not the deterministic consequence of situations or any collection of circumstances existing in the world. We use similar judgment when we make decisions regarding facts. The question whether the stock market will go up or down, whether the next war will be beneficial or harmful, and so on are fundamentally factual questions. The answer is yes or no, and they can be examined against the facts (at least after they actually happen). Still, because of the complexity of the facts, the answer is not a simple processing of them but the result of judgment. Here, too, two different people may arrive at different conclusions based on the exact same data, since their judgment or value systems are different. What about scientific questions? For instance, is a particular theory true? Or what is the appropriate scientific theory to explain a given data set? This is a question whose answer does not lie in the simple processing of facts. There are a number of possible generalizations one can make based on any collection of data (including an infinite collection), and therefore the choice of one of the generalizations is a function of our way of thinking and the consequence of judgment. This means that scientific decisions also involve judgment, and are not simply the processing of facts (this is in effect David Hume’s induction problem; see more on this next chapter). What about the meta-scientific question, like determinism or materialism? These are questions where the answer is certainly the result of judgment and not the simple processing of facts. On the other hand, these are also clearly factual questions, as it is clear that the world is either deterministic or not as a matter of fact; it is we who simply don’t necessarily know it one way or the other. So, we see that human judgment affects decision-making of various kinds, from value decisions (communism and capitalism) to factual questions (the rise and fall of the stock market) and scientific questions (quantum theory) and certainly when it comes to meta-scientific and metaphysical questions (determinism and materialism). But what is this judgment? We know it’s not simple processing of facts and deductive reasoning. When all the data lies before the individual, we must decide what is more convincing or what best fits the actual state of affairs in the world. There is a degree of speculation here, and the individual forms the answer from his psyche as well as with his way of thinking, logic, and even experience. The determinist, by contrast, believes, as we said, that every event in the world is entirely determined by the circumstances which preceded it. He sees the judgment we use when choosing one of a number of options as a kind of illusion, or at most as a consequence also determined by prior conditions (including the structure of our brain, which is itself determined by our biography and genetics). According to his approach, there is no actual component of judgment when it comes to processes of ethical choice or decision-making when it comes to facts. The individual is nothing more than a box which absorbs data from the environment, processes them automatically, and creates an output, which is either thought or action or both. This data processing system is itself the result of circumstances and an inborn genetic and neural/brain structure. According to the determinist, when two different people stand at the same fork in the road with the same data, their decisions may be different, but this is not due to judgment but rather their different brain structures (some might say personality or thought, but this is a metaphor), and really being from a different environment (internal and external). The automatic way they process data is different, and their output may therefore also be different. That different “programming” is really why they produce the answer they do. In sum, let us think of a situation in which two scientists face the same set of facts, and espouse two

different scientific theories. What creates the difference between their different decisions? The determinist will argue that their brain structure leads them (by force) to different theories, while the libertarian will argue that we are looking at different judgments, different assessments regarding the truth (much like the debate between them on value choices).

The Fool’s Cycle of Determinism How should we treat the conclusions of an individual in light of the deterministic picture? Ostensibly, we now speak only of the outputs of a programmed machine, and there is no reason to think them correct or credible in any sense. When an individual concludes that the world is round based on some set of data, does that mean the world really is round? This depends on the nature of the automatic machine he uses when he makes his decision. An automatic machine can produce yes or no answers for its own reasons (its electric circuits direct it to this or that answer), and there is no guarantee that the answer it produces has any correlation with the world itself or any external truth. If we go further, when an individual concludes that our world has a quantum structure, or that the physical world is based on a 63-dimension string theory, does that mean that the world is really like that? Or perhaps all we can learn from this is something about the automatic box which arrived at this odd conclusion? After all, another box would certainly produce a different answer. Moving one step further to the meta-scientific, we must ask: When an individual arrives at the conclusion that the world is deterministic, or materialistic, how significant is this conclusion? As we saw, here we are certainly no longer dealing with the simple processing of empirical information but rather a determination which involves some degree of speculation. No-one claims to have directly witnessed the deterministic character of the world or that it has no spiritual substance. These are conclusions he arrives at from various data or from some a priori logic. But does any of this say anything about the world, or just the box that said something about the world? Moreover, even our acquaintance with the box is not direct. Even this knowledge is arrived at via the box itself, with the same techniques it applies to external issues. We therefore have no guarantee that the box is not built in such a way that it forces us to trust it (not necessarily justifiably). Let me add another angle at the margins. Who is this “I” that is supposed to accept or not accept the results of this system, or form an opinion, deterministic or otherwise? After all, we saw in chapter two that according to the determinist, “I” am nothing more than the automatic system itself, and we must therefore remove any feeling of “me” which is accompanying the problematic issues defined here. This brings us back to Descartes’ cogito, which states cogito, ergo sum (I think, therefore I am; see the prologue at the beginning of the book). Here we encounter the other side of the coin: I don’t think, therefore “I” am not (“I” in the sense of some autonomous personality, which is supposed to form a position, consider data, and so on). We can put it this way as well: Is the illusion that our insights about the world are correct fundamentally different from the illusion of choice, moral responsibility, moral condemnation, self-awareness and the psyche, and all the other illusions which the determinist argues accompany us so powerfully with every step we take? Who will assure us that our “thought” on this issue as on others is not just another illusion in this series of mirages? It’s important to understand that this difficulty is not the same as standard skepticism. The skeptical argument raises the possibility that our conclusions are not true, and that our tools of thinking and recognition are not necessarily reliable. How do you know, it asks, that your conclusions are reliable? Perhaps your eyes are fooling you? Maybe your brain is deceiving you? Perhaps your feelings are but an illusion and distortion of reality itself (if there is any sort of external reality)? Or, as Descartes put it, perhaps there is some demon fooling you? There’s nothing one can respond to these challenges, of course, but most reasonable people are not actually bothered by them. People believe their conclusions and that’s it, at least so long as they do not have good reason to doubt them. Regular skepticism is appropriate for a situation in which the individual sees a sea and decides it’s real, and therefore concludes that he has arrived at the sea. The skeptic asks him: How do you know? Maybe it’s a deception? Can it not be that you’re dreaming? The individual pinches himself and concludes that he’s not dreaming and that there’s a sea here and that’s it. Thus is the skeptical nuisance silenced. For instance, when George Edward Moore64 was asked to prove the existence of an external world, he lifted his hand and said: “This is one hand, and this is another hand. There are therefore two external objects in the world. Therefore, an external world exists.” In his article from 1925, “The Defense of Common Sense,” Moore argues against idealism (a philosophical approach which denies the existence of an external

world) in which the reasons for accepting the metaphysical assumptions of the skeptics are far less good than those which support common sense. Moore wonders why their odd argument is more convincing than the knowledge about the world which the skeptics propose denying. He believed the fact that he had a hand rather than the assumptions of a strange argument presented in a philosophical lecture in favor of idealism. Moore refused to be bothered by something which didn’t really bother him. However, when it comes to the argument I presented above regarding the deterministic picture, the situation is entirely different. After all, in the deterministic picture, the individual already knows (?!) that he is subject to a blind and arbitrary system which makes his decisions for him (who?...). It is clear to him that his positions and conclusions are not really his and that they are not the result of judgment, and that they are instead the consequence of blind, mechanical, and arbitrary calculation. Now he must (?!) think about the nature of this coercive system. Is it sympathetic, ensuring that his conclusions align with reality itself? Or is it Descartes’ demon whose aim is to fool us? Perhaps this system doesn’t have a purpose at all, and it just produces arbitrary results regardless of what happens in the world itself, sometimes aligning with it and sometimes not? In contrast to the regular skeptical argument, the determinist needs to deal with a far greater difficulty. When I’ve already accepted the argument that my decisions are arrived at based on constraints not dependent on me, and certainly not based on my judgment, then he is compelled to question the nature of this coercive system. From his perspective, this is no longer an academic and superfluous philosophical nuisance but an obvious and pressing question. If we take different black boxes built randomly, each of which produce different results for the same input, obviously the outputs will usually not fit reality. The odds that any unknown box will produce a reliable output, matching reality, are almost zero. The assumption that the coercive system is not sympathetic or credible is therefore a logical one, and the determinist bears the burden of proof to argue that it is credible. Obviously he cannot present any evidence, as any such proof would itself be based on our thinking, which is nothing more than a consequence of the coercive system. Therefore, there is no independent way, according to his method, which would allow us to get out of it. A deterministic worldview is therefore a view which saws off the branch it sits on. If we adopt a deterministic worldview, then there is no real justification for our faith in our own decision-making process, and thus also not in the deterministic conclusion itself (as it was also arrived at based on those same arbitrary systems). It’s important to stress here that a situation in which these conclusions are reliable is possible in principle, even if the odds are very low. If our automatic system is sympathetic, credibly reflecting reality, then we can certainly learn about reality from our conclusions and thus can accept them as real. The problem is that even if this is true, we have no way of knowing this. We have no way of discussing the coercive system, as even that discussion itself will take place within its framework and the tools it dictates to us. For instance, an individual can come and say that we should not have faith in the laws of physics because they were forced on us by some arbitrary coercive system. To this the determinist will reply that this is true, but it turns out that the system is a sympathetic one. The proof of this lies in the laws of physics meeting the test of empirical observation. We test them all the time, and they work. But this argument itself — that the laws of physics should be believed if they are confirmed empirically — is actually a conclusion of ours, and is as such forced on us, like any use of our logic and intellect. So, how can we trust it? Moreover, even the tools we used to measure the phenomena were built based on that same physics and logic. Can we trust them? At most we can point to what we see as the internal consistency of the laws of physics, but there is no way to conclude from here regarding the degree to which they match reality itself. The determinist believes in the existence of a coercive system which we have no way of exiting, and therefore no way of examining if it truly is sympathetic. This is much like Bertrand Russell’s celestial teapot. Let’s assume that someone comes and tells us that the planet of Jupiter is orbited by a transparent, small celestial teapot, and therefore no-one can observe it. We will not tell him that the odds of this are necessarily zero and that this is certainly not true, but we will certainly tell him that this statement lacks any basis. It could be true, but neither we nor the advocate of the teapot has any way of knowing this, and there is therefore no basis for assuming this. This is one of the main arguments against the determinist. The determinist believes in the existence of a coercive system which we have no way of exiting, and we therefore have no way of knowing whether it’s sympathetic or not. Thus, the determinist might be correct, but the fool’s cycle argument tells us that in any event, we who are subject to the system have no way of knowing it. The argument is therefore without

foundation, even if not entirely out of the realm of possibility. Moreover, it is also almost entirely outside of the realm of possibility, as most of the systems we can think of are not reliable. Take the collection of parts of a television device, and think of all the possible ways to put them together. How many combinations will create a working system? The percentage is negligible. Therefore, lacking specific information and condemned to lack of knowledge by our very natures, there is no basis for thinking that our “device” (our recognition and thinking) works.

The Role of Evolution It seems ostensibly that the only way to overcome this difficulty is with arguments from evolution. The determinist can argue that while our system of cognition was indeed shaped randomly, the evolutionary processes saw to it that it would be reliable, and the conclusion is that despite the system being designed in an arbitrary and random manner, it is nevertheless credible. A credible system has a greater level of survival value, and therefore even if there was no guiding hand, a coercive but sympathetic system has been formed within us (or consciousness and thinking which are reliable). Others will say that God, with His great mercy upon us, is the one who created the sympathetic system, and therefore our system is reliable. We thus have a rare consensus between neo-Darwinists and believers. The two sides are in a state where they can be both determinists while at the same time believing that their consciousness and thinking are credible. Sadly, I must crash this party of peace and fraternity. Prima facie, these two arguments are problematic, and for the same reason. The assumptions of the determinists, both neo-Darwinistic and believing determinists, were arrived at with the same tools. The conclusion of evolution and its conclusions was itself made based on that arbitrary system, and it is therefore difficult to see this argument as the Archimedean point which will get us out of this jam. How did we conclude that evolution is itself true? How do we know that that the conclusion we derive from that fact on the reliability of our cognitive systems is also correct, even if evolution is true? After all, those same considerations were weighed using those same arbitrary systems, and are therefore dependent on their sympathetic nature or credibility. The same is true of a determinist believer. After all, his assumptions or conclusions that there is a God and that He created a coercive sympathetic system within us are the result of that same system. How would he know if they are reliable or not?

Further Clarification: The Dilemma Argument and Internal Consistency [This section is meant for those with a strong philosophical heart, as it has a number of non-trivial twists. It can be safely skipped.]

We can formulate this argument against determinism as follows: Accepting the view that the world is deterministic is done with the same tools forced on us, and if we adopt a deterministic picture then we have no basis for the faith we place in them, and therefore cannot place any faith in the deterministic conclusion itself. There is a delicate philosophical point here, which I will try to clarify in this section. Prima facie, the logical structure of this argument is problematic. After all, the determinist could respond with a dilemma-style argument of the sort we noted in the first part of this chapter: If the deterministic picture is itself deterministically forced on me — then determinism is true (as it forced that view on me). If the picture is not forced on me, then I freely concluded it — and then it is once again true, even according to the libertarian critic (as it is the result of freely used judgment). Therefore, my conclusion that the world is deterministic is indeed consistent. This is of course no proof that determinism is true, but only a possible way out for those who espouse this position, a way out which enables them to continue to hold onto this picture despite the problems we described above. But his formulation contains a logical error. The second horn of the dilemma, the possibility that the picture itself was based on free judgment, contradicts itself. After all, if this is indeed the result of free judgment, this means that I have such judgment, and therefore determinism cannot be true. The natural conclusion of this argument is the opposite: It is apparently an error on the part of a system with free judgment which mistakenly arrives at the conclusion that it is itself deterministic (judgment systems make mistakes, too). The determinist can therefore not really speak of the two horns of the dilemma. Moreover, even if this consideration was created by a coercive system, that does not mean it is reliable. It may be that the deterministic conclusion is correct but that the system which forced me to arrive at this conclusion is not credible, and vice versa. The dilemma argument is therefore not relevant in this case. What we can do here is examine whether the deterministic approach is coherent (i.e., internally consistent) — that is, see if we can speak of a

deterministic approach arrived at based on deterministic tools, and accept it. Above, we saw that approaches adopted in this manner are not acceptable and are indeed without foundation. But regarding the specific question of determinism, the answer is of course that they are. Specifically, the deterministic approach itself is acceptable, even if it results from the conclusions forced on it by the system. Why? Because even if determinism is forced on us, then this itself proves that our decisions our forced on us, thus proving determinism. There is therefore nothing preventing the adoption of the deterministic conclusion despite the difficulty of the vicious cycle described above. The intelligent reader is likely telling himself that this is begging the question. If this approach is indeed forced on us because of determinism, then determinism is true. At the same time, if it is not forced on us by determinism, then it is not true. This is ostensibly empty logical casuistry. But this casuistry does actually neutralize the vicious cycle argument. It shows that the determinist can remain consistent in his approach and it does not contain internal contradictions. But there is a fundamental error in this argument for the defense. Even if a specific decision was adopted due to a system which forces it on me, that does not mean that all my decisions are forced on me — and that is the deterministic argument. The deterministic argument does not just say that there are decisions that are forced on us, but that all decisions are forced on us. Therefore, even if this specific decision (that the world is deterministic) was forced on us, that does not mean that determinism is true. The deterministic conclusion is therefore not truly irregular, and it too, like any other conclusion arrived at by force, is not acceptable. Again, even if it is possible in principle, it remains without foundation, as I explained above. We can look at it from another angle and say this: We already noted that the conclusion regarding determinism doesn’t derive from direct observation. We don’t observe determinism and thus conclude that it is true. Determinism is also not logically derived from the laws of physics and evolution (to say nothing of the laws of physics and evolution themselves not being purely logically derived from observations). Determinism, at most, is a conclusion we arrive at through generalization or analogy from interim conclusions — these being evolution and the laws of physics. Therefore, if these interim conclusions themselves are forced on us and are not the result of judgment rather than the output of an arbitrary system, then we have no basis for believing that evolution or physics are true. If so, what’s the point in arriving at conclusions based on them? And even if we do adopt them, we will need to make another generalization to move from them to a deterministic conclusion (for instance, that the laws of physics also apply to will — see more on this in the sixth chapter on causation), so why should we trust it itself? It is therefore wrong to discuss the deterministic conclusion itself, which is indeed irregular. The discussion does not revolve around the conclusion but rather the interim conclusions on which it is based, and these are not irregular — when it comes to them, it is correct to say that if they did come to us via some coercive system, our faith in them is also without foundation. Bottom line, even if we accept the deterministic picture by force of the dilemma (which even if it is forced on us, then yes — our ideas are also forced on us), this at most establishes our faith in determinism. The associated cost is that we will have to throw out all our other insights, such as the laws of physics and science in general, as these are certainly not proven in the same logical way, and therefore the counterarguments against them (i.e., that they are forced on us, then there is no basis for the assumption that they are correct) remain valid.

Plantinga’s Argument Against Naturalistic Approaches Until now, we dealt with the a priori argument. Evolution arose as a possible explanation, and we rejected it since it leads to a circular argument. Here, I will present another argument, which points to the fact that even if we ignore the problematic circularity we pointed to previously, we are still left with a structural problem inherent in the explanation from evolution. In contrast to what many think, evolution does not actually offer us a way out of the philosophical problem (regardless of the question of its scientific validity). This is Alvin Plantinga’s argument against naturalism.65 This argument led to a wave of responses and challenges, and dozens of articles and even books have come out in favor and against it. In 1994, he published an updated version of the argument,66 in which he responds to the main challenges against it. First, he defines naturalism as an argument that contains nothing beyond the material and the natural — neither God nor any supernatural entities (including a soul or a psyche). Afterwards, he describes evolutionism as an argument that the human being (like other animals) developed in a natural evolutionary process without any outside guiding hand. His last definition is the definition of reliability, and this is the argument that our cognitive abilities — such as memory, thinking, and sensual recognition — are reliable, and that in most cases they provide results which fit what happens in the world itself (this does not mean that they don’t have errors sometimes, but we can have basic faith in them).

Evolutionism argues that our physical form, our character, and our behavior are determined according to considerations of adaptability, fitness, and ultimately survivability. What is beneficial for our survivability is the characteristic which becomes dominant during the process of survival of the fittest, and therefore becomes a dominant characteristic of the population of the next generation. Thus does our behavior become more and more adaptive. From here, the determinist will try to derive that it also becomes more and more reliable, as a response to the challenge we presented above. However, Plantinga argues that natural selection deals with behavior and not worldviews. The survival of any creature is determined by its behavior and not how it views the world. If an individual takes care to flee the tiger, he will be saved and will survive, regardless of whether he did it because he thought the tiger is dangerous or he thought the tiger wanted to play tag with him. What’s important is what the individual does, and not what he thinks. Evolution can therefore at most ensure optimal behavior in terms of survivability, but not necessarily reliable thinking and consciousness of the sort that are optimal in terms of their alignment with the truth. Put differently, the match between what we think and the truth is not a relevant parameter for natural selection. The fact that we flee tigers is indeed an important characteristic for our survival, but the motive for that flight is not. The same is true when it comes to our consciousness and our thinking: We have no guarantee that they actually reflect what happens in the world and are credible, as the content of our thoughts is not relevant to our survivability. Evolution is therefore no guarantee for the reliability of our thinking and consciousness, as it does not screen for correct thoughts, only beneficial behavior. If an unnatural factor was introduced, like God, who directs our development and who shapes our cognitive abilities, there would be room for our conclusion that these abilities are reliable as information. But within the framework of naturalism, which does not accept the existence of supernatural entities and which sees all of our development as deriving from evolutionary processes, the logical conclusion is that we cannot believe our own thinking and consciousness. The question asked by Plantinga which is relevant for our discussion is this: What are the odds that our cognitive skills are reliable, on the condition that we are basing ourselves on an evolutionary and naturalistic picture, accepting the arguments of naturalism as well as evolution. Plantinga’s argument is that the odds of this are very low. The reason for this is that if our cognitive abilities did indeed develop solely on an evolutionary basis, then they could have taken on any form. The processes of natural selection, as we already noted, do not give preference to reliable cognition (e.g., creatures who flee tigers because the tiger wants to eat them), but at most cognition for purposes of survivability (creatures who flee tigers). The possible reasons for correct conduct can be varied so long as they produce the correct behavior for purposes of survivability. Thus, natural selection will not create creatures who flee the tiger because it is dangerous, but only creatures who flee for a variety of reasons. The flight from the tiger may be based on the fear that it will eat me, or because I want to play tag with him, or because I think him ugly and I hate being near ugly creatures, or because it reminds me that I didn’t eat lunch yet. All the creatures who run for all these reasons would survive to the same extent, even though most of them (and perhaps all of them) have unreliable cognition skills. So why should we assume that we, as ones who survived the evolutionary process, are specially gifted with reliable skills? The odds that this would be the coincidental result, without any constraint, are negligible, and evolution increases the chances of this but little. After all, the number of possibilities of unreliable interpretation of data is far vaster than the option that they are credible, which is exactly one. We could go even further, saying that the idea that the tiger is dangerous does not in fact grant any evolutionary advantage, and it would be far more efficient if cognition did not emerge at all (instead of false cognition, and even reliable cognition). Let us think, for instance, of a creature which flees tigers because the presence of a tiger triggers his running muscles, without any cognitive basis. He has no explanation as to why he runs; he just runs because there is a tiger nearby. This is a different situation than the previous ones, as this creature does not have any sort of consciousness of causal thinking. He is simply programmed for proper conduct. Moreover, an instinctive reaction is usually faster and more credible in the struggles for survival than reactions which involve our cognition.67 Our cognition is therefore effectively superfluous in evolutionary terms. We could have created creatures which would move straight from the relevant physical-chemical state to the behavioral result, without any need for thoughts of any sort in between. For our purposes, this means that evolution does not deal with these dimensions, and can therefore not be an explanation for their reliability. Evolution thus does not improve the odds that cognitive skills are

reliable — but in fact, Plantinga argues that it is entirely irrelevant in this respect. In short, Plantinga addresses a possible argument that there is some connection between cognition and conduct, in which conduct deriving from correct cognition is also more optimal in terms of survivability. If this argument were correct it would be a response to his own argument, as to behave correctly one must first develop reliable cognition. To show why this is not true, he makes a complicated calculation regarding the relationship between cognition and behavior, concluding that there are no strong ties between the two. This is sufficient for our purposes here. However, one could make a softer version of this argument. Even if evolution leads to traits and skills which increase survivability, this is no guarantee that they are more reliable. And even if someone does trust them on a probabilistic basis, thinking that there is some correlation between survivability ad reliability, that is certainly insufficient to explain the absolute level of faith we place in our tools of thinking and consciousness. No-one considers what their eyes see to be a matter of odds. He does not believe that what his eyes see have a high probability of reflecting the world; he is certain that he is simply looking at the world. This certainly cannot be based on evolution.

Interim Summary: The Inherent Contradiction in the Deterministic Approach We see that the deterministic approach saws of the branch it sits on. If we adopt it, it renders itself without foundation. We are thus dealing with a kind of internal contradiction, which is not a standard logical contradiction: If it is correct, then it is without foundation (even if it might nevertheless be true, we ourselves have no way of knowing). Therefore, to be more precise, this is not an argument against determinism, but against our holding onto a deterministic position. We also saw that an explanation from evolution does not untangle this knot, for two reasons: (1) a priori: Even the consideration from evolution is a product of our cognitive system, and therefore cannot itself be accepted. (2) a posteriori: Even if we accept the theory of evolution, Plantinga’s argument shows that the odds that our cognition is reliable based on the acceptance of naturalistic evolution are very slim. We also added that even if the odds are not all that slim, they are certainly not a basis for our complete faith in our senses. So, even if we conclude that determinism is a necessary consequence of our scientific findings, this cannot be a reasonable basis for its adoption.

Another Level of the Deceiving Demon Argument So far, we have seen two levels of the deceiving demon: (a) standard skepticism — who said that all (or even part of) your insights are correct? Are these not the consequences of some deception? This argument can be made against any philosophical approach and any individual, and usually we treat such an argument as possible but without foundation. (b) The fool’s cycle — this argument appeals solely to the determinist: After all, you declare that you accept the existence of a coercive system, some demon. If so, how do you know that it’s sympathetic and credible? Maybe it’s a deceitful demon? Moreover, lacking any specific knowledge, most demons one can think of obviously deceive, so it is reasonable to assume that this demon is deceiving as well. As we saw, this argument is much stronger than standard skeptical ones. But here I wish to point to a third level of the argument, which directly attacks determinism: (c) The determinist argues that a long and impressive list of human insights and intuitions are nothing more than an illusion. Some of these have already been mentioned, like the illusion of free will, the illusion of moral responsibility, the illusion of weakness of the will, the illusion of the existence of moral or aesthetic norms, the illusion of commitment to these and the illusion of moral condemnation and praise, and we will later look at some other such “illusions.” Thus, this list is excellent positive proof that there is some demon deceiving us. And not only do we know there is a demon (= a coercive system), but we also know that it deceives us in many areas. So what is the basis for our faith in it in other areas (= in our system of recognition and thought)? Why should we trust it when it provides us with the causal-deterministic intuition? After all, we’ve already proved it deceives us. In terms of the distinction we made above, argument a is not hopeless but it is without foundation. Argument b is not without foundation but both options are still open in terms of the odds. Argument c, presented here, is already a very reasonable conclusion, to the point that its obverse (the faith in the sympathetic nature of the demon) becomes not only without foundation but also almost entirely without any chance of being true.

Another Argument From Another Angle: The Anthropological Argument One can perhaps present a similar argument against the determinist from a slightly different angle. If the world is indeed deterministic, then where did all our illusions of free will come from? Usually, we derive out insights and concepts from experience. Perhaps one could make generalizations and expansions at some level, but the basis is nevertheless our direct experience. We thus have to ask the question of where this idea and notion of free choice came from. Where did we experience such a mechanism? In the background there is also a question lurking about the mechanism of evolution regarding the survivability value of these illusions. But the anthropological question goes beyond this: How and when did these illusions sprout up? Does the very existence of the feeling of freedom not hint that we truly feel this freedom? This is a consideration similar to what Descartes called the “anthropological argument,” in which he proves the existence of God from the fact that we possess such a concept. In any event, it would certainly seem that this argument has some weight in favor of libertarianism. We will return to a version of this argument in chapter fourteen when discussing the illusion of choice.

On Thought Experiments Another a priori consideration against determinism is the Newcomb paradox. This is a thought experiment, which requires us to give a brief introduction on thought experiments, their validity, and their significance. Prima facie, it is not clear why we should make factual statements based on thought experiments (physicists, following Einstein, call it the Gedanken experiment, which is simply the German translation of the term “thought experiment”) or even a priori considerations in general. It’s important to understand that both determinism and libertarianism are factual claims, arguing something about the world itself, and this is true even if we conclude that we have no empirical way of demonstrating the correctness of either approach. How, then, can a thought experiment decide questions about the world without any observation thereof? In the period of ancient Greece, science was based on hypotheses and logic. Aristotle concluded that objects fall to earth at a speed appropriate for their mass (or weight), and that a heavy object will fall faster than a light one. This “logical” statement lasted some two thousand years, until Galileo came in the sixteenth century and simply tested it. He threw objects from the top of a tower, and thus, with a surprisingly simple experiment, he demonstrated that Aristotle made a basic factual mistake, thus ushering in the modern scientific age. This marked for physicists and scientists in general the importance of observation and lab experiments. It would seem that R. Eliezer’s cry/warning in the Babylonian Tractate of Brachot (28b): “Deny your sons from the logic” was further validated. We should therefore be wary of statements based on logic alone sans observation. This connects to the philosophical difficulty, which we already pointed to at the beginning of the chapter, which arises from making factual conclusions based on conceptual analyses and arguments. However, we do know of a number of thought experiments in the world of physics and philosophy. Plato’s Parable of the Cave is already a sort of thought experiment. Among the modern experiments in physics, we know of Maxwell’s Demon, who goal is to challenge the Second Law of Thermodynamics; Schrödinger’s Cat, whose goal was to challenge the concept of superpositioning in quantum theory, the Einstein–Podolsky– Rosen paradox (EPR), which challenges other elements in quantum theory (locality and causation), and so on. 68 Perhaps the most well-known thought experiment in the modern era is a thought experiment “conducted” by Einstein to prove the principles of the Special Theory of Relativity (and effectively to refute the principles of Newtonian mechanics). It is the most well-known because it was also one of the earliest thought experiments in modern physics, having been done at the beginning of the twentieth century, but primarily because it has far-reaching results and factual consequences, one might even say amazing ones. It’s very surprising to discover that a rich physical theory, so fruitful and reliable, like the theory of relativity, is based on an experiment conducted not in a lab but in the thinking of a physicist (Einstein, in this case), and that a rich and reliable and no less empirically verified theory, like Newtonian mechanics, was subsequently refuted. To demonstrate, I will bring one version (of many) for Einstein’s experiment, and we will see how factual conclusions derive from it.

The above schematic shows a car moving along the road at some fixed speed. When it passes point A, the driver throws a ball upward. The ball may have been thrown upward but the car is moving forward and it therefore gives the ball an additional component of speed in the car’s direction of movement. An observer from the side will see the path of the ball as it appears in the schematic: The ball will move forward at some angle (determined by the ratio between the speed of the ball and the speed of the car), and its speed will be calculated by a vector sum of its speed going up and the speed of the car going forward. This sum is calculated by the rule of the Parallelogram of Force described in the drawing below (for simplicity’s sake, we assume here that the speed of the ball is the same as the speed of the car):

In other words, if the car’s speed is 100 kmh, for instance, and the upward speed of the ball relative to the car is also 100 kmh, then the overall speed of the ball will be 100 X √2 (a little over 141) kmh. So far, we described the sending off of the ball. Now we will see in the above diagram that there is an elastic wall later in the ball’s path which returns the ball back down at the same speed and angle, and therefore the ball ultimately reaches point B. Who will reach point B first, the car or the ball? They will of course arrive at the same time, as the horizontal speed component of the ball always keeps it above the car (both of them are moving forward at 100 kmh), until it descends back down to it. We saw that the car moves at 100 kmh, while the ball is moving at a speed of 141 kmh. How then can it be that both of them will make it to point B at the same time? Simple: The ball has taken a longer path (the sum of the two sides in a triangle) from the car (which only went along its base). Therefore, everything works out fine according to Newtonian mechanics. Now, Einstein tells us, let us think of a similar experiment with a beam of light, and for simplicity’s sake we will discuss a car also moving forward at the speed of light. The car sends a beam of light upward at point A, it then hits a mirror (instead of a wall), which returns it to point B. We now ask again: Who has gone the farthest distance? Obviously the light (the two sides of the triangle). Who has moved the fastest? Here the answer is different than the previous question: The light moved at the same speed as the car (this is the principle of the fixed speed of light). So the light should not make it to point B along with the car. All this was described according to the reference point of the stationary observer (standing on the sidewalk). However, if we look at the reference point of the car, as far as the driver is concerned, the light is moving directly upward at point c, and goes down at the same speed. It follows the car, and it is therefore clear that when it comes down it “lands” directly on the car (which has meantime already gotten to point B). In other words, as far as the observer in the car is concerned, he and the ball did make it to point B at

the same time. So who’s right? The stationary observer or the driver of the car? Do they make it together, or one after the other? Einstein’s brilliant answer is that they are both correct. The time which passed as far as the sidewalk observer is concerned moves at a different speed than the driver of the car. Therefore the simultaneity of the two events, with both the car and the light reaching point B, has different significance depending on the reference point. What is simultaneous for the stationary observer is not for the driver of the car, and vice versa. We can see this more concretely if we look at the right side of the triangle from the above diagram.

This triangle describes half the path taken by the beam of light, and therefore covers half the time of the process. In this case, the base of the triangle is shorter that the vertical side, as the speed of the car (V) is lower than that of the light (C). The triangle is a right-angled one, and we can therefore apply the Pythagorean Theorem to it, which determines the relationship between the sides of the triangle. The length of the hypotenuse is the speed of light (C) times half the time which has passed in the system of the stationary observer (t). The length of the base is the speed of the car (V, the speed of the moving observer relative to the stationary one) times the stationary time. The second side describes the path of the light for the stationary observer (as far as he’s concerned the light went up and down in a direct pattern, at the same speed — C); its length is the speed of light times the time of the moving observer (t’). The Pythagorean Theorem gives us the following ratio: (ct’)2 = (ct)2 – (vt)2 If we simplify the equation and extract the root, we will get the following:

This is the formula of the special theory of relativity regarding the rate of time flow in the system of the moving passenger and the rate of time flow in the system of the stationary passenger (known as the “expansion of time”). While it’s true that we made assumptions here regarding the fixed nature of the speed of light, according to which the speed of light does not behave like the speed of a ball, yet still the experiment, entirely conducted by thought and not in the lab, succeeded in showing that the match between the various details of factual information and logic require a change of conception (or paradigm, to use the term of philosopher of science Thomas Kuhn). Thus, in a simple experiment, Einstein broke the principles of classic thinking which laid at the foundation of Newtonian mechanics (without anyone even noticing). These were seen as obvious, until Einstein came and showed that not only are they not obvious, they are actually incorrect. With his thought experiment, Einstein also found the alternative — one of the most revolutionary results in the history of physics. The result was an equation which serves physicists in entirely practical consequences, and which is supported by a great many experiments. It may have originated in thought, but the result is a factual claim which contains a great wealth of information about the world. The conclusion from all this for our purposes is that thought experiments have a great deal of importance even in the modern scientific era. They show us the coherence or incoherence in the relationship between assumptions and/or our facts, and sometimes uncover hidden assumptions we didn’t even know existed, and even deal with them.

Now, after we have become convinced that there is room for thought examinations of factual claims even in the scientific era, we can now move over to another famous thought experiment, which deals with the question of determinism.

Newcomb’s Thought Experiment In the previous chapter, we dealt with the argument of the idle individual, who according to the determinist should make no effort to achieve any goals. We demonstrated this with an individual with endless predictive capabilities who wrote the grade a student will receive in a future test on a piece of paper. Since the grade on the note is necessarily true (because this person has perfect precision in his predictions), there is no point in studying for the test since the grade is predetermined. We face a similar (in fact, identical) situation with Newcomb’s Paradox. Here, too, two people are involved: The chooser and the prophet. The chooser has two boxes in front of him: One is open and contains $1000, and the second is closed contains either $0 or $1,000,000 (the chooser does not know which option is correct). The chooser must decide whether to take the content of both boxes together, or only the content of the closed box. Now the prophet enters the game (we assume his predictive powers are perfect). He knows in advance and with certainty what the chooser will do, and based on this he prepares the contents of the closed box the day before the chooser makes his decision. Since the prophet is interested in giving a prize for those who suffice with less, he adopts the following tactic: If the chooser is about to take both boxes, he will see to it already now that there will be no money in the closed box. But if the chooser is an individual who settles for less and is set to only choose the closed box, the prophet will take care to reward him and put the million dollars in that box. The chooser himself knows all this. He knows the capabilities of the prophet, as well as his strategy. What the chooser doesn’t know is one single fact: What the prophet predicted regarding his future decision now (as he himself is not a prophet), and thus he also does not know what the prophet put into the closed box. The question is this: What strategy should the chooser adopt? Ostensibly he should only take the closed box, as this will earn him a million dollars, while if he takes both, he will only win a thousand (as the closed box will be empty). On the other hand, the box is already closed before him, and what is inside was already determined yesterday and will not change if he acts differently. Why, then, should he not take the closed box along with the open box with another $1000? After all, taking the second box cannot retroactively change the content of the closed box (our assumption here is that the prophet is all-knowing but not necessarily all-powerful). Thus, it appears that the taking of both boxes is the winning strategy in any event. But if the prophet is indeed all-knowing, he can see to it that in this case, the closed box will be empty. Thus, when the chooser takes both of them, he will win only $1000, having chosen wrongly. When this paradox was published, Newcombmania emerged, a real enthusiasm of people thinking on the matter. Philosopher Robert Nozick, in his article in 1969, which presented the Newcomb Paradox to the public, reported that almost every individual had a clear position on the question, and almost everyone thought that those who hold the opposite position are complete fools. But there is certainly no consensus on what the proper position is. I would thus not be an exception in this respect. I also have a clear position on this question. But for our purposes, it’s more important to stress that the assumption that there is such a prophet assumed a hidden deterministic approach. In the libertarian worldview, one cannot speak of such a prophet, as one cannot know for certain what the chooser will decide. If the information does not exist yet, how can one know it for certain? “To know information which does not exist” is an expression with an internal contradiction, thus a prophet who predicts the decision of an individual who freely chooses in advance is an oxymoron. This isn’t a question of the quality of his predictive powers. Predictive powers are defined where the information exists in principle and which can be derived from the present circumstances. However, in the libertarian picture, the information does not exist at all yet, and therefore no-one, no matter how good their predictive abilities, could ever predict the future (above, we claimed that God Himself cannot do so). By contrast, when it comes to the deterministic picture, the question stands. There we have an assumption that the information does exist now. Even if no-one can reach it due to its great complexity, and indeed no computer could predict it, no matter how strong, let alone a prophet, the information is there. There is therefore nothing preventing a prophet from knowing it in principle. This is a supercomputer, which may not exist today, but the concept does not contain a contradiction, as the computer knows information which does exist. Therefore, this argument is in effect a thought experiment conducted according to the assumptions of the determinist, no different than the thought experiment conducted by Einstein according

to Newtonian assumptions, which he used to refute them. We therefore see that the Newcomb Paradox is in effect an a priori argument against determinism, as if determinism is right then the information does exist and such a prophet is possible in principle. If so, by his lights, even though the content of the closed box is now fixed and unchangeable, it is still preferable for the chooser to forgo taking the open box and suffice with the closed one. Taking the open box with the extra thousand dollars will make the million dollars in the closed box “disappear.” Thus, according to the determinist, the chooser must make an utterly irrational decision (not take the thousand dollars in front of him), or conversely believe in causal influence working backward in time (that the taking of the box will retroactively change its content). This proves that determinism is entirely unreasonable.69 Some have sharpened this challenge by expanding Newcomb’s Paradox to include a glass box. If the closed box is made of glass, the chooser sees what’s inside it. If he sees it has a million dollars, he understands that the prophet predicted he will only take that. So now he will take the two boxes to earn another thousand dollars, thus contradicting the prophet’s prediction. If he sees the box is empty, then it is clear that the prophet predicted that he will take both. So to annoy the prophet, he may take only the closed box. In either case, his actions are the opposite of what was predicted. This is in effect proof that there cannot be a prophet who predicts the choice the chooser makes. Conversely, acting against the prediction of the prophet will simply magically change the content of the closed box, and then the prophet’s predictions will come true. Except in this case, we may have saved the possibility of prediction (= determinism), but at the cost of adopting an approach which allows causal influence backwards in time, from the future to the past. But this expansion is problematic, as according to the deterministic assumption, the chooser cannot change his choice. Even if the box contains nothing, he is compelled to take both boxes, and vice versa. Ostensibly, this is the solution to the original paradox, but another look above shows this not to be the case, as we will now see.

Further Clarification of the Newcomb Paradox Prima facie, it would appear that this experiment does not contain an argument against determinism, but at most demonstrates that freedom of choice does not fit with determinism, thus constituting an argument against compatibilism. The determinist shouldn’t panic at this challenge, as in the deterministic picture, the chooser doesn’t even have the option of freely deciding between the options. He will actually do what the prophet predicts, as he has no other option. This is in effect an obvious consequence from the deterministic world view, and it shows that one can remain a proper or at least consistent determinist. But another look above shows that the determinist’s problem remains in place. Our problem is not the question of what the chooser will do in this situation, but rather what is right for him to do. This is a mathematical question of strategy, belonging to the world of game theory, and not a question of the chooser’s psychology or what will happen in real life. Even in the deterministic picture, which argues that none of us have freedom of choice, we can ask the question: On the assumption that a creature in some imaginary world has freedom of choice, what would be best for him to do? This question has no answer, and this is the problem which the deterministic picture produces. The determinist may be able to be saved by arguing that that the question is not defined. Freedom of choice is a concept which contains a contradiction at the conceptual level, and therefore for him the hypothetical question of “the right strategy” cannot even be asked, to say nothing of the practical question of what any individual would do in that situation. But this is an unreasonable argument. For the sake of argument, I am willing to accept that it’s hard to define freedom of the will. I am also willing to assume for the sake of argument that perhaps such freedom does not exist for any creature, including human beings, in our deterministic world. However, the claim that such freedom is not possible at all at the logicalconceptual level, involving a logical problem and not just a physical problem, is a very unreasonable argument lacking a real basis. This would seem to me to be too heavy a price to pay to dogmatically adhere to determinism.70 One can put this a little differently: Even the determinist does not reject human rationality, or that individual human beings make decisions according to the best of their understanding and judgment. He only argues that this process does not reflect free choice and decision-making, but rather only a sort of calculation compelled on him. But if this is the case, the problem for his approach stands: In our case, we have a prophet who knows in advance what was chosen, and given this information it is clear that the chooser is supposed to use it to “calculate” his decision rationally. But what is the rational decision he is supposed to make, deterministically? Ostensibly, he should only take the closed box to win the million dollars. On the other hand, as a rational human he knows that the prophet cannot change the content of the box after the choice, and it is therefore better that he take both

boxes. So, to solve the problem, the determinist must now retreat one step and declare that the individual is compelled to act irrationally. Is the determinist willing to reject human rationality? If so, it’s not clear why he treats the considerations which lead him to his own deterministic position with such seriousness. Maybe he was compelled to arrive at those conclusions even though they are illogical?!

The Connection to Fatalism and Idleness As noted above, some have pointed to the connection between the Newcomb Paradox and the idleness argument. The later argues that since there is a piece of paper stating the grade the student will get on the test tomorrow, the idler therefore has no reason to study because the grade is set. Now we can see that in a deterministic world, he has two choices: be idle and not act to achieve any goals, or assume that studying for the test will retroactively change what the note says, adopting a position allowing causal influence backwards in time. In such a situation, if our idler is rational, he won’t do anything. Conversely, he can give up the rationality of the idler and act even though he has no reason to. We also saw above how determinism leads to the forgoing of the belief that human beings are rational animals.

The Connection to Logical Determinism Logical determinism points to an ostensibly similar, but effectively entirely different mechanism. Yesterday there was a note that said whether there will or won’t be a naval battle. From here, the logical determinist wanted to conclude that the occurrence of the naval battle must necessarily happen — it cannot not happen. Our response is that the occurrence of the naval battle is what retroactively writes the note. What was written on it a thousand years ago is determined by what happens now. Why is this possible in the context of logical determinism (even according to the libertarian), while it cannot be done by Newcomb’s prophet? Why can’t the determinist argue that here, too, taking the two boxes will retroactively change the content of the closed box? The difference is very simple. In the context of Newcomb, we are speaking of a note that contains information, or a box that contains money. Such a note cannot be written before the event, which is why I stated above that no such prophet exists. But I also explained above that the truth value of a claim is not information. The truth value of the claim “tomorrow there will be a naval battle” was always true, but that doesn’t mean the information about the occurrence of the naval battle always existed. As I explained, the truth value is our logical definition, not information or a fact about reality. As such, it is non-temporal or atemporal. But information and the presence of money in the box are physical facts and hence not atemporal. Therefore, the truth value of a claim describing an event is not a causal effect so it can be created retroactively, but the information about an event can be created only once the event occurs and not before, otherwise it’s a reversed causality.

The Connection to Theological Determinism If the Newcomb Paradox did indeed prove that no such prophet can exist, it should be seen as proof that despite His being all-powerful, it cannot be that God Himself will know in advance what any given individual will do in the future. To see this, we need only include God Himself as the prophet in this game, and allow Him to arrange the closed box in light of His knowledge of the future (God is also all-powerful, so therefore we must add the limitation that He accepts on himself not to change the content of the box after He organizes it. Conversely, He could hand the information to some individual, who functions as a prophet, and he will determined the content of the box in accordance with the information he has). This of course leads to precisely the same problem: Even God Himself cannot know what the future choice will be. As noted, this is not possible at the level of principle, and therefore does not point to a flaw in His being all-powerful. He simply “can not” do the impossible. Just as He cannot create a triangle that is a circle, or a bachelor who is married, so He also cannot know information which does not exist. Thus, Newcomb’s argument is proof against theological determinism as well. One cannot reconcile our freedom of choice with Divine foreknowledge. God’s being all-powerful changes nothing in this respect. The significance of the rejection of theological determinism is that the assumption that God knows everything in advance is incorrect. This does not contradict the assumption of Him being all-powerful, as foreknowledge of an event dependent on choice is a contradiction in terms, like a triangle in the form of a circle. The inability to create contradictions or act in a manner which contains a contradiction is not true inability — the inability here is logical, not physical. Here we once again see the paradoxical nature of this assumption from another angle. It is true: The Newcomb Paradox pulls the rug from under all sorts of evasive arguments on God’s knowledge of the future. The logical conclusion is that He indeed does not know it (at least if we don’t give up the libertarian picture).

Thus, Newcomb’s thought experiment rejects physical (so long as we preserve the assumption of human rationality), logical and theological determinism.

Summary In this chapter, we saw a number of a priori arguments in favor of determinism, most of which are fairly easily rejected. By contrast, we saw two types of arguments in favor of libertarianism, both of which point to a serious internal contradiction in the deterministic approach: The Fool’s Cycle and the Deceiving Demon, with its three levels, as well as Newcomb’s Paradox. This series of arguments is added to the fact that every one of us has a strong intuition regarding the freedom with which we act and the free manner in which we make decisions. Ostensibly, the situation is entirely in favor of libertarianism. What, then, leads so many thinkers to nevertheless cleave to determinism? To answer that, we will discuss the principle of causality and its significance for our analysis in the next two chapters. 64 British philosopher who operated in the first half of the twentieth century and who was one of the founders of analytical philosophy. He espoused the virtues of common sense and waged unrelenting war against philosophical idealism, an approach which argued for skepticism regarding the existence of the outside world. 65 American Christian thinker, one of the greatest analytical thinkers in our time and one of the most prominent philosophers of religion. 66

This

is

his

article

entitled

“Naturalism

Defeated,”

available

here:

https://www.calvin.edu/academic/philosophy/virtual_library/articles/plantinga_alvin/naturalism_defeated.pdf 67 Plantinga here quotes the well-known biologist J.M. Smith, who raised these arguments in Time magazine in the December 1992 issue. 68 One can propose an experiment in which the tester uses scans of brain signals to predict the choice of the test subject in a situation described by the Newcomb Paradox. We will discuss a series of experiments in chapter fourteen which use neural signals to predict behavior. 69 One can propose an experiment of the sort noted in note 16 above; in chapter fourteen, we will discuss Ariel Furstenberg from Hebrew University who spoke of how they are presently working on developing such experiments. 70 The seventh chapter is devoted to proving the possibility of freedom of will, showing that it does not involve a contradiction in terms. In a footnote there, I note that this proof raises serious questions as to the possibility that the determinist can defend himself from the Newcomb Paradox as presented here.

Chapter Five ON THE ORIGINS AND MEANING OF CAUSALITY If human will could be free, that is any individual could truly act as he wished, then history was a series of unrelated events.71 – Leo Tolstoy

In this chapter we will describe and define the causal relationship, so that afterwards (primarily in the next chapter), we will be able to examine the significance of causality for the question of determinism. The chapter is divided into two parts: The first deals with the definition and characterization of causality in general, while the second part deals with the causal conception of modern physics. I will note here that the present chapter is a little “thick,” as it also includes formal analysis. Those who have difficulty reading it can skip to the summary and from there to the next chapter. 71 I assume there is no need at this point to explain Tolstoy’s serious category error — he identifies freedom with liberty. I use his words here solely as a literary flourish.

A. What is Causality? The Origin of the Intuition of Causality It is a simple feeling among every individual that things do not happen without some sort of reason. When I hear a voice thunder behind me, I immediately turn around and search for the source of that voice. I do not even take into consideration the possibility that maybe that voice is uncaused thunder, which would mean I have no reason to search for the cause behind me. When some object begins to move, I immediately search for the force acting on it to move it. This is the principle of causality: There is a cause for everything that happens. Where did we get this feeling? Many people believe that our experience is what teaches us the principle of causality. Oftentimes in the past, we saw that when a log is placed in the fire, it burns, and when someone kicks a ball, it flies upward, and when someone blows a trumpet, it creates a sound, and so on. In addition, we never heard a sound created without a source and we never saw a ball go into the air on its own. It would appear, then, that a generalization based on this collection of observations is what produced the principle of causality for us. But is that really the case? Eighteenth century Scottish philosopher David Hume rejected this argument. He claims that there is no way to learn the principle of causation from experience. When we observe a log in the fire, we may see that it is always burned, but we have no way of observing the causal relationship between the log’s presence in the fire and its being burned. What we can see by direct observation is at most a temporal relationship: Logs are always burned after they are placed in the fire. This is a time-based relationship between events, but not necessarily a causal one. We have no way of showing that the fire is the reason that the wood turned to ash. This is the same when we kick a ball and see it go into the air, or any other causal relationship between two events. In fact, if we think about it, we will discover that the fact that we never saw an event which occurs without a reason is not quite accurate. It would be more correct to say that we always attributed some reason to every event we saw. Thus, even if we saw an uncaused event, our assumption was always that there must be some reason for it, even if one we don’t grasp right now. In other words, causality is an assumption of our way of thinking and intuition, and not necessarily the result of something we learned from experience. Hume himself offered a solution for his conundrum: We indeed have no way of determining a causal relationship between events. We can at most say that the two occur one after the other, and no more. Does that sound reasonable? Immanuel Kant, who we’ve already met, thought this sounded entirely unreasonable, and it would seem that most people think the same way. Moreover, it was apparently very unintuitive for Hume himself, as if this was his simple attitude towards the causal relationship, why did he even bring it up? Obviously, only the fact that one cannot derive the causal relationship from observation is what led Hume, a sworn empiricist in outlook, to the conclusion that there is no way to determine causal relationships at all.

The Three Elements of the Causal Relationship To clarify why this argument is utterly unreasonable, and also for our own needs later on, we have to go a little deeper into the definition of the various components of the causal relationship.72 As we will see, the causal relationship contains a number of components, and failure to distinguish between them may lead to serious misunderstandings when it comes to the concepts of causality and time. Here we will present the three central components of the causal relationship: The temporal component, the logical component, and the physical component.

1. The Temporal Component The relationship between causality and time is axiomatic. The simple assumption is that a cause must always appear before the effect, and it certainly cannot appear afterwards. Among physicists, when we say that there is a causal relationship between event A and event B, this usually means nothing more than that A was prior to B.73 As I already noted above, David Hume found an identity between these two expressions even at the conceptual level, meaning that he believes causality is nothing more than a necessary temporal order of events.

In philosophy, in fictional literature, and sometimes even in modern physics, some toy with ideas of reverse causality, in which the cause comes after the effect. The principle of causality may contradict this, but there are mechanisms in quantum theory which ostensibly enable such a thing to happen.74 For the last hundred years, a lively debate on the matter has taken place among physicists as well as philosophers. We will note that when we speak of “reversibility” in time of physical processes, we do not mean to argue that the cause appears after the effect, but rather that if event A can appear before event B (to be its cause), then according to the laws of physics a process will always be possible by which event B appears before event A (and now B might be the cause of A).75 The question of what this says about the causal relationship between the two events is not a simple one, as causation is a directional relationship, and it is therefore difficult to accept causal terms which are symmetrical in time. Obviously the flight of a ball is not the reason for the kick it gets afterwards. However, when you neutralize the element of causation from the causal dimension, and base this relationship solely on temporal precedence, as Hume did, then the determination is not as strange as it sounds. Indeed, in the second part of this chapter we will see that physics does not explicitly show a causal relationship between events; it deals with the description of events along the timeline, not in defining the causal relationship between them.76 Many question the assumption that the causal relationship assumes the temporal precedence of cause to effect, by force of the temporary symmetry of physics. Since the laws of physics do not differentiate between time which moves forward and time that moves backwards, argue the challengers, it cannot be that we will define causality on the basis of the assumption of temporal order, in which the cause is necessarily prior to the effect. This argument is very common, and we will deal with it in depth in the second part of the chapter.

2. The Logical Component There is also a logical relationship between cause and effect, as the cause is the condition for the effect. What does this relationship mean? We can say that if the cause occurred then the effect necessarily will as well (there cannot be a situation in which the cause occurred and the effect — or not). This means that the logical relationship between cause and effect is one of sufficient reason. The question of whether or not it is a condition which is also necessary or merely a sufficient condition is a matter of dispute among philosophers.77 Some claim that this is a necessary and sufficient condition, meaning that whenever there is an effect there obviously is also a cause. In other words: There cannot be another cause which engenders that effect (the cause must be singular). On the other hand, others argue that the logical relationship does not need to be necessary, only sufficient. In their view, there may be a situation in which two different causes produce the same effect.78 This debate will be discussed below. Whatever the true nature of this logical relationship, this is a phenomenological description, not a physical one. The argument is external to the events and is not related to the essential relationship between them. The logical relationship makes the following argument: If event A took place then event B did as well. It does not touch on the question of whether there is a relationship of content between the two events, in which A caused or engendered B. This is in effect another angle of the distinction between physics and logic which we discussed above in the context of logical determinism. We will further sharpen the issue later in the chapter.

3. The Physical Component The third component of the causal relationship is physical in character, and it determines that there is a causal relationship between cause and effect. This means that the cause does not just appear before the effect, and is not just a logical condition for it (i.e., without the cause there would be no effect), but that the cause is what engenders the physical effect itself. They have an essential-content relationship of physical interaction. If we discern two events which have a logical and a temporal relationship between them, but it will be clear to us for some reason that there is no causal relationship between them, we will not define them as cause and effect. We already noted the work of American logician Raymond Smullyan and his The Tao is Silent above. In chapter 38 of the book, he addresses astrology and offers to explain it based on a theory from the school of famous psychologist Carl Jung and Nobel laureate and physicist Wolfgang Pauli, known as synchronicity.79 Even though I am fairly convinced that astrology has nothing in it, we will examine this proposal as an example of the logic of synchronicity. Astrologists are often asked how a star, which is many light years away from us in distance, affects our lives. This contradicts the laws of physics, as there is a limit on the speed of physical influence, and if a change in a star affects our lives in some way, this change must have

occurred long before we were born. These are attacks on astrology based on the principle of causality.80 The astrologers, at least those who bother to answer such questions, usually give vague responses, which a rational person tends to summarily reject, and justifiably so. The logic of synchronicity, by contrast, argues that the stars really do not affect me, but for some reason there is a correlation between the structure of the stars in the sky when I was born and the history I will have, or between what happens in the sky and what will happen at the same time here on earth. This correlation takes place without any mutual influence; theoretically it may be that there are parallel processes in the sky and on earth whose occurrence is based on equivalent rules, and therefore the correlation is always preserved. Such a mechanism is of course not vulnerable to attacks based on causality of the sort we described above.81 In his book, Smullyan brings the story of a zen master who was deep in thought, and suddenly felt danger and turned around. He saw behind only his young servant and nothing more. In the end it turned out that the youth did indeed think at that precise moment on the possibility of stabbing his master with a sword. The explanation provided by Smullyan, who does not allow for the possibility of mind-reading, is that this event which caused the youth to think his thought at that moment is what caused the teacher to suddenly feel danger. The event could have occurred earlier, or at the same time, but this is a third factor which created a correlation between them. Again, the thought of the youth was not the cause of the teacher’s fear, but there is a correlation between them for some external reason.82 The conclusion is that even if we cannot point to the mechanism of influence between the two matching systems, this does not necessarily mean we must reject the fact that they do indeed match. True, basing this description on some coincidental miracle is not reasonable (even if possible in principle), but it may be that the correlation was formed deliberately, and not by coincidence, by a factor which constantly coordinates the actions of the two systems, or at least ensures that the rules by which they act are equivalent to each other. If we adopted such a picture regarding the stars, assuming that the laws which dictate their behavior are parallel to the laws which control what happens to us, would it be correct to see the stars as the reason for what happens down here? It turns out that this is not the case. This is ‘synchronicity’ or correlation, but not a causal relationship. What is the difference between synchronicity and a causal relationship? The synchronistic relationship contains the logical component (if A happened, then B happens), and perhaps also the temporal component (B after A). These two exist in the example of astrology, as well (if we accept it). What’s missing there is the physical component, as according to this explanation, the stars can serve us to predict what happens down here, but they do not engender what actually happens, and therefore the relationship is not a causal one. This is a good demonstration of the necessity of the physical component in defining the causal relationship. In the book by Arthur Koestler, The Roots of Coincidence, which discusses the synchronistic theory, the author mentions how Jung and Pauli were preceded by an Austrian biologist named Paul Kammerer. He was one of the last advocates for Lamarckism, and he had quite a bit of influence on Jung. Kammerer attributed a great deal of significance to coincidences, and he decided to carry out a systematic study of this phenomenon. As part of his study, he carefully listed and divided up the different types of coincidences which occurred to him from age twenty to forty. Kammerer’s main argument was that there was an “attraction” between similar events in some senses (which he lists there) along the timeline. The coincidences we encounter directly are just the tip of the iceberg of a very large scale phenomenon of similar events happening close to each other along the timeline. He compares it to the force of gravity which attracts masses to each other, arguing that there is a similar attraction between similar events. He published his findings in his book Das Gesetz der Serie. Jung and Pauli also claimed that the causal explanations are just a small part of understanding the ways in which nature operates, and that the synchronistic explanations are mechanisms which are no less important for understanding it. According to them, some of the events take place because of causes, and the others — based on “synchronistic attraction.”83 Koestler himself notes there of the similarity between this “attraction” and another principle of Pauli in the field of physics. Pauli is known primarily thanks to the exclusion principle which bears his name, which states that in certain types of particles (known as fermions, after the famous physicist Enrico Fermi), two different particles cannot be in precisely the same quantum state.84 To the best of my knowledge, Pauli’s exclusion principle has no explanatory mechanism in today’s physics, meaning we know of no force which ensures that the particles will never be in the same position. This is an ad hoc principle, and is therefore quite similar to the principle of attraction between similar events in Pauli’s synchronistic theory (Pauli’s principle speaks about rejection, not attraction). In any event, synchronistic relations make it clear to us why the physical component is so vital for the

causal relationship. Synchronicity is not causation, as it lacks the physical component, which directly engenders the effect. This is precisely the problem in Hume’s proposal to remove the physical dimension from the causal relationship and suffice with only the two remaining dimensions, the logical and temporal. Here, we saw why this proposal is unreasonable. Causation without actual causing is not really causation.

The Relationship Between the Logical Component and Causation: The Initiation Paradox As already noted, analytical philosophers have discussed the character of the logical component of the causal relationship. We must note that a necessary and sufficient logical condition has two characteristics: 1. Symmetry. If A is a necessary and sufficient condition for B, that means that B is a necessary and sufficient condition for A. In logical terms, necessary and sufficient means “if and only if.”85 2. Uniqueness. A necessary and sufficient condition is always singular. There are no two events each of which is a necessary and sufficient condition for that third result. Put differently: If A is a necessary and sufficient condition for C, and B is also a necessary and sufficient condition for C, then A and B are the same event.86 Yuval Steinitz, in his discussion on causality in the second part of his book The Tree of Knowledge (Hebrew),87 wants to prove the thesis that a cause can also be a sufficient but not necessary condition for the effect. To do so, he proposes what he calls the initiation paradox. To prove his argument, he starts from the opposite statement and shows that it leads us to an utterly unreasonable conclusion. Let us assume, Steinitz says, that the cause is indeed necessary and sufficient for the effect. It therefore necessarily results that every causal chain is at most made up of two links, meaning that there cannot be a causal chain containing more than one causal step. There cannot be, for instance, a chain in which A is the cause of B which is the cause of C. Why not? The characteristic of symmetry shows that in such a chain C is also a necessary and sufficient condition for B. Now we have a result in which both A and C are necessary and sufficient conditions for B, and this contradicts the principle of uniqueness (unless they are the same event). There can therefore be no causal chain with three links. But this is of course not possible, as this means that the causal chain will never be able to move forward. Events in the world would have to be divided into cause-effect pairs which come from nowhere and go nowhere. This also means that at least half the events in the world occur without prior causes, a contradiction of the principle of causality according to which nothing happens without a cause. Steinitz concludes that in order not to reach such an unreasonable picture, which is contrary to the aforementioned picture of causality, we must give up the assumption that the cause is a necessary and sufficient condition for the effect. His conclusion is that the cause needs to be sufficient, but not necessary, for the effect. But Steinitz’s argument is incorrect. Intuitively, there is no doubt that a chain of events with a causal relationship between each other is possible, and here Steinitz is of course correct. But at the same time, there is also no physical barrier to event A being the only cause of B with B being the only cause of C at the same time. Is there a limitation of the laws of physics, determining that there cannot be a chain of three events each of which is the sole cause of what follows it? Of course not. Let us assume that only rubbing a match lights fire (there is no other cause that can do so). Additionally, let us assume that only fire can burn a forest (there is no other factor that can do so). Does such a picture contain a logical contradiction? According to Steinitz — the answer is in the affirmative, but that is of course completely absurd. The picture presented by Steinitz does not properly describe the world. But his argument seems logically valid. What’s wrong with the argument? The mistake in Steinitz’s argument is his inattention to the collection of components we pointed to above, which together form the causal relationship. A necessary and sufficient condition is indeed singular. This means that if A is a logical and necessary condition for B then B is also a logical and necessary, and therefore singular, condition for A. But as we saw, the causal relationship contains other components in addition to the logical one. The paradox of initiation ignores the temporal and physical components of the causal relationship, and treats causality as though it were nothing more than a purely logical relationship.

The description in the above sections leads us to define the causal relationship in the following way: A is the cause of B, if and only if the following characteristics obtain: 1. Temporal: A is prior to B in time. 2. Logical: A’s occurrence is necessary and sufficient (I assume this here to show where Steinitz is wrong) for the occurrence of B. 3. Physical: A is the factor which engenders B in the physical sense. If we return now to Steinitz’s consideration, we will see that it fails at once. This complex relationship between A and B is not symmetrical, despite the symmetry of the logical component of this relationship. This means that if A is the cause of B in this sense, then B is necessarily not the cause of A. We can see this via the temporal or the physical component. After all, if A is the cause of B, then A is prior to B in time, and therefore B is clearly not prior to A in time. If so, B cannot be the cause of A. The situation is a little more complex when it comes to the physical component, as we will see in the second part of the chapter, as one could in principle conceive (even if unreasonably) that they would both be engendering causes, each towards the other. But the temporal component definitely breaks this symmetry. Steinitz refers to the causal relationship in its purely logical sense, and therefore he fails. If the reader still thinks we still haven’t solved the problem, we will complete the refutation of Steinitz’s argument in the next section.

The Atemporality of Logic But now we must face the opposite question: If the conclusion that the causal relationship is not symmetrical is true, meaning that the argument that A is the cause of B is not the same as B being the cause of A then the symmetry of the logical relationship between A and B is also broken. After all, one of the characteristics of the causal relationship between A and B is that A is a necessary and sufficient logical condition for B. But the symmetry of the logical relationship shows that B is also a necessary and sufficient logical condition for A. It is true that B is not the cause of A, as it is not prior to it in time, but the logical relationship between them should maintain symmetry. Let’s look at it from a different angle: Think of an event Z which is the cause of A. This means that the three components of the causal relationship between them obtain, meaning that Z is prior in time to A, physically engenders A, and is of course a sufficient and necessary condition for A. Now the difficulty reawakens: the emerging picture leads to both Z and B being necessary and sufficient conditions for A, and that cannot be, as a necessary and sufficient logical condition must be singular. Steinitz’s argument would therefore appear to nevertheless be true. But this difficulty is also based on an error, and is in fact the same error which confused us above. There is a failure here to distinguish between the logical and temporal components of the causal relationship. According to the necessary and sufficient interpretation of causality, we must in this case say that Z is indeed the only cause of A, but the necessary and sufficient condition for event A is the combination of both B and Z. In other words: Every time we have an event Z followed by an event B, there will be an A between them. The reverse is also true: Whenever A occurs, Z will occur before it and B after it. But as we saw, the fact that the necessary and sufficient condition for A is Z and B does not mean that B must be prior to it in time, or even be its cause. It is a condition for it only in the logical sense, meaning that when it occurs then that occurs, and the reverse. The temporal relationship between them is as described on the chain: Z and then A and then B. We see that causal relationships are subject to the timeline, meaning that the causal relationship between A and B means that A is prior to B in time, but the logical relationships are not subject to the timeline. The fact that B is a necessary and sufficient condition for A does not mean it precedes it in time. This means that causality is related to time only via the temporal (and perhaps also physical) component, but not via its logical one. This conclusion connects to what we saw in the previous chapter on logical determinism, i.e., that logic is atemporal. We saw there that the truth value of a claim can be retroactively established by some future event. We can therefore say in general that logic is indifferent to time. The timeline is not relevant at the logical level, and this means that logical relationships can work in the opposite direction of the timeline. Another important consequence of this discussion is that the cause is at least a sufficient condition for the effect, and perhaps also a necessary one. Here we discussed the logical component to demonstrate the consequences of the identification of causality with logic. But for our purposes, it is sufficient to note that

the logical component of the causal relationship must be sufficient; the question of whether or not it is also necessary is unimportant for our discussion.

Back to Hume’s Challenge to the Physical Component We saw above that David Hume offered an analysis of the causal relationship, with his empiricist outlook leading him to the conclusion that it is solely a temporal and logical one. He said that if event A is the cause of event B, that means that A is always prior to B and is a logical condition thereof. Hume omits the physical component from the causal relationship, and this is precisely the problem with his system. But Hume was right that one cannot always learn of the causing component of the causal relationship from experience. Observing a log being burned in a fire does not reveal to us any causal relationship between the presence of a fire and burning. We see that the log is burned after being placed within the fire (the temporal component), and we see that it always happens (the logical component). But the physical component, the relationship of causation between the two events, is our addition to an analysis of the situation. That was not learned from experience. Hume, who as noted was an empiricist, concluded that this is therefore a fiction, and in his opinion the causal relationship does not include a physical component of causation. In effect, he identified causation with synchronicity. As we will see later on, his analysis has greatly influenced scientific thinking until today. Ironically, it is physics which has adopted the picture which sees the causal relationship as not containing a physical component. As already noted above, when physicists speak of a causal relationship, they are referring to temporal precedence, but not necessarily causation. The reason for this is that the equations of the laws of physics and their solutions do not indeed contain components of causation. As we will see below in the second part of the chapter, physics does not even know how to describe causation.

Fourth Intermezzo: Beginning with Crime and Ending in Compulsion To finish this section of the chapter, I will demonstrate the relationship between the different components of the causal relationship via a Talmudic discussion. A number of places in the Talmud refer to a situation where an individual starts by committing a crime and ends up in a state of compulsion. For instance, in the rules applying to guards, when an individual receives an object for safe keeping, he must take care to protect it properly. If he doesn’t do so, he risks having the object be damaged or stolen, and he will then have to compensate the owners. However, if he protected the object properly yet something still happened to the object, the guard is exempt from punishment, as he was under compulsion — something happened to him outside his control. For instance, an individual receives a wallet with bills to guard. He must guard the wallet carefully to make sure nothing happens to it. We know that if he puts it in the forest it might be burned, and that if he leaves it in an open house, it might be stolen. Let’s assume now that the guard places the wallet in the forest, and a fire does indeed break out there. In such a case, he was negligent in his duties and must now compensate the person who gave him the wallet. The case is the same if he left the wallet at home and it was stolen. But what if he left the wallet at home and the house burned down? That is compulsion, and he is exempt. But what if he leaves the house open, which is negligence due to the fear of it being stolen, but in the end there was a fire not a theft? A similar question arises when he buries the wallet in the woods (which is negligence in case of a fire), and in the end a fire doesn’t break out but the wallet is stolen. These two cases are known in halachic jargon as “beginning with crime and ending in compulsion.” The situation began with negligence because of the fear of event X, but ultimately the unexpected event Y occurs, which is itself defined as a situation of compulsion. Talmudic Tractate Baba Metzi’a 42a has a discussion of a case in which a guard receives coins as a trust, he buried them in a shack in the woods, and the coins were stolen. The gemara brings a debate there as to whether the guard is guilty or exempt from punishment. The ultimate legal ruling is that he must pay, and as the Talmud puts it: If it begins in crime and ends in compulsion — he is guilty. The Tosefot commentary there writes that this is the case only if the compulsion would not have happened without the crime. In the example brought there, if he hadn’t committed the crime and buried the wallet in the forest, it wouldn’t have been stolen. Therefore, he is considered as one who caused the theft and must pay. But when there is no connection between the crime and the compulsion, then the guard is exempt. We find such a case in Baba Metzi’a 36b, when the guard didn’t properly lock up an animal he was guarding. The beast then went to the lake and died naturally there (not due to encountering any obstacle, it simply died. This is known there as the “angel of death”). The assumption is that the animal would have died even if it had stayed in the shed; its time had simply come. In this case, it was ruled that the guard is exempt

from paying the owner, since the animal would have died regardless of how well he guarded it. As the Talmud puts it: “The Angle of Death [comes], what of him being here, what of him being there.” We should note that in case the animal had encountered some obstacle in the lake and died from that, everyone would then agree that the guard would be obligated to pay. This is a case in which there is a causal link between criminality and compulsion. The former caused the latter. In fact we are not at all dealing here with a case of compulsion, and therefore it is not defined as a situation of “beginning in crime and ending in compulsion,” as the aim of the guarding is precisely to protect the animal from just such a case. A more extreme case would be one in which the animal was stolen after the guard didn’t lock the shed properly. This is certainly a clear case of criminality (“It is entirely criminality,” as the gemara puts it), and everyone agrees he must pay. We will end with another halachic innovation. The system of the Rif is that according to one of the Amoraic figures of the gemara, Abayeh, the guard is also obligated to pay even in the case where there is no connection between criminality and compulsion. We don’t rule like him, but that opinion does exist in the Talmud.88 So far, we presented three kinds of cases: a. The guard got sloppy and buried the wallet in the forest, and a fire broke out. This is complete criminality (as the fire is prone to fires) and everyone agrees the guard must pay). b. The guard got sloppy and buried the coins in the forest, and thieves came. This is a situation of beginning with criminality (re: fires) and ending in compulsion (theft, which is not common when it comes to coins buried in the woods), and there is a dispute in this case. In practice, the guard must also pay in this instance. c. The guard got sloppy and left the door open and the animal left to the lake and died naturally, and indeed would have died had it not left, as well. In this case there is an irregular opinion that the guard must pay, but in practice all agree that he is exempt. These three situations express the three elements of causality which we spoke of here. To see this, we will examine the relationship which exists between the actions of the guard and the damage which occurred in his wake, in every one of the three cases. Case c is one where there is no connection between the criminal action of leaving the door open and the compulsion of the animal dying a natural death. The only component of the three which characterizes the causal relationship which is relevant to the relationship between these two events is the temporal one. The death of the animal occurred after the criminality of the guard. The logical relationship does not obtain here, as the logical relationship is always an expression of the type: “If… Then…” Here it is not right to say that “If the guard opens the door then the animal dies,” as the latter would happen regardless of the former. This means that there is no logical relationship between the two events. There is certainly no physical relationship (i.e., of causation) between them, as it is simply wrong to say that “The animal died because the guard opened the door.” What about case b? Here, too, we of course have the temporal relationship, as the theft occurred after the burying of the wallet in the woods. There is also the logical component, as in this case it is correct to say that “If the guard hadn’t buried the wallet in the woods, it wouldn’t have been stolen.”89 But the physical relationship is not there, as it is not right to say that “The wallet was stolen because of criminality.” In the legal sense there is no causation here, as we saw above that burying the wallet in the woods is considered proper guarding when it comes to theft. What about case a? Here, too, there appears to be a temporal relationship (because it is right to say that “The burning of the wallet occurred after it was buried in the woods”), a logical relationship (because it is right to say that “If the guard had not buried the wallet in the woods then it wouldn’t have been burned”), and a physical relationship (“because it is right to say that “The burning of the coins happened because of the burying in the woods”), with “physical” being replaced here with “legal.” Thus, the Talmudic discussion on the question of “beginning with crime and ending in compulsion” precisely distinguishes between the three elements of causality as we described them here: If there is a full causal relationship then the guard must pay damages. If only the logical and the temporal relationship are in evidence without the physical component, then there is a dispute, and in practice the guard must pay in

this case, as well. If only the temporal relationship is in effect, then we are not dealing with causation at all, and the guard is exempt (aside from Abayeh’s opinion per the Rif).90

Causality in Modern Physics As we noted, ironically, it is physics which warmly embraces the Humean conception of the causal relationship, omitting the physical element therein. As we will see now, equations of physics do not express relations of causation or even causal relationships in general. We will try to demonstrate this by discussing some simple examples within physics.

The Meaning of the Physics Equations Bertrand Russell already pointed to this in brief,91 who demonstrated this on the dynamic of two bodies attracting each other based on the power of gravity. The equations of motion of the two bodies do not give even a hint as to what is the cause or the effect. They simply describe the relationship between forces and their location and motion. But one can see this in a much simpler and surprising example. The second law of Newtonian mechanics determines that there is a regular relationship between force (F) and acceleration (a), with the proportion coefficient being the mass of the body on which the force (m) operates:

F = m × a This means that if we apply a force of 5 newtons on a body whose mass is 1/2 kg, the resulting acceleration will be 10 meters per second squared. Now we ask ourselves what the cause and the effect is: Is the application of force what leads to the acceleration, or is the acceleration the factor which leads to the creation of force? Intuitively, everyone will say that the application of force is the cause, and the result is the acceleration. It sounds strange to say that the acceleration of the ball “causes” it to apply force to it. But if we look at the equation itself, we will see that there is not even a hint of such a relationship. The equation is not directional, and it prefers neither side. It expresses an entirely symmetrical relationship between force and acceleration, without any difference between them which may hint to us which of them is the cause and which the effect. Moreover, the relationship between both parts of the equation is simultaneous, meaning that there isn’t even a temporal relationship between them (they do occur around the same time, but not with one following the other in order). Actually, a look at this equation points only to a logical relationship between acceleration and force (if force then acceleration and if acceleration then force), and perhaps also a temporal component in some sense (while there is a force there is acceleration and vice versa), but there is certainly not even a hint of a physical component of causation. The equation does not tell us what is the cause or the result, or even if either exist. That is something we bring “from home”. Causation is nothing more than our interpretation of the equation, not an actual result thereof. To become further convinced of this surprising argument (that physics equations do not deal with the physical-causal aspect), let us examine another equation. Coulomb’s Law determines the force applied by a local electric charge (q1) on another local electric charge (q2). If the distance between them is r (and K is the Coulomb constant), then the force they apply to one another is:

Here, too, there is no expression of a causal relationship or any indication of what is the cause and what is the effect. What, then, is the significance of this relationship? One can see it as a description of a physical force created by the interaction of these two charges. This force operates on each one of the particles and accelerates it. But one can also interpret the law differently, if we continue the move by introducing Newton’s second law, which as stated determines the relationship between force and acceleration:

Now we can see that the significance of Coulomb’s Law is that when we have two charges at distance r from each other, acceleration will be created among each of them, whose value is given in the above formula. Is there a force which causes this acceleration? We cannot determine that. We see only that if

there are two charged particles at distance r from one another, this leads them to start to move and accelerate. We cannot see the force with our eyes, and there are therefore some who treat it as a fiction, a sort of form of expression as though there is some force which leads to the acceleration of these particles. The physical content of the equation is not the force but only the acceleration which each of the charges causes its neighbor — i.e., only the measurable quantities. By contrast, most if not all physicists assume that this phenomenon is driven by a force which the charges apply to each other, and which leads to acceleration. Why should we assume this? We saw that the equations themselves do not necessarily establish the existence of the force, but only the phenomena it engenders (acceleration). In addition, we have no direct way of approaching, observing, or measuring the force. We can only see it through the acceleration it causes. Still, we assume that such a force exists. Why? It’s pretty clear that this is a causal interpretation for a physical equation. If a force was not in the background, then the acceleration described in the equation would effectively be left without a reason, an event created without any cause. But the common assumption is that there is no event which happens without a cause (the principle of causality), and we therefore prefer to interpret the physical equation as describing a theoretical object, that being a force (or a force field). So, the assumption that this theoretical entity (the force, in our example) really exists, is not the result of the equation but simply the manner in which we interpret it. In effect, we are talking about interpretation based on the a priori assumption of the principle of causality. This is another example of physical equations not necessarily expressing causal relationships in the fullest sense (including causation), meaning that causality is an a priori assumption of the intellect, not an empirical observation.92 This analysis is true for all equations of physics. As noted, the assumption of most physicists is that these equations describe not only the relationships between observed physical quantities, but also theoretical entities which are objects which exist in reality. This assumption is a priori, as it was there even before we observed the force fields themselves. Already Newton, when formulating the law of gravity, referred to it as describing the force of gravity and not the phenomenon of two masses creating acceleration on their neighbor. It turns out that physics is transparent, or indifferent, to relations of causality. It simply doesn’t know how to describe them. A physical equation points to equality between two quantities, meaning a simultaneous logical relationship, and not a relationship of causation. The clear indication of this relationship is that a relationship of causation is directional (i.e., asymmetrical: from side a to b, and not vice versa), while a mathematical equation is symmetrical in essence and therefore cannot describe a directional relationship. Moreover, equations also do not necessarily say that the theoretical entities which appear in them do actually exist. It is our interpretation which assumes the existence of these entities.

The Field Theory and Coulomb’s Law: What is the Cause and What is the Effect? Coulomb’s equation, which determines the relationship between force and charges, determines identity. Are the charges a cause and the force the effect, or the other way around? We cannot derive this from the equation, as both sides of it are entirely equal. We know that when we place a charge a force field will be created, and when we place a force field (even though there is no option to place a force field in a space without charges), charges will not suddenly emerge. We therefor assume that charges are the reason and the force field is the result. But this is a result of our reference point and not of the equation itself. As far as the equation is concerned, if there was an independent way to create and electrical field, charges should be generated by it. Thus, it would ostensibly seem that the development in time is what might represent causal relationships. What happens when we introduce charge points into a space at a particular point in time? How do the force fields develop around them? To understand this, we need to go into a description of field theory. In field theory, the relationship between the two sides of the equation in Coulomb’s Law is not simultaneous or static. When we place charges a force field will be created around them which starts to advance outward from the charges, at a high speed (the speed of light). The moment we laid a charge at any point, another charge still does not feel the force applied to it, as it takes the field time (even if a short time) to cover distances. So, ostensibly the more precise description of field theory does point to the charge being the cause and the field being the effect, as the field was created by introducing the charge and it appears after it in time. But it turns out that this is also not a derivative of the equations themselves. It turns out that field

equations always have another solution (known as a retarded or delayed solution, as opposed to the normal solution which is advanced), according to which the charges swallow the field. This means that in the first moment, the field is in the form of a huge ball with an infinite radius, which begins to shrink. When it reaches the point in the center of the ball, that’s where the charge is suddenly born (the field is swallowed by the charge point). Thus, field theory also fails to introduce causality into the equations of physics. This means that causality is introduced by us into physics on an ad hoc basis, and its origin is a priori. It is not the result of observations or equations in and of themselves.

Kinematics of a Local Particle Intuitively, it’s hard to imagine the opposite relationship of what we described, in which a force field creates the charge. To understand this better, we will demonstrate this effect itself with a simpler case — the movement of a local particle. Let us assume that we have a local particle with a given mass in front of us. Newton’s First Law (the law of inertia) tells us that so long as no force operates on it, it will move at a constant speed in a direct line, without changing. If we describe the path of such a particle on a graph of space against time, we will result in the following diagram:

Graph 1

The angle of this graph is the speed of the particle, meaning the distance it covered divided by the time that passed. If we draw the speed of the particle as a function of time, we will get a straight line with a fixed value:

Graph 2

Is there a temporal direction for this physical process? Certainly not. We can read the first graph in two equal ways: Movement forward in time: Between time t1 and time t2, the particle moves right. Movement backward in time: Between time t2 and t1, the particle moves left. So, we see here an expression of the temporal symmetry of the physical laws of movement. If a particular path matches the laws of physics, the reverse path (when time is pointed in the opposite direction, or the film is on rewind) is a legitimate solution for the laws of movement. All laws of movement in microscopic physics are symmetrical in time, and therefore every path has a reverse path which is a legitimate solution of the equations. Having trouble imagining this? Think of a billiard table. When an individual hits ball A in the direction of ball B, ball A hits ball B and sends it onward to the right. If we play this film backward, what we will get is the following clip: Ball B moves from right to left, hits ball A, stops, and sends A onward to the left. This is

an entirely valid process in terms of the laws of physics. What nevertheless changed here? It turns out that when we see the original clip, we will say that the rightward movement of ball A is the cause and the movement of ball B is the effect. By contrast, when we watch the second clip (the reverse one), we will define the movement of ball B leftward as the cause and the movement of ball A leftward as the effect. Why do we reverse the roles? Obviously because of the temporal relationship. So, physics does not determine what is the cause and what is the effect — that is our interpretation of physical events. Physicists tend to describe the movement of a physical system (in our case the point particle) on a different frame of axes, known as the “phase space.” In this system, the axes are the location and the speed of the particle at a given point in time. This is then drawn for every point in time, resulting in a path in the phase space, in which time is the parameter which accompanies the path. In the case of the aforementioned particle, what we get is the following:

Graph 3

This path reflects the fact that the speed of the particle is identical in all locations.93 How is the symmetry expressed in this graph when it comes to temporal directions? The question is whether time passes rightward or leftward. In other words, the path is the same path, and the parameter t which accompanies it prima facie flows rightward or leftward. Put differently: On a path in which time flows forward, the particle is at a low point and over time advances rightward (its x location increases). On the path in which time moves backward, the particle starts from higher locations and moves left (its x location decreases). The difference between them has to do with the starting point of the system. This is precisely the distinction we saw above in field theory between the advanced and retarded solutions, but it’s easier to see and understand here.

Interim Summary: The Problem of Causation in Kinematics The description so far did not include a causal relationship anywhere. We showed that if a physical path in which event A (a particular speed and location) appears after event B (a different speed and location) is valid (fits the laws of physics), then the reverse path time-wise (in which event B appears after event A) is valid as well. But the relationship between the events is in terms of chronological order, not causal relationships, and we should therefore not be surprised when an asymmetry emerges between the different directions in time. None of the events can be defined as the cause of the other event, as if it is, it cannot appear before it in time. But the laws of physics state that if it appears before it, then it can necessarily also appear after it. So, the temporal symmetry of the laws of physics and the valid paths are two aspects of the same phenomenon: Physics in its mathematical formulation cannot describe causal relationships, meaning relationships of causation. Causality is introduced into the equations by us ad hoc. This is a way in which we look at them, which derives from an a priori point of reference rather than empirical facts, and certainly not the equations themselves. It is therefore also clear that it is wrong to conclude from the symmetry of the laws of physics that the principle of causality contains no physical or temporal component. At most we can say that physics does not address this component.

The Dynamics of a Local Particle So far, we dealt with the kinematics of the particle (movement without force). But what happens when a force operates on the particle? Here we enter into what physics calls the dynamics of the particle. In such a case, we will ostensibly see what should be temporal direction, as in this picture we already have a causal direction. The force is what influences the speed of the particle, meaning that the force is the cause and the change in speed or acceleration is the effect. If so, the results of the force (the acceleration) should appear at a time later than when the force acted. We would expect the body to change its speed only in times after

the force acted upon it, and not before it. What happens if we rewind this clip? Will the acceleration become the cause and the force the effect? For simplicity’s sake, we will assume that the force operates on the particle for one moment in time, t1. Our body under observation moves at a speed of 0.5 meters per second, and the force adds another 0.5 meters per second to the body, leading to its speed after the force’s action being: V = 1 meter per second. In such a case, the graph describing the speed will look as follows:

Graph 4

The location graph in such a case would be (the change in speed being expressed in the change in angle):

Graph 5

The path in phase space to this situation is:

Graph 6

Let’s try and see if there is direction here along the time axis. How do we do that? We need to compare the dynamic graph (with the action of the force; graph 6) to the kinematic graph (without the action of the force; graph 3). The difference between the two graphs appears only at times which are later than the time of the action of the force (i.e., only after t1). Before the action of the force, the two graphs are identical. Thus, in dynamic contexts, it appears that physics does indeed show us that the force is the cause and the change in speed or acceleration is the result. But that’s not true. Surprisingly, we can also see graphs 4-6 in another way, which describe a path which expresses influence backward in time. In other words, dynamics (the response to force) is also symmetrical in time. How can that be? When we looked above at graph 4, we assumed that we would have a particle at location x=0 at time t0, with its initial speed being V=0.5. It moves at this speed the whole time, and suddenly a force acts upon it, and this changes its speed, and therefore also the rate of its change of location, expressed in the angle. We interpreted this as influence forward in time. However, we can see the same graph in the opposite manner: The particle was at time t3 in location x=2.5, and its speed was V=1. We flow with the time backward, and suddenly a force emerges which pushes the particle to the right, and therefore the particle slows its initial

speed, and starts to move at the speed of V=0.5, until by time t0 it reaches location x=0, and its speed in the entire second section is V=0.5. We thus had a particle at a given speed, we flow backward with it in time, and when we reach an earlier time, a force acts, and the influence appears only at times which previously were before the force acted (which we reach afterwards with the flow backward in time). The conclusion is that in contrast to what we thought, the dynamics are also symmetrical in time. Just like the path for the kinematics were symmetrical, and could be interpreted for both directions of time flow, dynamic paths also have the same symmetry. The conclusion so far is that even the influence of force (which is ostensibly a causal influence) can occur backward in time. Not just a series of successive events (like in the kinematic case), but also the force which causally engenders changes in speed and the location of the particle (the dynamic case) can do so backwards in time. Once again we encounter the symmetry of the two solutions: retarded and advanced. This is the picture described by Bertrand Russell in his article on causality. He argues that the symmetry of the physical laws of movement in relation to the directions of the flow of time do not necessarily lead to symmetry in terms of the causal direction of influence. He derived from this that causality is a fiction, as it has no relation to physics, while we lack any other source for such asymmetry meanwhile. It was born within our own minds, just as David Hume argued. But here, the argument is even more far-reaching: Not only is there causality without physical causation here — in physics, there are not causal relationships at all, not even in the sense of the temporal component thereof.

Note: Does Physics Really Not Describe Causality? So far, we saw that empiricism of the sort advocated by Hume leads to the conclusion that the causal relationship only includes a logical and temporal component. We also saw that modern physics, as a clearly empirical field, adopted the Humean definition. The symmetry of the physical equations in relation to the timeline and their structure as equations between two sides dictate the result that the causal relationship will not appear therein. We therefore see that the causal relationship is not the result of scientific study but a given which we introduce into science with the interpretation we give the equations. It would seem to me that it is certainly correct to see the force as the cause and acceleration as the effect, but this is an interpretation of equations and not something which lies in the equations themselves. It’s interesting to note that what Russell did to the laws of science was done two hundred years earlier by Hume to scientific observations. He showed us that the determination that the force is the cause and acceleration is the effect is not a result of the observation, but rather our interpretation of the observation (and as an empiricist, he rejected this). Why do the laws of physics not actually describe causal relationships? It turns out that this derives from the fact that the language our physics uses is mathematics. The problem is that mathematics lacks the tools to capture the causal relationship. It can describe a chronological order, but there is no mathematical tool which expresses a relationship of causation between event A and event B; we already saw that an equation is always symmetrical. If we found a form of formalization which would clearly represent the causal relationship, it may have been possible to introduce causation into the laws of physics and not leave it to interpretation external to physics. It is therefore incorrect to derive the idea that there are no causal relationships between events from the mathematical presentation of the laws of physics. The intuition of everyone (including Russell and Hume themselves) screams in the clearest fashion that the causal relationship also includes a physical component. A cause is not just a condition for the effect, and it is not just prior to it in time. It also engenders or causes it. The mathematical form of the equations which describe the laws of physics are simply not a tool to express this, and therefore this would ostensibly seem to be a deficiency in our physics. That is left to our interpretation of the equations, but does not appear in the equations themselves.94 How does this sit with the empiricist challenge? After all, Hume was right when he stated that we cannot learn of causality from experience. This has a simple, more intuitive solution, which is that the principle of causality cannot indeed be learned from experience, but this does not mean it is incorrect. In some form, we understand that things do not happen without a reason, but this is not the result of observation but rather a principle which is actually prior to any observation. Our conclusion regarding a relationship of causation is forced on the observation by our intuitions. I argue that Hume and Russell were captives of a conceptual system, and they erred in making conclusions about the world based on that system. The fact that the laws of physics cannot express a complete causal relationship (which includes causation), or the fact that sensory observation cannot create a direct connection with the causation component of the causal relationship, is a limitation of the language

and the conceptual system which we use (as well as by Hume with his empiricist assumptions). But it would be a mistake to therefore conclude that this component does not exist in the world itself. In many cases, our conceptual system prevents us from seeing something directly, but that does not mean it is correct to conclude that it does not exist.

Another Note: The Mirror Paradox To demonstrate the argument that we are sometimes captives of a conceptual system which prevents us from seeing objective reality, I will briefly discuss a paradox published in the Israeli newspaper Haaretz ten years ago, which bothered many people, primarily mathematicians and physicists, who are used to considerations of symmetry. Let us imagine a person wearing a watch on his right hand and who stands in front of a large squarer mirror. The picture he sees before him is of a visually identical individual, but the latter wears the watch on his left hand. The shirt pocket of the figure in the mirror is on the right side, instead of the left side. This is the normal symmetry of a mirror, which turns the right side into the left side and vice versa. On the other hand, the reflected figure does not stand with his head instead of his legs on the ground. This means that the mirror does reverse right and left, but it does not reverse up and down. Prima facie this is a very strange phenomenon. After all, the mirror itself is symmetrical. There is no difference between the horizontal axis and the vertical axis (for convenience’s sake, this is a square or even circular mirror). The question must be asked, then: Why does the mirror always turn right to left while retaining the up-down axis where it is? How and why is the “symmetry” of the mirror broken? The obvious explanation of this phenomenon is that this is a consequence of the structure of the human body. The human body is symmetrical from right to left, but it is not symmetrical top to bottom (at least among most human beings, whose head is fundamentally different than their legs). This means that right does not actually get turned left. It just appears that way to us, since these are two equal parts of our body. The conclusion is that the mirror doesn’t actually flip anything, and it is just an illusion of ours based on the structure of the human body. The problem is not the mirror, which is entirely symmetrical, and which doesn’t distinguish between the vertical and horizontal axes, but rather the human being reflected in it, whose symmetry is only between right and left and not up and down. To see why this is not a satisfactory answer, we will continue our thought experiment, and place another object in front of the mirror — an entirely symmetrical ball. We will mark a point on its top and another point, in a different color, on its right. When we look at the mirror we will see an identical ball, but in terms of the vertical axis, it has a point on its top (not its bottom), while on its horizontal axis, the colored point will be on its left side (and not its right). This means that once again the mirror has reversed the lateral parts but not the vertical ones. Here we are already talking about a ball, which does not suffer from partial symmetry, as we argued above regarding the human body. If the previous argument was correct, the mirror should now be able to switch the top and bottom parts of the ball, as the ball also has symmetry on this axis as well. So, the problem does not lie in the mirror nor does it lie in the object before it and which it reflects. So what are we left with? Is there another component in this situation which is the cause of the paradox? The answer appears to lie in the difference between the terms right-left and the pair up-down. Man has built a world of terms for himself which includes a number of spatial relationships: Up and down, right and left, forward and backward, north and south. We have an illusion that each of these pairs have meanings parallel to one another, but this is a mistake. “Up” means a direction distant from planet earth and “down” means the direction closer to planet earth. This means that this pair has definitions in geographicalphysical terms, like the definition of the four directions of the world (north, south, east, and west). All of these are objective definitions, which are not dependent on the individual to define them. By contrast, “right” and “left” have no such definitions.95 It’s very hard to explain to someone unfamiliar with these terms what right is and what left is, without using the terms right and left themselves. There is also no way to do so by pointing to something, as these terms have no root in physical space. The only way I can think of is pointing to one of the hands, right of left, and tying a ribbon on the right hand, for instance. As every child knows, there is no way to remember these definitions until you have absorbed them, and even this is only possible after these concepts are embedded in the child (seemingly due to the symmetry of his body’s physical structure). A creature which lacks these traits will seemingly not be able to understand these concepts. Except that after defining the subjective terms of right and left, the individual usually does not distinguish between them and the directions of the world, or between them and the pair up-down, which have an objective meaning. We do not notice the essential difference between up-down and right-left,

meaning that we are unaware of the nuances which exist in the world of concepts we use, and that is what creates the confusion in the wake of the mirror paradox. The problem is not the mirror, and not the figure it reflects. It turns out that there is another player involved: us. The individual looking in the mirror and thinking about the problem. The problem, then, would appear to lie in this factor. The mirror flips nothing in the physical world; it seems that it is only the relationships defined by us — right and left — which are flipped. The mirror itself is entirely symmetrical, and it also preserves its symmetry. Only the picture which emerges in our minds is not thus, and this is because it is based on this unconscious conceptual system and principles which are implicitly embedded in us, and which we sometimes don’t even notice. Because of this unconscious conceptual confusion, we end up with paradoxes like the one involving the mirror. It is true that this phenomenon occurs because of the symmetry of human beings, but we are not talking about an individual looking at the mirror and being reflected in it, but an individual who defines terms and uses them, who looks at the scene and thinks about this problem. Therefore, the problem exists even when the mirror reflects an entirely round and symmetrical object. Does this mean that the relationships of right and left characterize us and not the world, meaning that they would not exist if we didn’t? Not necessarily. It is more correct to say that if people did not possess a symmetry of right and left, they would seemingly not successfully understand, let alone define such a pair of terms, and they would be left only with the geographical ones. Our structure thus helps us recognize an aspect of reality, and it may well be that this aspect exists not in us but in reality. One of the indications of this is that the distinction between left and right has consequences for physics; there are right and left phenomena which are distinguished from one another, and this points to our pair of terms belonging to reality itself and not just how we look at it. Think of a creature who has no symmetry between right and left. There is some chance that he will not even be able to conceive the relations we call right and left. Does this mean that these actions do not exist in the world? Of course not. If he had an intuition which grasped the right-left relation in some way, even if he couldn’t point to it concretely, he could certainly adopt those definitions even though he lacks the tools to directly conceive and define it.

Back to the Causation Component of the Causal Relationship How does all this tie into what we’re discussing Above? I proposed that the mathematical way we use to present the world in physics (e.g., equations), forces us to ignore the causation component in the relations described in the laws of physics. The causal relationship between force and acceleration, for instance, is not represented in the equation of Newton’s Second Law; it is rather a result of our interpretation of that equation. We learn from the mirror paradox that the manner of presentation and the conceptual system we use can confuse and deceive us, and we must be very careful not to be their captive and end up within paradoxes. In exactly the same way, the fact that physics does not allow us to present the causal relationship may deceive us into thinking that such a relationship does not exist at all. The very fact that physics cannot express a causal relationship can be a problem of our language and conceptual system, but it is certainly insufficient to force us to adopt such a conclusion against our clear intuition.96

A Surprising Conclusion: Does Science Provide Explanations? Before we finish here, let us combine the two parts of the chapter, and see how they lead to a surprising conclusion. Scientific explanations are usually based on providing a cause for an occurrence. When we want to explain some event, we find the causes for it. For instance, we explain the tides with the power of gravity exerted by the moon on the sea. The power of gravity is the cause for the tide ebbing and flowing, and therefore we have a valid scientific explanation. Which of the three components of the causal relationship provides an explanatory power to the cause? Let’s explain this in terms of Raymond Smullyan’s proposal to explain astrology by way of the mechanism of synchronicity between the behavior of the stars and the appropriate events on earth. In the framework of this picture, if we try and “explain” events on earth with the movement of the stars, this will not count as an explanation. The movement of the stars is logically related to what happens on earth, and perhaps even allows us to predict what will happen there, but it cannot be considered an explanation of it. Only if there is a physical component in the relationship between the stars and planet earth, meaning that the movement of the stars cause what happens here (like astrologists claim), then we might be able to see the movement

of the stars as an explanation for what happens down here. So, the explanatory value of causality rests precisely on the physical component of the causal relationship. Without it, the cause would not have any explanatory value in the logical sense. If we add the fact that we learned from David Hume that this component of the causal relationship is not the product of observation, meaning that it is not scientific but rather a priori, we will reach a fascinating and surprising conclusion: The explanatory power of science lies not in science but in philosophy. Science can at most point to logical connections between facts, but not connections of causation between them. Science does not itself provide explanations, and its explanatory power lies in the a priori philosophical assumptions we bring with us, not the scientific findings themselves. In the second part of the chapter, we saw that the laws of science do not deal with causal relations but rather at most with correlation (meaning the logical component of the causal relationship). The causal significance, and therefore also the explanatory significance, we derive from them is at most interpretation we provide for them. I will note as an aside that we can indeed see a scientific theory (as opposed to scientific facts) as a sort of explanation for the facts, but David Hume taught us that even scientific theories are not the result of observation, but rather represent our choice of one of a number of possible generalizations on the basis of observed data.

Summary In favor of those who skipped over the formal analysis, I will present our main findings here. We described three components of the causal relationship: (1) temporal — the cause is prior in time to the effect; (2) logical — the cause is the (necessary, and sometimes also sufficient) condition for the effect; (3) physical — the cause engenders (or causes) the effect. We saw that giving up any one of these components may lead to errors and conceptual confusion. For our purposes, we need only understand that the logical relationship between cause and effect is at the very least sufficient. If the cause is there then so must be the effect, and the effect is singular, the result determined in a single-value manner by the cause. David Hume adopted a different approach to the causal relationship. In his opinion, causality doesn’t include the physical component. Surprisingly, as we saw modern physics has adopted the Humean approach. The equations of physics do not express the physical component (!) of the causal relationship. This means that we cannot learn cause or effect from studying physics equations. The equation will only point to a logical-temporal relationship between the two events or the quantities it describes, and no more. At the end of the chapter, we expanded further on this conclusion, showing that explanations are not at all within the purview of science, always being the result of interpretation we grant to the findings and laws of science. Bottom line, it turns out that the common intuition which produces the principle of causality is indeed not the product of observation (as Hume showed), and it also does not emerge from the equations of physics (as Bertrand Russell showed). On the other hand, there is no necessity to give it up like Hume and Russell did. They allowed a too narrow conceptual system to dictate their picture of reality. A more reasonable alternative is that this is an a priori insight of ours, through which we think about the world. Our intuitive feeling leads us to the relationship between cause and effect having to contain a physical component and not just logical and temporal ones. I will only say in brief that this is a solution similar to that proposed by Kant for the problems raised by Hume. Kant argued that concepts like causality derive from the assumptions and structures of reason and not observation in their simple sense. In the next chapter, we will reach the core of the discussion: The relationship between the principle of causality and the question of determinism and free will. 72 A further expansion on this subject can be seen in Temporal Logic in the Talmud, mainly in the second chapter. A survey of the components of causality, their various representatives, and their conceptual implementations can be seen in Stanford University’s Encyclopedia of Philosophy: Dowe, Phil, “Causal Processes,” The Stanford Encyclopedia of Philosophy (Fall 2008 Edition), Edward N. Zalta (ed.). http://plato.stanford.edu/archives/fall2008/entries/causation-process 73 In the space-time terms of the theory of relativity, it is more correct to say: B is within A’s cone of light, meaning it can be influenced thereby. 74 In quantum mechanics, there is a connection between the subject of the causality and locality of processes. For further study of the question of causality and time in quantum theory, see also: L.S. Schulman, Time’s Arrows and Quantum Measurement, Cambridge University Press, 1997. 75 See, for instance: William R. Wharton, Understanding Time and Causality is the Key to Understanding Quantum Mechanics. http://xxx.lanl.gov/abs/quantph/0310131 76 We are dealing here with the world of physical facts. By contrast, in normative contexts we find quite a number of mechanisms which appear to experience reverse causality from the future to the past. This is true when it comes to legal, halachic, and perhaps even moral systems. A clear example of this is the legal mechanism of a retroactive condition. The question is whether such causality is possible outside the laws of

physics and what does this mean. At heart, the question is what logic runs these mechanisms, which is what our books on Talmudic logic dealt with. We showed there that halacha recognizes reverse causality, meaning influence backwards in time. 77 See more on this in Steinitz’s book Etz Hada’at [Hebrew. The Tree of Life], Dvir 1994, in the second part and related bibliography. Steinitz’s argument in favor of the position of non-necessity will be discussed later in the chapter. 78 To the best of my knowledge, there is no philosophical approach which suffices with a necessary but not sufficient relationship between cause and effect. At least not when speaking of a complete, rather than partial cause. 79 Physicist Wolfgang Pauli contributed a great deal to developing the concept of synchronicity within the system of archetypes of Jung and corresponded with him on the subject. See: Atom and Archetype, The Pauli/Jung Letters, 1932-1958 [Pauli, Wolfgang, Jung, C.G., Ed.], C.A. Meier. Princeton, New Jersey: Princeton University Press, 2001. 80 This difficulty is only an apparent one, as the match is between the state of the stars in the past and the state of human beings in the present, and such an effect is possible according to the laws of physics. I also present this difficulty and its proposed explanation only as a demonstration of non-causal correlation. 81 I will not again that my intention here is not to defend astrology, but only to point to the theoretical possibility that its arguments should not be rejected due to the inability to point to a mechanism of direct influence between the stars and ourselves. 82 See in this matter Leibniz’s clock parable, in his book, The Monadology (also at the net translated by Robert Latta). There, he proposes three different mechanisms for matching two synchronized clocks: (1) clock A influences clock B; (2) Clock B influences clock A; (3) there is a third factor influencing them both. 83 It’s important to stress that the expression “attraction” is reminiscent of a physical force, but this is just an analogy. It is a description which is not causal at all. In fact, it is an alternative mechanism to causality. The fact that we are captive to the thinking of causality is what “attracted” Kammerer, Jung, and Pauli to describe this phenomenon in causal terms. Koestler in his book describes the fierce criticism of philosopher and mathematician Alfred North Whitehead on this inconsistency. Whitehead’s approach on the matter is presented in his book: Idem, Process and Reality: An Essay on Cosmology, New York: Free Press, 1978. 84 See more on this in chapter sixteen, when we discuss the identity of Leibniz’s monads. 85 We can see this in the following way: Assume A is the necessary and also sufficient condition of B. Based on the assumption that A is a necessary condition for B (meaning that there is no B without it), we see that given B, then obviously A. Meaning that B is a sufficient condition for A. The same is true in the reverse: the assumption that A is a sufficient condition for B leads to the conclusion that B is a necessary condition for A. We thus conclude that B is a necessary and sufficient condition for A. 86 The proof of this is very simple: Let us assume that A and B are each a necessary and sufficient condition for C. This means that if A occurs and B does not then C must occur (as A is a necessary and sufficient condition for it). On the other hand, if C occurred, then necessarily also B does, as B is a necessary condition for C (without B there would be no C). This of course contradicts the assumption that A without B is a necessary and sufficient condition for C. 87 See note 8 above. 88 See more on this in the Tosefot, starting verse “Hochma,” Baba Metzi’a 78a. 89 True, in the simple physical sense this is a necessary but not sufficient condition, and as we saw such a logical relationship is certainly not enough to create a causal one. But even when the coins in the woods were burned there isn’t sufficiency. It is therefore clear that at the level of legal responsibility, this sufficiency is taken as given, and what is left is only the discussion of necessity. The very fact that I received an object to guard imposes responsibility on me to ensure that my actions are sufficient for the result, and the only question is whether they are necessary. Therefore, the discussion here has the logical relationship expressed only in terms of necessity and not sufficiency. My thanks to Yehudit Ronen who drew my attention to this point. 90 Let us quickly complete the picture for the curious reader. Why does the Rif think that one can obligate the guard to pay even when there is only a temporal relationship, not a causal one? It seems that in his opinion the very criminality of the guard is what obligates him to pay, as though he himself stole or made the animal disappear. But if he committed a crime and nothing happened to the animal, he “pays” with the animal itself, and therefore de facto does not need to pay. But if the animal died, and even died naturally (in a matter unrelated to his crime), he is still obligated to pay. He cannot return the animal itself, and therefore must pay damages in money instead. As noted, this position was not accepted in practice. 91 Idem, “On the Notion of Cause,” Proceedings of the Aristotelian Society, New Series, v. 13, 1912-1913, pp. 1-26, especially pp. 13-14. 92 The analysis here is a little simplistic and inaccurate, as in principle physics does have ways to view the fields themselves. For instance, the light we see is nothing more than an electromagnetic light (which we also view via the movement of the electrons it generates in our system of vision). The gravitational force field is treated by us as an existing object, and there are therefore theories which predict the existence of gravitons, or gravitational waves, which are expressions of the force field and not its visible phenomena (meaning, the movement of the particles with mass created by it). It is true that meantime we have not actually observed gravitons (apparently due to the insufficiently sensitive equipment we have today), but the fact that such substantial efforts are made to find them and to formulate a theory which describes them means that physicists tend to accept their existence on an a priori basis. 93 As noted, time is the parameter which accompanies the path. For instance, the point above x=1 belongs to time t2 (as we can see from the graphs above). 94 In my article “Zeno’s Arrow and Modern Physics (Hebrew),” Iyun 46, 1998, p. 425, I pointed to a similar fallacy. There I showed that the mathematical formalism of our physics does not properly describe the concepts of motion. We are therefore used to seeing dynamic processes in a distorted way — movement appears to us like standing every moment in a different spot, and the same is true for any situation in time. I showed there that taking note of this point can easily solve the paradox of the arrow of Zeno of Elea. Here my argument is that this attention could contribute to a greater understanding of the concepts of causality and their connection to physics. 95 Here’s an interesting exercise for the reader: Is the forward-backward relationship subjective or objective? It’s worth thinking about the question of what a mirror does to this conceptual pair. Is it like the switching of right and left or the objective pairs which are preserved in the mirror’s reflection? 96 Of course, even our a priori assumption that there is a causal relationship could be a deception. I wish to leave that question open and neutralize the considerations due to the form of presentation of physics. I cannot prove that objective reality contains causal relationships, and I certainly cannot prove the principle of causality.

Chapter Six ON REASONS, CAUSES AND FREEDOM OF WILL All education is a prison. The educated locked within, is harmed by its bars, but the result is true freedom. – Goethe In the first four chapters, we saw that the simple feeling of every individual is that all his actions are free, meaning that our initial intuition as human beings is libertarian. We also saw a number of a priori considerations (and a thought experiment) which confirm libertarianism (or which at least place determinism in a tough spot, forcing it to recognize the possibility of causality which moves backwards in time). In addition, we saw that determinism leads to quite a few problems for our basic moral intuitions, effectively preventing us from judging, let alone condemning wicked people or praising the righteous. All this would ostensibly lead to the conclusion that one should not hold a determinist position. So why is it that so many intelligent people arrive at this problematic and non-intuitive position? Until recently, so long as the problem arose solely in philosophical contexts, the response was simple: The principle of causality. This is the most important aspect of the deterministic position. Causality is the most important, strongest, and perhaps sole justification for determinism. It is no wonder that almost all philosophical discussion regarding determinism revolves around the issue of causality. In recent years, the situation has changed somewhat when the question also arises in scientific contexts, something we will discuss in the final chapters of the book. This chapter will primarily focus on a philosophical discussion of the principle of causality. We prepared the ground for this discussion in the previous chapter, describing the principle of causality, its origins and logical and physical significance. In this chapter, we will discuss the consequences of the principle of causality for the question of determinism.

The Contradiction Between Causality and Libertarianism: A Prelude The principle of causality states that every event which occurs in the world has a cause. If something happens, something must have brought it about. If a ball starts to roll — someone or something must have applied force to get it rolling. A voice can be heard — obviously someone or something created it. A war broke out, a political decision was made, it started to rain — we treat all these events as the results of prior causes. True, as we said in the previous chapter, this is an a priori assumption on our part, and not the necessary result of empirical observation or the laws of physics, but obviously any rational person operates based on this principle. In the previous chapter, we saw that the cause is sufficient reason for the effect, meaning that if the cause exists, the effect must appear. Moreover, the effect is singular (there are no two continuations of the same cause, although there are philosophers who believe that there could be two different causes which could separately lead to the same result). In the previous chapter, we called this the “sufficiency” of the cause and the “uniqueness” of the effect, two characteristics of cause and effect which are agreed to by all. On the other hand, we saw that libertarianism assumes that there are events which occur without a cause. When an individual chooses to freely think or do something, the libertarian argues that this is not due to any cause, as if it was we would not be talking about free choice. To speak of the physical causes of the decision is to say that the physical causes caused this “decision,” not individual human will. In the libertarian picture, free choice is an event which appears ex nihilo, without a prior cause. The individual acts because he decided, he has arguments of what he wants to achieve and where he wants to strive, and there are of course various influences on the decision, but the action does not have causes which bring it about (i.e., determining it in an unequivocal manner). Put differently: There could be situations where the cause of a given result is present, but the result will not emerge (because the individual chose not to do so). This, of course, contradicts the principle of sufficiency of the cause for the effect. The conclusion is that libertarianism is directly opposed to the principle of causality, and this in itself is the most powerful consideration in favor of determinism. In light of what we saw in the previous chapter, we can now understand that this consideration is also an a priori consideration on our part, which appears to be just as strong as the libertarian intuition: The principle of causality, according to which things do not happen without a cause which brings them about. It’s important to stress that in light of what we saw in the previous chapter regarding Hume’s principle, causality is also not the result of empirical observation; it is not a scientific fact, but an assumption of our reason or intuition.97 Thus, the clash between causality and freedom of will is a clash between two intuitions, not a clash between an a priori assumption and an empirical observation as it is sometimes presented. Still, we mustn’t dismiss either of these two intuitions of causality and freedom of will, and we must therefore seek a solution which leaves both as intact as possible.

Determinism and Causality We already demonstrated that determinism is not bound to the concept of causality and does not require it. If we assume that a state of affairs at time t unequivocally determines the state of affairs in the time which follows, then we hold a deterministic position. There is no need to assume that the state of affairs at time t causally determines the state of affairs which follows. Per the terminology explained in the previous chapter, it is sufficient for us to have the logical and temporal elements of the causal relation to establish determinism. The argument “If A, then necessarily B,” even without saying that A causes B, expresses determinism (as it establishes that the present circumstances determine what will happen next).98 But the approach according to which the situation at time t unequivocally determines the situation following it is not reasonable unless we assume causality. If the first situation does not causally determine the following situation, then why should we assume that the following situation has indeed been determined? Therefore, even though determinism is not necessarily tied to causality at the logical level, the justification of a deterministic position is very much tied to causality in its fullest sense. Causality is the basis for holding a deterministic position, even if not determinism itself. We will demonstrate the problem libertarianism suffers in light of the principle of causality with a description (a schematic one, but sufficient for our purposes) of a human decision which ends in a mechanical action. Not every human decision is like this; we can also make decisions which do not involve the mechanical movement of any limb. By contrast, when an individual decides to carry out a physical action, there is a chain which leads from the decision to the final resulting action. We will now examine how that takes place in practice. When a man decides to walk, his foot starts to move. Walking is a physical-mechanical event, and this means that we have movement of objects with mass (the feet, and the rest of the body with it). According to the laws of physics, objects with mass do not move without cause. There must be a force operating on them. In other words, when a stationary object starts to move, then at least in the beginning it accelerates (increases in speed). According to Newton’s Second Law of Motion, which we already encountered, for a body with mass to accelerate — force must be applied to it. When you go one stage backward, there is a force which leads to the movement of the hand or the foot — the force which the muscle operates

on the moving limb. The muscle applies force on the hand and thus it moves, thus carrying out Newton’s Second Law. But now we must go another step back: What moves the muscle? After all, it moves, contracts and expands, and this must also have some physical cause. The physiology of the muscular system teaches us that this movement is due to an electrical flow which reaches it via the nervous system. Thus, this is once again a proper physiological process: The movement of the muscle is due to a physical cause which engenders it, the muscle converts the electrical current into mechanical movement, which is definitely possible according to the laws of physics. But the electrical current which moves the muscle is also a physical event, and therefore is also clearly the result of some cause. There must be a physical cause for this current. An electrical current is the movement of charged particles (electrons), and what drives it is usually an electric field (though not necessarily). We know today that the mechanism which sends the electrical current to the muscles occurs in the brain. The brain forms a physical-chemical reaction which ultimately creates the electrical current which reaches the muscle via the nervous system, moves it, and the muscle moves the hand or the foot, and so we move. So far, so good as far as the laws of physics are concerned. So, now we are at the point in our chain where we must ask what creates the neural activity (the electrical field) in the brain, the one which started the whole chain we described up until now. Effectively, we are asking what generates the electromagnetic field which accelerates the electrons in the brain, which are sent to the muscle and move it, and which move our mechanical limbs (and so on and so forth)? If we go all the way back within the human brain (which is of course far more detailed and complex than what I described here), we must at some point reach an event at the beginning of the chain, which is the initiator of the whole series of events. What’s going on there? Did something happen there ex nihilo? We will now examine the two answers proposed to this dilemma, and which stand at the foundation of the debate between libertarianism and determinism.

The Deterministic Approach The determinist argues that the initial cause is an event which occurred outside of the individual. There is some event which starts the internal process we described, and which ends with a mechanical event in the form of the movement of a limb. For instance, someone threatens me, and I identify the anger via rays of light which hit the cornea of my eye. My brain processes this visual information (we are speaking of the physicalchemical reaction; we will deal with the mental question below), creating an electromagnetic field in response which generates an electrical current, which ultimately reaches my arm muscles, which raises itself to strike my assailant or defend myself from him. This is the deterministic description of the human “decision.” Determinism paints a picture of the human being as an automated machine driven by external factors. This entire internal chain of events which I described is merely part of a much longer chain which reaches back to the Big Bang, reached the individual who threatened me (who of course also underwent processes similar to what I just described), continued with the brain activities I presented, and ended (for now) with my arm moving to strike my assailant or defend myself from him. These actions will of course continue onward in further causal changes, by myself and others, in an unending series until the end of days. It’s interesting to ask here: What will determine whether I defend myself, strike first, flee, or restrain myself and stand my ground? The external stimulus is the same, yet people will react differently to it. The determinist will argue that the difference is the physical structure of the brain, which is a result of both his prior experiences and his genetic makeup. The response formed is a combination of the external factor (the threat) and the structure of the machine generating the response (the brain). Together, these two factors unequivocally determine my response. Only different machines can create different outputs in response to the same input. Despite the difficulty in seeing the human individual as a type of machine, this picture is ostensibly necessary. It is a conclusion which is forced by the laws of physics and the principle of causality. How can the libertarian hold to a different picture without rejecting causality itself? Does he believe that it is possible that there be a physical action without a physical cause? More concretely: Does he believe an electron in the brain can move without an electromagnetic field moving it? After all, an electron in the brain is just like any other electron in the universe; it’s the same matter, with the same traits, and is therefore subject to the same laws of physics.

The Libertarian Approach The above description seems to leave the libertarian with no way out; he must accept the possibility that the electromagnetic field was created ex nihilo, without a physical cause. Human will or decision making starts this chain, generating a new physical-chemical chain of events like the one we described above. The libertarian effectively argues that there are electrons in the brain which move without an electromagnetic field making them do so. As strange as it sounds, they start moving just because. Human will can make electrons move without applying force. Magic. It seems more likely that human will moves electrons by creating a force field. This means that there is a force which operates on the electrons and makes them move, but this field is created ex nihilo (in the physical sense), simply by virtue of the fact that I willed or decided it. An electric force field is usually created by electric charges. A gravitational force field is created from gravitational charges, these being bodies with mass. Meanwhile, the electric force field which moves electrons at the beginning of the decision process, in contrast to any other force field we are familiar with, is created without any charge distribution creating it. This force field has no physical origins. Here we reach the main weak point of the libertarian approach. It necessarily leads us to the conclusion that force fields can be created without sources (charges). We may have indeed moved the libertarian’s problem from one involving an electron moving without a force applied to it to the creation of a force field without an origin, but this still leaves the libertarian description of events contradicting the laws pf physics. The idea of physical objects beginning to move without any physical cause contradicts both our intuition and our experience. It’s very hard for us to accept such a possibility, and this is a major reason which leads people to accept the deterministic position, despite the existence of such good arguments in favor of libertarianism, and even though all our intuitions point to libertarianism.

An Initial and Unreasonable Solution: Humean Causality To preserve the picture of free will, it may be convenient for the libertarian to reject the physical component of the principle of causality, reject the idea that causality includes direct causation. He could adopt the approach of David Hume, according to which event A does not directly cause event B but only precedes it in time and is a logical condition for it, leaving only two of the three components of causality which we described in the previous chapter in place. According to this conception of causality, no physical event is actually the result of any cause. All we can say is that there are events which precede an event and which are a logical condition for its occurrence. We thus empty the principle of causality from its physical content. The brain event is thus not irregular in any way — it is also not caused by anything. But this solution is unsatisfactory, for at least two reasons. The first is that even if such an approach saves the libertarian, the result is a pyrrhic victory. As we have seen throughout our discussion, while the determinist is willing to treat many of our intuitions as mere illusions, the libertarian holds to a position which adheres to an intuition of ours (i.e., our experience of freedom), and is not willing to view our basic intuitions as illusions. Thus, the libertarian will reject this convenient approach for the same reason he rejects determinism: It is contrary to our intuition. It is not reasonable to defend one intuition by forfeiting another one which is no less fundamental. The second reason why the Humean solution is insufficient is that even if we would be willing to ultimately become Humeans, even Hume argues that every physical event has a logical component which is a sufficient reason for its occurrence. Even if a physical equation does not describe a causal relationship between its two parts (since as we saw, this is our interpretation of the equation, not an inherent conclusion therefrom), it still determines a logical relationship between them: If there is force, there is acceleration, and if there is no force, there is no

acceleration. By contrast, in a brain event of free choice, the chain starts by itself without any event which “engenders” it, not even in the logical-temporal sense (meaning it was prior to it and served as a logical condition for it). Ostensibly, electrons thus start to move without there being a force present. Or there is a force present without charges which create it. Either way, we have a contradiction with the laws of physics, even if we gave up the component of causation in physical causality. Thus the adoption of a Humean approach towards causality does not actually solve our problem. The meaning of all this is as we saw above: Determinism is not dependent on there being a causal component within causality. Even if the state of affairs at time point t is not the causal factor for the following state of affairs, but only a logical condition for it, that in itself is sufficient to state that the state of affairs at time t unequivocally determines the state of affairs which follows. The logical and temporal components are sufficient to establish determinism. Therefore, in the discussion below we will assume the intuitive assumption of causality (with all three of its components: temporal, logical, and physical), and nevertheless seek out a way to hang on to a reasonable and intuitive libertarian approach.

Between the Principle of Causality and the Laws of Physics It is important to distinguish between two terms I have used here somewhat loosely. When I say that an electron moves or a field is generated ex nihilo, I mean that it occurs without a physical cause. That does not mean there is no cause at all. There certainly is one: human will, meaning my desire to act. Thus, this is not something from nothing, but something from something; it’s just that the first “something” is not physical. In the view of physics, it is something from nothing, because physics treats non-physical objects and events as “nothing.” We can put it this way: Libertarianism at most contradicts the laws of physics, but not the principle of causality. The physical event does happen due to a cause, just not a physical one. But if we take this one step further, then we ostensibly have something here which also contradicts the principle of causality: the formation of our will or decision itself. This may not be a physical event, but rather a mental one which engenders physical events, but we still need to admit that it in itself happens without a cause; as we saw, any cause be it physical or not contradicts its being free. In the second chapter, we saw that if an individual’s decision is caused by any previous factor, then it is not actually freely created by my judgment. There may not be a contradiction here of the laws of physics, as the event is not physical and not subject to the laws governing physical events, but there is a contradiction of the general principle of causality. In other words, we distinguish here between the principle of causality and the laws of physics. The principle of causality means that there must be a cause, but it does not necessarily need to be a physical cause. It also does not distinguish between physical events or other kinds of events. By contrast, the laws of physics only deal with physical objects, and as far as they are concerned, a specifically physical cause is needed. After we have made this distinction, we can return to the above description, which speaks of a will that creates a physical force field which starts to move particles and create mechanical actions. In sum, libertarianism makes two statements: (A) The creation of will is not a physical event, and therefore does not clash with the laws of physics. On the other hand, it seems to be contrary to the principle of causality; (B) The formation of a physical force field by an act of will is a physical event without a physical cause. This stage does not therefore contradict the principle of causality, but certainly contradicts the laws of physics (as they require a specifically physical cause, and will is not such a cause).

The Principle of Sufficient Reason To reconcile these contradictions, we must introduce a new factor: The principle of sufficient reason, one of the most interesting and confusing philosophical principles formulated by Gottfried Wilhelm Leibniz.99 In his book on monads, Leibniz states that: And that of sufficient reason, by force of which we think that no fact can be real or exist, and no statement can be true, without there being sufficient reason, why it is this way and not otherwise; even though these reasons, we can usually not know them at all. Leibniz is expanding and to some extent weakening the principle of causality. He deals with the question of why something even exists, why does our universe exist? He wishes to prove God’s existence from here, and this is what Kant will call, a hundred years later, “the cosmological proof for the existence of God.” The principle of sufficient reason states that there must be sufficient reason for the existence of the universe, and it of course must be something outside the universe itself, and he thus arrives at the conclusion that there must be a transcendental factor which created the world, meaning God. It’s important to note that Leibniz’s principle does not speak of a cause, but of a sufficient reason for anything, be it an entity or an event. This is an assumption which is not based on observation or experience, but is rather an a priori assumption of rational thought. Just as we saw that our experience is based on the presumption of causality, not vice versa, so, too, the principle of sufficient reason is not necessarily the result of our experience but is rather an a priori assumption. What is the difference between a reason and a cause? As many have argued, there isn’t necessarily a causal explanation as to why something even exists. There are a few reasons for this: (a) The relationship of cause and effect is part of our world; (b) This cause is necessarily outside our perception (as it leads to the existence of everything we know), and therefore the very question of causality becomes unintelligible; (c) If time itself was created simultaneously with the universe, then it is unclear how we can speak of a cause prior to the world (i.e., operating before its formation). The term “reason,” meanwhile, refers to something broader. Even if we accept that there are things or events which are not caused, they need to have a reason. For instance, an eternal object which always existed is not really something which can be discussed in terms of what caused it. As cause precedes effect, and the object in question was always here, there is no real possibility of speaking of the cause which engendered it. But there is still room for searching for the sufficient reason of the existence of the object. Take the four fundamental forces of physics: Strong, weak, electro-magnetic, and gravitation. Some see them as eternal laws which always existed (perhaps even prior to the Big Bang, if that has any real meaning). Is there significance to the question of who created them? If they are eternal — then certainly not. No-one created them because they always existed. Is there significance to the question of their reason? Not necessarily. But when we ask why the laws of physics are as they are and not otherwise, that is certainly a question with meaning. It searches for sufficient reason for these laws, and not necessarily a cause which created or engendered it. Obviously, if the laws were arbitrary, a meaningless combination of principles, the question would then be weaker; the laws are the way they are, just because, and could have been entirely different. But when we are dealing with a unique set of laws like the laws of physics in our world (enabling, for instance, the emergence of chemistry and biology, and therefore also life; also known as the “fine tuning” argument), the question of sufficient reason definitely becomes stronger. Such a unique set of laws makes us feel that there must be some reason why these laws are in force and not others; this is a question of sufficient reason and not of causality. How does one answer questions of sufficient reason? As quoted above, Leibniz noted that we are unable to do so in many cases. But this does not mean that the question is unintelligible or that sufficient reason does not exist. Others argued that the sufficient reason for our universe, or the laws which control it, is that God ensured that the world would run according to those laws. If possible one should also point to his direction, as then the reason will be far more complete and satisfactory. We should note that the statement is not that God created these laws, as they always existed, but that he ensured that these would be the rules and not others. My aim is not to convince anyone of the correctness of this particular answer, but to demonstrate another form of question and answer. This is an example of an explanation based on sufficient reason — a

non-causal explanation (not a description of creation). This is an explanation based on a reason and not a cause. Another such example is quantum theory. There are events in quantum theory which at least according to present-day definitions seem to occur without a cause. The decay of an atom, a move between quantum states, the formation of particle pairs from the vacuum (in a manner in which the charge of one cancels out the charge of another, thus preserving the overall charge in the world). What causes all these events? Many claim there is no cause. It happens spontaneously. Does that mean these events occur without any context? Should such events be expected throughout the entire world, regardless of the laws of nature which are in force? Of course not. After all, in principle we rule out the formation of something ex nihilo. Only the quantum nature of our world (according to quantum theory) is what allows these events. Quantum theory explains these events, but that explanation is not causal. No physical object or event we know of causes or engenders these events, yet we still see quantum theory as an explanation for their occurrence. Another example of a non-causal explanation is a teleological or purpose-driven one. Such an explanation interprets an event based on the purpose it is supposed to serve. For instance: I plant a tree in order to reap the fruits in the future. This is an explanation which refers not to the cause of the planting but the purpose. Although many physicists do not admit this and many might not even be aware of it, it is accepted and sometimes necessary to use teleological explanations, even in physics. I will not delve deeper into this complex issue, as our focus is human decisions according to the libertarian approach. As we saw, a decision of the libertarian type can only have a teleological kind of explanation. For instance, Joe decided to call his friend John to arrange a meeting the next morning. Did the decision to conduct the telephone call occur without a cause? The libertarian will answer in the affirmative, as it was the freely chosen decision of Joe, and free choice is an event without a cause. On the other hand, it is correct to say that this decision lacks sufficient reason? Of course not. Joe had excellent reasons to make the call, as he wished to arrange a meeting with his friend John the next morning. The judgment in the background of the decision is what gives us the sufficient reason for it. Judgment determines goals and values, and decides how to act in order to realize them. Thus, the libertarian may not accept a causal explanation for freely made decisions, but this does not mean he rejects any explanation at all for them. The decision has a sufficient reason, this reason being teleological. This brings us back to the discussion in chapter two, where we saw that free will is different than randomness. A random event occurs ex nihilo, without cause or reason. A deterministic event, by contrast, occurs due to a cause. Meanwhile, an event of free will lacks a cause (ex nihilo), but has a reason. The distinction between reason and cause once again shines a light on the distinction between determinism and choice, and between these two and randomness, demonstrating how they are three different mechanisms. I should note here that I am not dealing with the question of whether Leibniz was right that God’s existence is a natural conclusion of the principle of sufficient reason, and certainly not the question of whether the principle of sufficient reason is everywhere and always correct. My goal is simply to present the principle of sufficient reason and distinguish it from the principle of causality. Why? Because I wish to replace the causal explanation with another explanation, which is perhaps “weaker” but nevertheless provide us sufficient reason for reality. The libertarian thus suffices with the principle of sufficient reason for when it comes to the free desires and decisions of individual human beings, and he does not demand that they have a causal explanation. The sting of the determinist’s attack is thus greatly dulled. It’s true that the emergence of will occurs without a cause, and this may contradict the principle of causality, but it certainly does have sufficient reason. It is incorrect to say that it “just happens.”

The Uniqueness of Human Will So far, we discussed the formation of will or a decision. But what about the formation of the force field which starts the aforementioned physical chain of events? As noted, in the libertarian picture, will creates the force field which starts this chain. We saw that this does not contradict the principle of causality but it does contradict the laws of physics. Does libertarianism have a solution for the problem of the emergence of a force field without a physical factor being the cause? It doesn’t seem so. The libertarian does indeed seem to wish to argue the existence of a unique human ability, which succeeds in creating a force field without origins (charges). It’s important to understand that the very search for a solution to this problem contains a degree of question-begging, as it assumes that there must be a cause for our will and action, which is something libertarianism rejects out of hand. We could perhaps put it differently: Will can transfer energy from the body and create a force field. I don’t mean to say that will is physical, only that it can act and be acted upon in the material world, including taking energy and using it to create something physical. In other words, if we check the energy balance of a human being before and after the decision and the action, we will not necessarily see a violation of the Law of Energy Conservation. To the contrary, it is more reasonable to assume that the energy used by human will to create the force field came from elsewhere in the body, thus preserving the balance of energy. Put this way, will is only unique in that it is a non-material entity which can transfer energy and bring about activity through it. This formulation dulls the sting of the other objection regarding the ostensible violation of the laws of physics (the formation of a force field). It would appear that this formulation also fits our intuition. It is true that particles do not move unless some force acts upon them, but we have no experience which says that force fields cannot be created by a mental capacity such as will. We know that electrical charges create an electrical force field, but there is no reason to assume a priori that human will cannot also do so. The contradiction with the laws of physics thus looks rather weak. Obviously, if we assume that a force field cannot be created without charges, then our conclusion will be deterministic. But this once again begs the question, as this is precisely the point being debated. This is precisely what the libertarian is arguing: Human will is unique and irregular, and unlike any other physical state, can create force fields without charges. Thus, the libertarian sees human will as a unique event, or a very unique fact, in at least two aspects: it can create a force field without a physical cause, and it is created by force of a sufficient reason but not a cause. Here I wish to describe this uniqueness from a slightly different angle. Above, we described a decision to make a phone call to set up a meeting. How does the causal chain appear in this case? The determinist would draw it up like this:

By contrast, the libertarian sees the chain as follows:100

In a way, the first event in the libertarian causal chain is actually the future meeting (or to be more precise, the plan to set up a meeting). The will “sees” the (so far hypothetical) future and decides to strive for it. It creates a force field which engenders the mechanical action, ultimately leading to the plan being carried out and the meeting taking place. Thus, will is also unique in that it is affected by a future event, which creates a causal chain leading to its realization. Will translates a hypothetical future into physical causes which create a causal chain which turns a hypothetical future into a reality. Put differently, purpose becomes the inverse of the cause. The future purpose, the meeting, moves back and becomes the cause of the causal chain of events, which leads us to the purpose itself. The libertarian argues that this ability to plan, reversing the direction of causal influence, is a fundamental and unique characteristic of human will. In other words, the libertarian sees all the other human skills, as well as all the physical entities (forces and events), like the determinist’s linear chain. As far as he is concerned, the entire world is causal, moving in chains of cause and effect from past to future. The only exception (that we know of) is human will, which can create a new chain. But he does not create it ex nihilo, but based on a sufficient reason: The desire to set up a meeting. Why was this desire formed? Because the individual wants to derive some benefit from this meeting (this is even further in the future; it deals with what will happen after the meeting, but meantime this future is also hypothetical). An event lacking a reason, as we saw, is an event which actually does emerge ex nihilo, and it is an expression of randomness. What makes choice unique is judgment and decision making (i.e., action for the sake of a purpose), which provides sufficient reason for the mechanical action even if it is not its cause. Instead of a cause producing an effect, we have a theoretical effect producing a cause which then realizes this effect. This is the libertarian mechanism of free choice, as opposed to determinism and randomness. Here we have put our finger on an important point. In the previous chapter, we noted that when we demonstrate indeterminism, we may have ruled out determinism, but we have not necessarily ended up with randomness. Our argument was that free choice is also a mechanism which is not causally determined by the present. How is it different than randomness, if both kinds of events occur without a cause? The difference is that human choice has sufficient reason, while random events lack either cause or reason. A random event just happens and that’s it, while a human decision occurs because of a reason even if not a cause. In chapter two and four, we called this “judgment.” The teleological explanation may not be a causal one, but it does provide sufficient reason, and the libertarian is therefore prepared to suffice with it. Indeed, as I noted, the libertarian requires the principle of sufficient reason as an alternative to causality (and does not just suffice with it), as without the element of turning purpose into a cause, we would have a random event and not a mechanism of judgment expressing free choice. This transformation created by will, is the essence of human free choice, and it is what separates it from random occurrences. Here we need to remind the reader that already in chapter two, we saw that even libertarians admit not all human actions are like this. The first such example is the “irrepressible” impulse, where the human being is compelled to act by forces he is unable to resist. The other, far more common example is actions done without judgment, which are thus not based on sufficient reason but rather causes, and are therefore not counted as actions based on free will.

A Note On Conscience and Ethical Decisions We already noted that the causal intuition also refers to moral decisions. When an individual decides to do a moral act, we immediately ask why he did so. Our assumption is that there is some cause for his action, and that if there was no motive for his actions, he would not have done them. The proposed explanation usually revolves around conscience. If the individual had not done the moral deed, he would have had pangs of conscience, and he therefore decided to act. Except that if he does so only to escape pangs of conscience, then his act is not truly moral, as his motivation is his own drives and not the needs of his fellow man.101 Moreover, conscience would appear to be something which arises on its own, not by force of any individual decision, and an action based on conscience does not appear to be based on a freely made decision. If conscience is a sort of instinct which naturally arises, and it is the cause of the action in question, then it also has an engendering cause, meaning that the action is not the result of a freely made decision of the individual but that particular cause. But if the moral act is not based on conscience, why was it done? The libertarian argues that there is no cause for such an act. It is the result of judgment, but judgment is not a cause, but rather a sufficient reason in the sense of directing oneself towards a purpose. A moral act is an act an individual does in order to achieve a purpose, and not due to a reason, and such an action is therefore the result of free choice. According to this interpretation, one can perhaps say that conscience is not an instinct which simply arises by itself, but is in fact the result of our judgment and autonomous decision. This brief discussion returns us to the debate between Smullyan’s Taoist ethics and those of Kant which we discussed in chapter three. We saw there that Kant believes that an act is moral only if it was done for a purpose, not if it was done due to a cause. For him, actions which are caused are not autonomous deeds of the individual, and they therefore contain no moral significance. Another possibility is to identify conscience with the framework of the internal psychological conditions the individual operates under. Conscience can only point to the morally correct course of action, but it does not engender the individual’s actions. After the awakening of conscience, the individual must now autonomously decide whether to act or not. This decision is the result of free will, not the automatic result of the awakening conscience. In terms of the hiker parable used in chapter two, conscience is part of the individual’s mental layout, but it is not the cause of his action.

Libertarianism’s Attitude to Causal Intuition So far, we saw how the debate between determinists and libertarians revolves around the contradiction between two intuitions: The intuition of freedom and the intuition of causality. The determinist chooses the latter and finds himself having to reject the former. The libertarian embraces the former and thus finds himself having to reject the latter. So which one is right? How can we decide between the two? Is it a purely subjective matter of taste, or is there an a priori way of demonstrating the superiority of one over the other? To the best of my judgment, we are dealing with a case of superiority in principle, not just a matter of taste. At the fundamental level, we saw at least three powerful arguments in favor of libertarianism: Newcomb’s Paradox, the fool’s cycle of determinism, and the intuition of moral judgment (as well of humanity being more than just a machine or an animal). These in themselves

definitely make libertarianism preferable, and their weight aids us in rejecting the second intuition, despite all the difficulty in doing so. But there is another consideration, this time at a more methodological level, which also leads us to prefer the libertarian option. It is accepted in legal systems that when there is a clash between different principles, various doctrines for deciding between them are implemented. One of the foundations of these is that the specific law must always be preferred over the general law, or in Latin: Lex specialis derogat legi generali, or simply: Lex specialis. An halachic example will get the point across. Halacha prohibits killing another human being, but when someone murders their fellow man, their punishment is execution. Ostensibly, we have a contradiction here between the general halachic prohibition against taking another human life and the particular norm of executing a murderer. Which norm wins out? The Lex specialis determines that the particular overcomes the general, meaning we should kill the murderer. Why? Because if we give up the particular norm, then the law stating that a murderer must be executed is emptied of meaning. In such a case, the law has no legal force and is utterly invalid. By contrast, if we prefer the particular norm over the general norm, the general norm does not subsequently become a legal dead letter: There is still an halachic prohibition on killing, but there is an exception when it comes to murderers. This is therefore the preferred legal interpretation. In other words, when the legislator enacted the particular law, he clearly did not attempt to thus invalidate the general law. We therefore prefer the interpretive option which leaves both laws as having legal validity and force, with the particular law being an exception of an otherwise valid general rule. This seems to be the case when it comes to our dilemma as well. We have a conflict between the intuition of freedom and the intuition of causality. Which is preferable? How can we implement the doctrine of Lex specialis? We see clearly from this doctrine that the libertarian intuition is preferable to the causal one, as even if we adopt libertarianism, we will not have therefore emptied causality and the laws of physics of meaning, but rather made free will an exception and special case. In any other context, they remain as valid as ever. By contrast, preferring the causal intuition would empty our intuition regarding free will of any meaning. We can make the argument as follows: We prefer to assume that the principle of causality and the laws of physics are usually correct, and only that their application to free will, not based on empirical observation but an a priori assumption, is incorrect. After all, no-one has shown that free will works just like any other physical phenomenon. This is therefore a generalization which is not really properly established, and there is no reason in principle not to reject it, or at least restrict it, and determine that while it is true that force fields are usually created by charges, there is a possibility that human will can create a force field by itself. This possibility is certainly preferable to choosing (?!) the deterministic picture, this being the approach that causal generalization must also apply to will (which, as noted, is not based on empirical observation), thus requiring a complete rejection of the intuition of freedom, moral judgment, and the idea that humanity is more than machine or animal. In other words, this methodological consideration also leads us to the conclusion that the libertarian position is preferable to the deterministic one, despite the cost involved in doing so and which we must not deny.

The Principle of Causality: Summary Above, we presented three mechanisms which bring about events in the world: determinism, randomness, and (judgment based) free choice. Already in chapter two, we pointed to the feeling that the deterministic mechanism is familiar and intuitively recognizable to us. We are thoroughly familiar with causes and effects. We saw that the second mechanism which we are ostensibly familiar with is the random mechanism, by which things just happen, without a cause. Prima facie, the mechanism of choice is less familiar to us, and we therefore try to map it out based on one of the other two mechanisms. This is the root of most, if not all misunderstandings regarding freedom of the will. The problems usually start when we try and discover the cause of a particular decision we made, and fail to do so. But as we already noted there, this is already a mistake, as there is no cause (!) for mapping one mechanism based on other mechanisms so different from it. Just as we will not try to base determinism on randomness or free choice, it would also be wrong to base the “mechanism” of free choice on the causal-deterministic mechanism. The reason people tend to do this is due to the feeling that determinism and randomness are familiar and intelligible, while free will is vague, unfamiliar, and unintelligible. If we look again at the three mechanisms, we are in for a surprise. The only mechanism which we are directly familiar with in an unmediated manner is free choice. The other two mechanisms, which are ostensibly more recognizable and intelligible, are speculative hypotheses lacking any real empirical basis, as we established above. None of us has seen something happen without a cause (a random event), and as we learned from Hume, no-one also has seen, with their own eyes, something happen because of a cause (i.e., did not see the causal relation between cause and effect). The conclusion is that both causation and randomness are mechanisms unknown to us from experience. What about free choice? Our most intimate experience would seem to be our freedom to choose our own path. If this is not an unmediated and empirical familiarity with the mechanism of free will, I do not know what empirical recognition looks like. I don’t think this is less clear than something I see with my own eyes. The conclusion is that from among these three mechanisms, it is precisely free choice, which seems so unintelligible and problematic, that is the only one we know in an unmediated manner. The other two, which for some reason seem more familiar and intelligible, are the result of a priori assumptions and generalizations which are not necessarily properly established. They are certainly not based on our empirical experience. If that’s the case, why should we try and base free will on the mechanisms of either causation or randomness? Whence the implicit assumption in many of the discussions on free will that we must always seek out a cause for every event and action, otherwise it must be defined as a random and arbitrary event? Why do the other two mechanisms seem to us entirely natural and obvious, such that we thus try to base choice on one of them? It seems to me that the misunderstandings regarding freedom of will have to do with the fact of physical thinking dominating our logic. Determinism is the result of an unclear preference for unfamiliar mechanisms over the familiar ones, and the denial of the existence of the familiar ones simply because they cannot be based on the unfamiliar ones. Sound strange? It is. It’s very strange. The denial of our most intimate and clearest feelings is a strange phenomenon. When it’s done without any real justification, it becomes so sevenfold. The libertarian is effectively telling us to return to our most basic intuitions. We have freedom of will, and as we saw, the problem of its clash with the principle of causality can be resolved fairly easily. The unmediated feeling of freedom means that our will can indeed engender physical phenomena. This is precisely what our unmediated feeling is telling us. The determinist must present very, very good reasons for us to reject such a clear feeling. If the only reason is a shaky generalization that force fields can only be created by charges and that will is no exception in this, this is certainly not sufficient reason. To the contrary, our most basic intuition says that the principle of causation is incorrect when it comes to my own will, and I see no dilemma here which would force me to reject it. The causal intuition is only relevant for situations and events unrelated to will, and there the laws of physics are indeed in full effect. We have returned to the principle of Lex specialis. 97 We may have seen matters only in relation to cause and effect, but here I will add that the logical relation between them is also based on scientific generalization, and this is but an a priori intuition (the principle of induction, which Hume also undermined in the same manner, demonstrating that it is also an a priori intuition and not observation). 98 This can be seen in the discussion under the entry “causal determinism” in the Stanford Encyclopedia of Philosophy: http://plato.stanford.edu/entries/determinism-causal/, as well as Norton’s article: Norton, J.D., “Causation as Folk Science”, Philosopher’s Imprint, 3 (4), 2003. 99 A highly accomplished 17th century mathematician and philosopher. Discovered infinitesimal math alongside Isaac Newton, a brilliant mathematician and physicist in his own right. 100 This is just a schematic description. There are factors which influence decision making even in the libertarian picture, and they are supposed to appear before the circle on the left. We will develop this important point in chapter fifteen, where we will draw the fuller picture and see its consequences. 101 A real person’s motives are usually mixed, but we are dealing here with theoretical, pure types.

Chapter Seven A MATHEMATICAL MODEL OF THE THREE FUNDAMENTAL MECHANISMS Don’t be concerned about your difficulties in mathematics — I assure you mine are greater. (Albert Einstein in a letter in response to a girl telling of her troubles with math) So far, we have described libertarianism and have presented a priori arguments in its favor and against it. We saw, primarily in chapters two and six, that there are three models for describing how events and actions occur in the world: Determinism, randomness, and free choice. From the next chapter onward, we will deal primarily with the scientific questions regarding the relation between the brain and cognition. However, before we do that, I wish to use this interim chapter to describe another mathematical angle which at the very least proves the consistency and possibility of the third mechanism, free choice. I will discuss the mathematical argument presented by my friend, mathematician Prof. Moshe Koppel, in the second chapter of his book Meta-Halacha (pp. 15-32).102 His book deals with halacha, its character and nature, and he makes uses of the autonomy of human action and freedom of will to demonstrate the autonomous character of halacha itself. Although he uses formal thinking, I will present it in an accessible manner, without making use of real mathematical symbolism. However, readers not fond of math can simply skip this chapter.

The Fundamental Contradiction Koppel begins his second chapter by pointing to two contradictory characteristics of the idea of autonomy: On the one hand, autonomous behavior is not open to prediction. When an individual acts autonomously, we cannot predict what he will do. Every prediction is based on recognizing existing reality, but autonomous behavior by its very nature is not deterministically set by present conditions. On the other hand, it is entirely clear that autonomous behavior cannot be random, as it would then not be true autonomy. The result of a lottery is not the same as a decision made by an individual human being. The difference is that the individual has judgment and justifications for what he did and thought. But if he does, why can’t we predict it? Put briefly, he says: All human behavior is random or causally determined. If free choice is neither, then it would seem not to exist. I already dealt with this issue in chapter two, when I said that choice is a third mechanism, different from randomness and determinism. In chapter six, I added to this and showed that this distinction regards judgment, which ties into what I called “sufficient reason,” as opposed to causality-related determinism. We saw there that deterministic behavior is based on reasons, random behavior on nothing, and free choice on reasons. In other words, the assumption at the heart of the problem as presented by Koppel is incorrect: It is not true that all behavior is determined causally or randomly. There is a third possibility — behavior determined by reasons and not by causes. Therefore, even though our behavior has justifications or reasons, it cannot be predicted in advance (as it has no causes).

Methodology To explain his way out of this dilemma, Koppel argues that autonomy, randomness, and determinism are not traits of the formation of behavior but of behavior itself. Moreover, he claims that these also do not refer to particular conduct at a particular time, but to behavior along the timeline. I will state already here that my argument in the book runs in the opposite direction. I argue that there are also three possibilities for the formation of a single momentary decision: based on a cause, judgment, or randomly. There is no need to go so far afield as to discuss behavior along the timeline to characterize these three types of behaviors. Still, Koppel’s picture is illuminating and allows us to sharpen this triple distinction, and I therefore consider it worthy of being presented here.

The Basic Model Koppel presents an example which accompanies him in his analysis. Let’s look at a person who tends to

sleep every afternoon. This is behavior which is entirely predictable. It is likely that he will sleep this afternoon, as well. Koppel argues that this is not autonomous behavior, as it is entirely predictable. I will note here that this statement is problematic: There is an option by which a person decides freely to sleep every afternoon, and this is certainly autonomous behavior. It may be predictable, but it is still autonomous. This can be explained in two ways: (1) The prediction is not certain, as these are decisions made every day; (2) The prediction is certain, yet his behavior is still based on conscious decisions to sleep every afternoon, and that is certainly autonomous behavior. It is more correct to say that such behavior in and of itself is not necessarily interpretable as autonomous behavior; it can be, but it is not necessarily so. In other words: Observing such behavior does not in itself point to autonomous activity, but it still may be autonomous. As noted, Koppel tries to characterize autonomy based on conduct, as an outside observer of human behavior. Koppel searches for a definition of behavior which can be objectively diagnosed as autonomous, beyond the autonomous and psychological mechanism. By contrast, here I define it based on my internal view of myself and how I make decisions. The sleeping man himself could certainly know if he decided to go to sleep or was forced to do so. Only from the outside does this appear to not be clear-cut and understandable in two ways. Koppel goes on to say that when a person sleeps in the afternoon only on some days without any regularity, then there is no fixed pattern we can point to. Here, too, he argues, we should not consider this autonomous behavior. Again, Koppel deals with the matter as an outside observer, and the observer cannot know if this person decides autonomously or not. But the person himself can certainly explain to himself why he was tired and decided to go to sleep on certain days, while he preferred to spend his afternoons on other days differently. Koppel then goes on to say that these two examples are extreme: One is entirely fixed and the other is entirely random. But if we survey all the options of an afternoon nap, we will encounter patterns which are partially fixed and partially random. For instance, a person sleeps every Saturday afternoon, and on other days he sometimes sleeps without any defined pattern. In such a case, we can of course predict his behavior on Saturdays, but not on other days. This is a mixed pattern, of causally fixed behavior and random behavior. It is therefore partially predictable, but not fully. We can represent such a series of data in a numerical series of 0s and 1s, with 0 representing when the person does not sleep in the afternoon and 1 representing when he does. For instance, the series (1,1,1,0,0,1,0,1,0,0) describes a situation in which the person sleeps on Sunday, Monday, and Tuesday, but not Wednesday or Thursday, on Friday but not on Saturday, on the following Sunday, but not on the following Monday or Tuesday.

Possible Interpretations Let us think now of a person who sleeps on all Saturdays, but who sleeps or doesn’t sleep on all other days in an entirely random manner. When we search for an explanation for this conduct in terms of regularities, we can describe it in a few ways: (A) We could say that he sleeps every Saturday but sleeps randomly on all other days; (B) We can of course say that he sleeps randomly on all days, and it just so happens that the “lottery” resulted in him sleeping on all Saturdays; (C) We can find regularity which would explain the entire series causally, if we create a function which gives us a 1 on all the days he slept and 0 for the others. We can of course create such a function for any finite series. If we adopt explanation A we could predict his situation only for Saturdays, but not for the rest. If we adopt explanation B, we will not even be able to predict his behavior for Saturdays (as even Saturdays are seen as a coincidental result). According to explanation C, we will be able to predict his situation on all days. Explanation B is called “underfitting” in mathematics, as we ignore a clear structure which exists with some of the data. Explanation C, by contrast, fits the definition of “overfitting,” as it assumes that there is no randomness even though we have no clear indication of this (as any such numerical series can be explained in this way). Another form of overfitting is to assume that the explanation is that the person sleeps on all the Saturdays on even years (the year happened to be even), and on other days it was a random decision. This also fits the data, but we have no indication from the data itself that his sleeping on Saturdays is just on even years. To the same extent, we could have decided that it only happens in years which divide by 2012, for instance, and it would still be true. This is sometimes called “overcomplication.” To avoid explanations of this sort, we use Ockham’s Razor, which tells us to adopt the minimal explanation, meaning the shortest and simplest one. When one can explain a given phenomenon in a simple manner (i.e., that he sleeps every Saturday afternoon), there is no need to add another, superfluous

data point (that it only happens in years divisible by 2012 without a remainder). The data is explained just as well without this addition, and is therefore unnecessary. It makes the explanation more convoluted, and we have no indication that it is correct. That doesn’t mean it isn’t, just that we have no proof that it is (in the terminology we used above in chapter six: It’s not hopeless but it lacks foundation). Now think of this sort of numerical series describing the behavior of a given individual, whose behavior we must now classify as autonomous or not. We will start with the following series which contains two million numbers: 0,1,0,1,0,1,0,1,0,1…0,1,0,1. This is a series in which 0 and 1 alternate regularly, and there are a million such pairs. Every reasonable person will tell us that this is an alternating series, and this will predict that the next number will, in this case, be 0. In other words, interpretation and prediction are connected. When we have interpretation, we have prediction, and the other way around. We can of course offer a different interpretation: The series alternates every two million numbers. The next two million numbers will only have 1s in places divisible by 17, and the rest will be 0s. Obviously the first interpretation is shorter and neater, and we therefore assume that it is more correct based on Ockham’s Razor. To derive a prediction from the second interpretation is a complete absurdity; it has no basis, and it is therefore unreasonable to base further predictions on it.103 By contrast, a random series has no shorter way to describe it except going one by one, meaning explaining where a 1 is and where a 0 is. There is no general rule, such as alternating numbers, which could describe it in a briefer manner. Of course, there could be longer descriptions, such as describing all the numbers and adding a description of its continuation. These are also possible descriptions which entirely fit the data, but they are longer than one-to-one descriptions. This is what characterizes a random series: Its briefest description is the series itself, as opposed to a non-random series, which has a shorter description (e.g., to speak of the first number 0 and then its alternate, instead of explicitly listing all two million numbers).

A Two-Part Series: Random and Structured A mathematical theorem states that every finite series of such data can be divided into two parts: The first is structured, interpretable and predictable, and the second is random (as we saw above regarding the series of regular sleep on Saturdays and randomly on other days). For instance, a random series of a thousand numbers, which is duplicated again and again. We can describe it more briefly than just listing every single number — we list all the numbers of the first thousand, and then note that the following sets of thousands are duplicates. This is a more complex combination of randomness and regularity, which stands out in the fact that its description can be shorter than enumerating every single number, and therefore contains a regular component beyond the randomness. The length of the description reflects some part of the series and which part is determined by the rules. The mathematical theorem states that any finite series can be described in the shortest manner by combining a description of the structured (regular) description with a random description (a case which is entirely random — where the regular part is empty, and an entirely regular part — where the random part is empty). Thus, the behavior of such a finite series can never deviate from a combination of the structured and the random. Therefore, one cannot speak when discussing a finite series of a third element beyond the deterministic and the random — something which ostensibly makes it difficult to accept the libertarian thesis. But this is because the series of events is finite. The picture is different when it comes to an infinite series.

Modelable Infinite Series In infinite series, the full description of the one-to-one series is infinitely long. Therefore, the criterion of the minimal description is problematic. If there are two infinite descriptions, then it is difficult if not impossible to determine which of them is shorter. What we do in such a case is to take lengthening sections from the series and try to find a minimal description for each such section. Let us look at the following series, 8 numbers in length: 1,0,1,0,1,0,1,0. If we had to describe it in the briefest way possible, we would say that it is alternating, and that the next number would be 1. Let us assume that we accept that the next number in the series will actually be 0. If so, we must forgo the simple description we have clung to thus far and seek another minimal description which will fit the series of data points we have now. In such a situation, we can say that even numbered –places are certainly 0, and describe the odd-numbered places one-to-one. Our predictions for the future results will be determined from now on according to the new interpretation. Now comes the next number, the tenth in the series: Once again, it’s 0. This of course fits our prediction,

which is somewhat confirmed in light of this result. The eleventh number to come is 1. This does not of course contradict our prediction (we have no prediction when it comes to odd-numbered places), but it also does not strengthen our theory or interpretation, as the numbers in the odd-numbered places have no significance in terms of their predictive capability. We will now jump ahead to where we have 32 numbers already. The additional numbers are all 0s, aside from the number 1 in the following places: 13, 17, 19, 23, 29, and 31. This is still consistent with the assumption that there is a 0 in the even places and the rest is random. The proposed description according to this interpretation is: 0 in the even numbered places and random results in the rest. But we should take note that now there is another description which also fits the entire series: 0 everywhere, aside from prime numbered places, where there is a 1. This is of course a shorter description, as the previous one required a one-to-one description of all the numbers in the odd-numbered places. In fact, we could have made this proposal at any prior stage and it would of course fit. But it would not be unequivocal after the first 10 numbers, and then we would enter a situation of incongruence (raising a hypothesis which might fit but which lacks sufficient basis). After we have enough numbers, meaning that the description of 1 in the prime-numbered places has become short and significantly neater than the description of each and every number, here we are already ready to adopt it as the best description. What did we have here? At first we offered a hypothesis of simple alternation. It did not pass the test, and we therefore replaced it with the proposal that the even numbers are 0 and the odd numbers are random. After a while, nothing was harmed, we still had full congruence, yet still, the consideration of overfitting led us to adopt a neater description which also fits the data, with 1 in the prime numbered places and 0 in the rest. Meaning, after we have a certain hypothesis, any additional number which we receive can fall into one of three categories: A. Refuting the previous hypothesis. B. Fitting the previous hypothesis, but leading us to a neater hypothesis, which also fits the data. C. Supports the previous hypothesis. After receiving a great deal of data, we remain with an interpretation or hypothesis which does not change. This interpretation can be entirely structured, like the last case we described, or divided into a structured component and a random component, like the previous hypothesis (0 in the even numbers and random in the rest). In the second case, we only have a partial predictive capability rather than a complete one, but it is the best description (the neatest and briefest) which we have for this numerical chain. We treat these two types of chains as modelable. We have a model for this chain; sometimes it has a random component and sometimes not, just like finite chains. Chains of the first type, entirely structured, are called computable chains. Such chains allow us to predict every number which will appear in the chain in the future. A chain of the second sort is non-computable, since we cannot predict every number within it, but only some of them. As noted, these two chains are modelable, but one is computable and the other not.

Unmodelable Infinite Chains Are there unmodelable chains? Such things are certainly non-existent when it comes to finite chains, but it turns out that infinite chains have a third kind of chain, of the sort which is not modelable. Think of a chain where we get number after number. We start with a specific proposal, after which we add a random component. The more the numbers flow in, the more we suddenly find a neater description which also includes part of the random numbers, and the random element thus reduces in size. The numbers continue to come in and once again we find a neater description, which further erodes the random part, and so it continues. With the chain of one in the prime-numbered places, for instance, we saw that this happened twice. Let us think that now after 32 numbers, random numbers start to appear for the next 32 places. Afterwards we once again see the previous structure repeat for the following 32 numbers, followed again by 32 random numbers. We will therefore change the description to the following one: We will divide the series into sections of 32 numbers; 1 in the prime-numbered places in all the 32 numbers in the odd-numbered sections, and randomly in the even-numbered sections. This way we can continue to increase the

dimensions of the description (e.g., a collection of five 32-number sections becomes a new super-section). This is a series of numbers which the more it goes on, the more we find neater and neater ways of describing it which erode the random part, but it never actually ends. There will always be some element of randomness. We will never be able to reach a complete model for this chain, and not even a model made up of a structured part and a random part. We can’t permanently determine the random part of this series. This process could go on forever. Each time, we find a deeper theory which describes a larger section of the series, but we never reach a final and complete theory for the entire set of numbers. The part subject to prediction always increases but is never complete. In such series, the structured part is infinite, but always leaves us with a random part, Such a series is not modelable; we have no final model for it (not even a model which contains a defined random component).

Solving the Autonomy Problem: Back to the Topographical Layout and Probability Thus, unmodelable series solve the formal problem of autonomy. Infinite series don’t necessarily divide into random and structured components, nor a combination of the two. In the description thus far we saw a third kind of series, an unmodelable one. It is neither random nor structured. There is a creativity here which nevertheless continues onward, and is not predictable. On the other hand, it is not entirely random, meaning it is not something which lacks any discernable structure or order. There is some form of structure, but it is not modelable. Thus, it is a phenomenon which is truly parallel to human action through choice, as opposed to random activity (which is not predictable at all) and deterministic activity (which is entirely predictable). In such chains, no final series of numbers, no matter how long, can be sufficient for predicting what the following numbers will be. On the other hand, it is clear that use of the partial models and data we did arrive at will produce better predictions than simple lotteries. In this sense, this series is fundamentally different than a random series (where a lottery would give us the best shot at predicting the results). At the same time, when we know someone and how he addresses matters and acts, we can try and predict his future behavior. Such prediction would be better than a mere lottery, or a prediction of the behavior of an unfamiliar person, but we would still never be able to perfectly predict his actions. There will always be the possibility of surprises at points where he chooses a new path. By contrast, a machine or natural process which operates according to regularities — can be predicted, at least partially. Even if the computable portion of the process is incomplete, meaning that it has a random element as well, we can predict the conduct, at least at the partial level of the computable, structured part, and do lotteries for the rest. So, human behavior is characterized by the lack of modeling ability, as opposed to machines which operate solely based on randomness and/or deterministic mechanisms.

The Significance of Koppel’s Proof Koppel identifies autonomy with unmodelable behavior. Although the mathematical distinction he proposes allows us to sharpen the meaning of freedom of will as a mechanism (!) compared to determinism and randomness, we still see a clear distinction between his description and my own. We can understand it in the following way: A person lives for a certain length of time (certainly not an infinite length). If we take the sum of his behaviors over his lifetime, we can in principle understand him completely, as it is a finite chain of events, and every finite chain is modelable (even if not fully computable). Thus, according to Koppel’s formal definition, a real person cannot truly be autonomous. Only a creature which lives an infinite amount of time can be autonomous according to this mathematical definition. It’s true that we may not be able to predict the future behavior of a person in light of what we know so far, but this characteristic also exists when it comes to chains which are entirely computable and are of finite length, if we only form their complete model after we have all the data (i.e., when the last data point is also important for understanding the model at the basis of the data chain). As opposed to Koppel’s formalism, I define autonomy or free will in terms of the singular act, and not through a phenomenology of a chain of actions. The inability to build a model is a metric of autonomy, but it is not the definition of autonomy itself. Put differently, an action which cannot be modeled is autonomous, but an autonomous actions is not necessarily always unmodelable. For instance, if a person chooses to act in a coherent manner for certain values (sleep every afternoon, for instance), we can predict his behavior with a great degree of certainty. On the other hand, it is certainly dependent on his choice, as he could choose otherwise. So, in practice his chain of behavior is modelable, and even entirely computable, but he is still clearly an autonomous person who makes decisions based on his free choice. Koppel himself points this out when he explains that autonomy is defined by the ability of a person to act

in a manner which is unmodelable. If a person chose to act in a modelable manner, this does not make him a non-autonomous being. His fundamental character is determined according to potential and not effective behavior. This means that Koppel deals with chains as an indication of autonomy, but he also agrees that autonomy can be defined by a single action. He effectively only proves the possibility of autonomy but does not define it. Koppel’s description helps us define the third phenomenology, and prove in a formal mathematical manner that randomness and structure are not the only two possible categories. In this sense, its formalism has great importance. It proves the possibility of freedom of will and of autonomous action as a distinct category, different than both causality and randomness. But to directly understand what autonomy really is, then after we mathematically proved the existence of such a third category, we must implement this distinction regarding an isolated action. Indeed, as we saw in chapters two and six, one can make such a distinction when it comes to a person’s single action as well: If it is done due to a cause, it is predictable, meaning it is computable. If it is done without any cause and arbitrarily — it is random. And if it is done for a purpose or due to a reason — it is autonomous. According to this definition, computability does not characterize a series of actions but a single action. In the determinist-computational picture, if a person looks inside himself (or we find a way to look inside him) and he knows what drives him, he can calculate the results of his actions based on that. The outside observer, who does not know how to go inside his thoughts, is the one who requires the entire series of previous actions to prepare a prediction (build a hypothesis or an interpretation), but this is solely a technical matter. In principle, according to the deterministic picture, the one who knows the person at this moment can predict what he will do the next. By contrast, a choosing person is one who acts autonomously. Regarding such a person, acquaintance, whether it be momentary or done based on generalization from past actions, can indeed be beneficial and lead to better predictions, as it ultimately reveals the topographical layout in which the person acts at this given moment, but it cannot perfectly predict what he will do within it. In the second chapter, we already saw that it is clear even to the libertarian that probabilistic results also work regarding the actions of people (because of their surrounding topography). But we also saw that acquaintance with the entire situation in all its details is still insufficient to give us a guaranteed prediction of the future action of an individual person. As we saw, even after taking the influence of topography into account, there is still the element of that person’s choice which will additionally affect the final outcome. This cannot be predicted, even though this mechanism is not random (meaning that the prediction will provide a better result than a lottery, but not a perfect one). My arguments here are not directed at those of Koppel. Meaning, Koppel was discussing the nature of halacha, and not the nature of individuals and their choices, and therefore his characterization is certainly appropriate and correct. He is dealing with an area of ethics, and not an individual acting within that framework. His main argument is that halacha is built in a manner which is unmodelable. He compares it to terms like beauty or aesthetic quality which can also be said to be unmodelable. The same is true of the series of true statements in number theory, or the collection of correct norms of what is allowed and not allowed according to Jewish halacha, to which we can apply the same perspective.104 But we are interested here in the person, and in this context we must characterize autonomy via an individual and momentary action by that individual, and not based on a generalization made by an outside observer based on a collection of actions. To characterize halacha, we must use all the data, examine them from the outside, and consider if the field is modelable. This is much like the relationship between the question of whether the collection of correct halachic values is modelable and the question of whether an individual is acting autonomously when he chooses any given value. There is definitely a connection between the two questions, and the autonomy of these two areas is not detached from the autonomous character of the individual bringing them about. Thus, it is obvious that the character of halacha is the result of the character of the person who creates it; a person who thought and acted autonomously can create a group of (halachic) arguments which are unmodelable. But the first question is objective, as it deals with a collection of arguments or behaviors, and must therefore be dealt with via a mathematical analysis based on an outside observation of a collection of data. By contrast, the second question is subjective, as it deals with the decision-making or judgment of the individual person. By the nature of things, one cannot deal with such a question in a direct, mathematical manner, and there is therefore importance in mathematically coping with the first question, if only for the purpose of proving the existence of a third category — but in the end, we must implement it and apply it to the thought and action of the specific individual.

Intuition How can one think of areas which are unmodelable? If we don’t use rules or some sort of model, to analyze the beautiful, the good, or the halachically correct (we certainly would not use lotteries), then what tool do we use? The tool people use to act in complex areas is called intuition. This is the human tool for dealing with unmodelable areas. Moral intuition creates values, and cognitive intuition creates scientific generalizations. Intuition, which is the human skill which aids the individual to deal with questions and contexts which are unmodelable, can create collections of arguments or behaviors, which are also unmodelable. No-one disputes that there are intuitions which can be modeled. These are situation in which the individual intuitively reaches an answer, but he is unaware that he has actually conducted an unconscious series of calculations to reach it. Here, intuition is merely a shortcut for systematic logical (and modelable) thinking. But the argument of the libertarian is that there are also other situations, ones in which there is no possibility to calculate based on a model. According to this proposal, our intuition succeeds in doing something which no computing model can do and for which no computing model is applicable. In my book Two Carts and a Hot Air Balloon, I pointed to human intuition as being the antithesis of analytical thinking, the form of thinking based on rigid calculation based on models (in the terms we used here). There, I called intuition synthetic thinking. This is the cognitive equivalent of free choice when it comes to values. In chapter four above, in the discussion of the fool’s cycle of determinism, we assumed that our thinking is also done freely (based on judgment), which is equivalent in some way to free choice in value contexts. We saw there that without this assumption, there would appear to be no basis for our faith in the laws of nature and the various generalizations we carry out. Just as there is an understanding of decision-making in the area of values as something based on calculation, while the libertarian sees it as the result of judgment, so, too, when it comes to thinking about facts. When it comes to the latter, we can see acts like abstraction and generalization as analytical actions (computation based on modeling) or as intuitive-synthetic acts.

Are There Processes Which Are Unmodelable? In this chapter, we saw how unmodelable processes are defined processes (at least negatively so), and that they are possible in principle. Koppel proved that the assumption that only random or computed processes or their combination exist in the world is incorrect. His argument leads to the possibility of consistently defining unmodelable processes without running into contradictions.105 But proving the possibility and consistency of such processes is no proof of their actual existence. The question we face now is whether such processes actually exist. Specifically, is our cognition as we understand it? Do our choices of values or other cognitive decisions have such a character, or are they perhaps the result of processes based on a structured model? The Church-Turing hypothesis on the physical world is that physics is computational, because all the laws which control it are deterministic and structural. If we wish to include random dimensions within it, we must expand this thesis and say that physics is modelable (we may not have a complete model, and we may never have one, but this is due to a lack of information. In principle, it is modelable, meaning such a complete model can exist in theory). This thesis is not proven; it is only a hypothesis, based on the fact that it appears that science works pretty well, meaning that we have been successful in discovering structural laws which control physics and use them to predict future behaviors (future results of experiments). The laws of physics are the models we build for the chains of data provided by nature (and our labs). But it’s important to avoid becoming too beholden to this thesis; those involved in the area known in mathematics and physics as “complexity” repeatedly warn us on this point. This is an area which discusses emergence (see below in chapter thirteen) and phenomena which cannot be placed on a basis in which it is entirely calculated (meaning, it cannot be reduced to microscopic laws). On the other hand, in light of the success basic physics has had with simple phenomena, this assumption seems fairly convincing to us. The accepted assumption among physicists is that we can base complex phenomena on microscopic laws (i.e., reductionism), but that we meantime lack the knowledge or the computational ability, and we can therefore not actually do it in practice. We may never create a complete structural description in the form of a perfect model, but the assumption is that this is possible in principle. The argument of many in the field known as artificial intelligence is that this is true regarding all physical phenomena, including the human individual, as a physical-biological creature subject to the Church-Turing thesis. They argue that the cognitive actions of human beings can be entirely modeled. Douglas Hofstadter, in his monumental and fascinating book Gödel, Escher, Bach: An Eternal Golden Braid,106 defends this

thesis in a very enjoyable manner. It continues to be strengthened in light of the recent advances in the neurosciences. This assumption is based on various arguments which tie our cognition to the activity of our brain. This of course raises questions regarding the relationship between the brain and mental phenomena, and we will deal with this extensively below. 102 Moshe Koppel, Meta-Halakha: Logic, Intuition, and the Unfolding of Jewish Law, Jason Aronson inc., USA, 1997. 103 This is precisely the problem which Wittgenstein presents as following a rule. See, e.g.: Saul A. Kripke, Wittgenstein on Rules and Private Languages, Harvard University Press, 1982. 104 There the issue is Gödel’s theorem, or Turing’s stopping problem, or a class of productive sets, meaning a series of approximations which approach but never reach a full description of the area under discussion. 105 It’s worth remembering Newcomb’s paradox which we mentioned in chapter four, where we noted that in order to defend his position, the determinist must conclude that free will is not defined conceptually, not just that it does not exist in reality. Here we have a strong argument against this position. 106 This thick book explains number theory and the paradoxes it involves, in Gödel’s theorem and Turing’s computation problem, while weaving it all through the painting methods of Escher and the compositions of Bach. This fascinating triple braid is the subject of this wonderful book, which — justifiably — earned Hofstadter a Pulitzer Prize.

Part Two The Whole World is Physics? In part one, we presented the two pictures, determinism and libertarianism, next to each other. The discussion focused on the primary difficulty encountered by libertarianism: How does it sit with the principle of causality? Many modern discussions of this question have made use of “holes,” non-causal regions in the framework of physics, specifically chaos theory and quantum theory. These defenders of libertarianism effectively admit (not always knowingly) that the materialist-physicalist picture is true, according to which the whole world is physics, and they therefore try to introduce freedom of will into the gaps which exist within physics itself. But as we will see in this part of the book, these attempts are based on a misunderstanding of freedom of will or a misunderstanding of chaos theory and quantum theory. We will see that these attempts are doomed to failure, and that in a sense, the whole world is indeed physics, at least at the material level of reality. The non-causal gaps in physics, to the extent they exist, cannot truly contain free will, and libertarianism can therefore not exist alongside physicalism. We will see here that the libertarian cannot avoid a dualist position, according to which our world (or at least ourselves) contains something beyond the material, which as we saw in chapter six is exempt from physics. Moreover, as we will see in this part of the book, there must be interaction between this other substance (spirit) and matter, which leads us from materialism to interactionist dualism, and within the human person this is even more sharply expressed, via the interaction between body and spirit, which will lead us later on (in the next part) to deal with the neurosciences. So, in this part we will see the connection between the two questions which so far have been presented as ostensibly independent of each other: The question of determinism VS libertarianism and the question of materialism VS dualism. We will see here that there is a connection between the questions, even though it is not logically necessary. In the end, the ultimate conflict will be between deterministic materialism and interactionist dualism. In chapter eight, we will present a philosophical map of the relations between matter and human spirit, and we will reject some of the theoretical possibilities which arise every so often in philosophical discussions of these questions. In chapter nine, we will deal with chaos theory and see that the libertarian’s salvation will not come from here. Chapter ten will similarly examine quantum theory. After we connect the two questions, we will return in chapter eleven to examine the philosophical questions dealing with materialism.

Chapter Eight BETWEEN THE MENTAL AND THE PHYSICAL: A PHILOSOPHICAL MAP If you lose your soul and are aware of it — you still have a soul left to lose. Charles Bukowski In chapter six, I pointed to the way of thinking which led people to adopt a deterministic approach. We saw there that the fundamental principle at the basis of the deterministic view is the principle of causality, which leads to the rejection of the intuition of free will. In chapter three, I presented a different basis for the deterministic approach, in which by force of theological considerations. God’s omnipotence and omniscience leads us to future events being dictated in advance and unchangeable. In both cases, we are dealing with determinism, but both approaches rely on different assumptions. Causal determinism is usually based on a physicalist worldview. Its foundational assumption is that all physical process are solely determined by material factors operating according to the laws of physics. In chapter six, we explained that freedom of will is considered problematic according to this view since it is contrary to the physicalist approach. Theological determinism, by contrast, is based on God knowing everything in advance. His knowledge is not limited to physical events, and it turns out that if he is allpowerful then he can also know what we will experience and what we will feel, meaning future mental events. Therefore, a deterministic approach on this basis does not necessarily assume physicalism; from a theological perspective, even if it appears incorrect to say that all the world is physics, meaning that the world also contains a spiritual dimension, we still have not extricated ourselves from determinism.107 Our interest in this book is causal determinism, if only for the reason that as we saw, the theological consideration is not really convincing. Moreover, the basis for most deterministic approaches nowadays, certainly those based on science, and especially neuroscience, is physicalism. In this chapter, I would like to go deeper into the connection between determinism, physicalism and materialism.

What is the Mental? Almost no-one disputes that mental events occur within human beings. An individual human being wants, thinks, identifies, decides, considers, plans, remembers, loves, becomes sad, and so on. Every such event is also divided into a number of specific phenomena. Thinking can be applied to a math problem or a historic event. Will can be the will to marry someone, fly to the moon, or eat chocolate. A decision can involve choosing a profession, buying a house, or watching a movie. Emotion can be love for Shimon, hatred for Levi, jealousy of Rachel, and so on. All these are events which the individual experiences in an unmediated manner, and it is therefore difficult to deny their existence. When I see a body move, the bodily movement is of course a physical event, since the body is a physical entity. But the sight which I am watching, the moving body, exists only in my consciousness. It is a result of my system of consciousness, which is not necessarily a simple reflection of external reality but a processing and presentation of the outside world in a unique language and framework. This sight is a mental phenomenon or event. The table before me has color. That color is the result of the light bouncing off the table being of a certain wavelength. The electromagnetic wave is a (somewhat abstract) physical object which exists in external reality. But the color was only created within my consciousness. The world itself has no colors, or at least, we have no reason to assume that they are out there. Color is the form which my brain chose to represent electromagnetic waves to me. Different wavelengths along the spectrum are expressed in my consciousness as different colors due to sensors in my brain. If we decide to be more general, this is not just true of colors. Even light as light is not a phenomenon which exists in the external world. There is an electromagnetic wave at a visible wavelength (meaning that our sight can notice it). Light, like the sight it creates, exist only in our consciousness. The same is true of sound, smell, sense of touch, and other experiences. If we think of a creature whose brain wiring is different — say, that their eyes are connected to the hearing center of the brain — then he will hear sights instead of seeing them. What separates me and him is not what happens in the outside world, as we both experience the same electromagnetic wave. What makes us different is our brain and consciousness structure,

resulting in a different sense of perception (what sight I see, and whether it’s a sight or a sound or something else entirely). We can demonstrate this with a question which seems to confuse many: Does a tree which falls in an empty forest (i.e., empty of creatures which can hear) make a sound? The answer is of course in the negative. It would appear to create an acoustic wave (air movements and pressure changes), but sound is a mental phenomenon which exists only in the consciousness which absorbs this acoustic wave (e.g., when the acoustic wave hits the human ear drum). Without such a consciousness in the area — the tree makes no sound; it only moves air. The confusion regarding this question is a result of the failure to distinguish between the acoustic wave (which is a physical phenomenon) and the sound it makes (which is a mental phenomenon). It is not just the sights and the sounds (my various conscious recognitions) which are mental events. My watching and listening to all this counts as a mental event. It is an action which I do, and it takes place within me, in some way. It is not accessible to others. The external object is accessible to all, but the internal experience we will feel is not like this. Every individual has access solely to his internal experience and not that of others. This subjective dimension of our consciousness is commonly called qualia. Hence the problem known as the “philosophers’ chestnut”: Is the color I call “red” the same color you call “red”? It may be that what you see and call “red,” I see and call “green.” Thus, when I look at a watermelon I describe it as green, but I effectively see it in the color you call red. Meanwhile, you see it in another color, but you also call it “green.” We’ve become used to this terminology from childhood, and we therefore both speak of “green,” but what we see in our internal consciousness is entirely different. Each one of us continues to use their term, without noticing the difference. This is a possible picture, and we have no way of knowing if it’s true, as all these sights are the individual experience of every one of us; they are inaccessible to any measuring device, as measuring devices can deal with the outside world but not our internal experience. This is just an example, and we could of course bring others regarding any other form of sensory perception.

The Interaction Between the Mental and the Physical It’s fairly clear that there is some mutual interaction between mental events and the body and the physical world. When I receive a blow, I feel pain. The blow is physical, but the resulting feeling of pain is a mental event. Pain is a kind of experience, and it certainly cannot be identified with the blow, which is the factor which engendered it. This is a clear example of how the physical affects the mental. By contrast, when an individual is depressed he often eats less, and as a result starts to lose weight. This is influence in the opposite direction: From the mental to the physical. When an individual decides to go watch a play, he goes there. This is also a case of the mental (the decision) affecting the physical (the going). Or, to take another example, I see a person and a feeling regarding that person emerges within me. The vision process begins with physical events; He is standing before me and the light waves from him reach the cornea of my eye. This information then makes it to the brain, and we are still dealing with physical events. But now the mental events begin: At some stage I identify the person before me, an identification. Afterwards, I think and form an opinion about him, this being judgment. Finally, I feel all sorts of emotions about him, as a result of that judgment. Meaning, the physical started a chain of events which continued on to the mental. It is therefore not just difficult to deny the existence of the mental; it is also very hard to deny the mutual influence of the mental and the physical. It is no wonder that there is almost no-one, including the greatest of the materialists (aside from a few confused souls), who denies this. There is wall-to-wall consensus at least as regards the existence of mental events, and perhaps also that we experience them as interacting with the body and the surrounding material world. The question is then: What is the nature of the mental? Must it necessarily assume the existence of spirit, or can it also be explained within a materialistic conceptual framework? This is the modern, contemporary version of the question of materialism.

The Correlation Between the Mental and the Physical So far, we saw the interaction between the physical and mental levels, but that’s not all. There is also a simultaneous correlation between the two. Today we know that mental events accompany physical events in the brain. Two types of events, brain and mental, take place more or less at the same time. When I identify a person, and as noted identification is a mental action, physical-chemical events take place in my brain. We spoke above of the relation between the person and the light waves bouncing off him towards me and the formation of an identification as being a causal relationship. The physical factors (the light waves) create mental events (the identification). But here we are dealing with the identification action itself,

meaning the result of the light hitting the cornea. When I conduct a mental identification, my brain is conducting very complex physical-chemical activities at the same time which accompany the identification, and some would say even engender it. As we said, this happens more or less at the same time as the identification itself. The situation is much the same when I want, decide, think, feel, and so on. As far as we know today, there is an unequivocal relationship between the mental and the physical — every mental event takes place when there is a parallel neural event in the brain of some kind. There are regions in the brain which are “responsible” for our various actions. There are regions responsible for identification of people, decision-making, love, mathematical thought, artistic creation, aesthetic judgment, and so on. In many cases, we are dealing with a multi-region brain event, meaning a mental event whose associated brain event occurs in a number of regions of the brain, at the same time or one after the other. We will go deeper into this issue in part three of the book. For now, we will suffice with the statement that it is commonly accepted that every mental event (love or jealousy) has a physical correlate in the brain (an electrical awakening of a region or regions in the brain). Thus does every specific mental event (love for Leah or jealousy for Rachel) have a specific physical configuration in the brain to express it. Above, we saw that the relationship between the mental and the physical (such as between pain and a blow) is a causal one. But what is the nature of the relation between the mental and the neural (such as between the pain and its neural picture in the brain). Is this also a case of cause and effect, or perhaps there is a different relationship here? This correlation between the mental and the (physical) brain is the driver of the modern, contemporary body-soul question, and it also touches on issues of causality and determinism. To sum up, there are two main questions which we must deal with in the context of the relation between the mental and the physical: A. Do these two levels truly exist, or does only one of them exist while the other one is fiction? When it comes to this question, there are three different answers based on three philosophical approaches: Materialism (only matter), dualism (spirit and matter), and idealism (only spirit). Even if we are materialists, this does not necessarily mean that we deny the existence of mental phenomena. There are those whose approach entirely denies the existence of the mental and identifies it entirely with the physical brain. But this is simply a conceptual confusion, as I will immediately explain. As noted above, it’s hard to deny the very existence of phenomena which accompany each one of us at the most intimate level and at all times. B. The second question is dependent on the first: If they both exist, what is the relationship between the two? This is a body-soul question, or put differently: The psychophysical problem. The question of the relationship between these levels also leads to a number of possible answers: Is one of the levels the cause of the other, or do they operate in tandem without relation to one another (like the synchronicity we discussed in the first part of chapter five), or is one level simply an aspect, or a reflection of the other? As noted, these two questions are interrelated. If one of the levels is just an aspect of the other, then it’s not clear if we can say that both exist or whether they are effectively the same phenomenon. If one is the cause of the other, then it is clear that both exist in a relationship of cause and effect. The best way to deal with these questions is therefore not to separate but to briefly present the different answers to both of them at once.

I am My Body: Hard Materialism Richard Taylor, in the second chapter of his book Metaphysics, which we’ve already mentioned more than

once, begins his discussion of the body-soul problem in a slightly different manner: He discusses the relationship between me and my body. Some identify the two concepts; they think I am nothing more than my body. In response, Taylor argues that when we identify two things as one, like David Ben Gurion and the first Prime Minister of Israel, our argument is that it is the same person, who is first identified by name and second by his position. Thus, any description on the subject of the name needs to fit the position holder in some way. If Ben Gurion tended to stand on his head, the first Prime Minister of Israel also tended to stand on his head. If the person by this name lived with Paula, the person described as holding this position also lived with Paula. If there is a description which fits one and not the other, this is a sign that we are dealing with two objects and not one object or person. Now we must ask, can everything that can be said about my body be said about me, and vice versa? For instance, I believe that it is now night (an utterly false belief). If this belief is nothing more than a brain state, then we must say that my body believes it is now night. That is an odd statement in and of itself. We usually do not say that physical objects believe in something or not. At the same time, the statement that my body wants, thinks, decides, and so on, also sound pretty strange. But let us go one step further. I said that my belief that it is now night is false. What can I say about the brain state? Can a state of a physical object be false? It’s a physical state, a collection of chemical-electrical currents and tensions and nothing more. Falsehood is not a trait of physical states or objects. To sharpen this point: Try to tell a physiologist to look at some brain state, and define it solely based on that observation whether it is true or false. Taylor thus concludes that the identity between me and my body is nothing more than a conceptual confusion.

Dualism Usually, such considerations such as Taylor’s lead us immediately to the dualist alternative. The dualists, one of the most prominent and ancient of these being Plato, see the human being as including two different components: Body and soul. According to their approach, the soul is a separate, non-material substance, within which all mental activities occur. Alongside it is the body, which is responsible for physical activities. The question raised by such simplistic dualism is how the body and soul constitute a single totality; or put differently: Why do we think they belong to the same person? Is there a relationship between what happens in the body and what happens in the soul? What is the nature of this relationship? How is it created? Plato tended to think that the person is the soul, and that the body is a sort of prison holding the soul captive. This does not refer to a physical presence inside the body, as the soul is not a physical entity, and it therefore is not located in any specific place. What this means is that the soul has ownership of the body in the legal-possessive sense. It means that there is some connection between soul and body, a connection which imposes various limitations on what the soul can realize. According to this approach, the person is the soul, and the body is nothing more than one of the objects or tools that the person uses, like a hammer or a car. The body may be forced on us, and we can’t get rid of it or replace it (although today we can replace parts of it), but that doesn’t change its status as nothing more than a tool. To this dualist approach we add what we saw above, that the mental and the physical act on each other in a causal manner. A dualist approach therefore often appears along with interactionism. This aligns with the description we encountered in the previous chapters, according to which libertarianism necessarily leads to an approach by which the (mental) human will can move electrons or create force fields. This is precisely the meaning of interactionism, as we have influence which works in both directions: From the mental to the physical and vice versa. These two levels, which exist at the same time, influence each other. But this is of course not a necessary approach of the dualist framework. Dualism can appear in other variants which are not interactionalist. Thomas Henry Huxley,108 for instance, proposed the epiphenomenalist approach already in the nineteenth century.109 This is an approach which argues that spirit cannot influence the body as it contradicts the laws of physics. Epiphenomenalism only recognizes influence in one direction: Body influences spirit, not the other way around. We should note that this thesis may sound reasonable at first glance, but it is in fact a very strange one, entirely contrary to our intuitions. We saw above that at least when it comes to experiences, it is clear that influence is a two-way street. The epiphenomenalist will explain every influence of spirit on body as influence which moves from one physical event to another physical event via the spirit — there is state b1, which creates mental state m1, and at the same time or a little afterward, physical state b1 also creates physical state b2. Our feeling is that state b2 was created by mental state m1, but this is just an illusion. In practice, according to this system, b2 only comes after m1, but the real creator of b2 is physical state b1.110

Naturally, epiphenomenalism is deterministic (or compatibilist, which as we noted is the same thing), as the implicit assumption is that only physical causes create physical states, leading to an entirely physicalist picture. Mental events only accompany physical events in a passive manner, but they are influenced but do not influence. They are side effects, epiphenomena, of physical conduct, but they have no part in it. Others argue that physical and mental events occur at the same time, in a synchronized but not caused manner (see the discussion of synchronicity in the first part of chapter five). According to this approach, there is no influence in either direction. Just as spirit does not influence matter, so matter does not influence spirit. This is the parallelist approach. Leibniz even proposed the idea that this parallel was created from the beginning by God, and this is what is called occasionalism. Here, too, we have a speculative hypothesis, which does not at all fit our intuitions. In fact, such an approach includes almost a logical contradiction, as if there cannot be interaction between the material and the mental, then how can we know of the existence of the material? After all, knowledge of the material exists in the spirit; the recognition of something is done by interaction with it (like the exchange of light rays or the absorption of sound waves and the like). These wild speculations are attempts to reconcile dualistic intuitions with the assumptions of physicalism, in a manner which will remove the need for an interactionist argument (which requires recognition of a causal effect of spirit on matter); attempts which, as we see, do not go well.

Softer Materialism I already described the approach identifying me with my body which Taylor critiqued. I did not use the word materialism, and deliberately so. Contrary to what Taylor wrote, materialism does not necessarily identify me with the body. This materialistic position is a straw man, too easily attacked by critics. The argument of the materialist can be softer. He can argue that I am not an entity in addition to the body, but rather a state or a trait of the body. For instance, we say that the table’s color is green (we do not mean the painted color, but the natural color of the wood). Now we will ask: Is the color of the table an entity in addition to the table itself? We will usually not define it in this way. Color is not an object or an entity but a trait of the table. We will therefore not say that we have two things before us — the table and its color, but that we have one object, the table, which has this or that trait, like a green color. In the same vein, my intelligence is not a separate object than myself, it is a trait of mine. The same is true of my good-heartedness, my tendency to get angry, my love of Leah, and so on. According to this proposal, the mental is not a separate entity but a trait of the physical. Such materialism is based on a phenomenon which physicists and neuroscientists have in recent years called emergence. This is a name for states in which some material collection has a collective trait which does not characterize its parts. John Searle, in his book Mind, Brains and Science, proposes an example to clarify this: Water at room temperature has the trait of liquidity. Every collection of water is a collection of water molecules (H2O). Is liquidity a characteristic of a single water molecule? Of course not. A single molecule cannot be a liquid or solid or a gas. Such states of matter are only characterized as being collections of molecules. If so, argues Searle, then we have a trait here which only characterizes the collective, but none of its parts. The emergence argument in our context is that this is also the relationship between the physical and the mental. The mental is a collective trait of the body. Materialism does argue that the body is nothing more than a collection of organic molecules, and a single molecule certainly does not possess consciousness, emotions, thought, or will, but these traits do exist at the collective level; the collection of molecules we call the human body (or the human brain) is graced with them. We will return to this issue in chapter thirteen and expand on this matter. Another formulation of soft materialism is the theory of the double aspect. This is an approach which sees the physical and the mental as two sides of the same thing. Others put it a little differently: The body is the thing itself (noumenon, in Kantian terms), while the mental is the appearance of these events in our consciousness (the phenomenon). When we observe our physical or mental occurrences, a mental picture is created in our consciousness. This is how we conceive of our material totality, just as color is a characteristic of the picture which forms in our consciousness when we observe an electromagnetic wave which is outside ourselves. The problem which this formulation raises is, of course, who is the observer, exactly? Whose consciousness forms this picture? After all, acquaintance with the mental occurs only when I observe myself (I have no way to experience another’s mental state in an unmediated manner). Does that mean that the observing consciousness is also a material object? In such a case, we’ve simply pushed the problem back one stage, but are still stuck with the same dilemma. Now, we must ask: What is this observing

consciousness? After all, it does a mental act. Does it also exist only in another consciousness, third in number, which observes it? This is an infinite regress. Therefore, the common approach today, primarily among neuroscientists and philosophers, is the emergent one, and we will entirely ignore the double aspect theory from now on. Be that as it may, these approaches of soft materialism are not exposed to the same objections Taylor made and which we noted above. We can say things about the color which we don’t say about the table. The color is green, the table is not. The color is bright or dull, and the table is not. The color is in our consciousness, and the table is not. Similarly, my moral tendency is to the good, but that is not the case with my body’s tendency. A statement that this table is dull is seen as shorthand for the statement that the table’s color is dull. The statement that I tend to do good means that my soul, meaning that collective trait which describes my body or my brain, includes a tendency to do good. If so, soft materialism remains alive and kicking even after Taylor’s offensive against its hard counterpart. On the other hand, this approach is exposed to another attack: If spirit is simply a form of the body’s appearance, then how can spirit affect the body? After all, spirit is not another entity which can have influence. It is merely the form by which we conceive the body, not something separate from it. As we saw above, the materialist will treat this as an illusion. What we have here is a process of physical causality, meaning state b1 which is accompanied by mental state m1 causing physical state b2. We therefore erroneously see mental state m1 as causing b2. If so, we are back to epiphenomenalism or emergence. We will come back to this question later in the book (primarily in chapter 13).

Summary As we’ve said, this book is interested in the question of determinism and not that of materialism, except to the extent that the latter relates to the former. We will now summarize what we’ve seen so far. We have removed the option of a direct identification of the body with me. This identification is indeed not possible. Even idealism, the idea that the material and bodily are an illusion and that the only thing that exists is spirit, is not a serious alternative for our discussion. This position is certainly not prevalent in our time, and certainly not in light of modern neuroscience (although the prologue does examine the possibility of seeing it from a unique and different angle). Epiphenomenalism is not essentially different from emergence, as both agree that what goes on is controlled by the laws of physics; the question is only whether the mental is another substance which is influenced and not influencing, or whether it’s a trait of the material body itself, but this difference is not significant for our discussion. Leibniz’s approach, according to which we are dealing with synchronicity, appears somewhat mystical (like Smullyan’s astrology) — scientific thinking teaches us that when two things regularly appear one after the other, the former is usually the cause of the latter. Parallelism and occasionalism are therefore also off the table. Even though these approaches appear as options in almost every philosophical text dealing with these issues, I know no-one who seriously believes in them. The conclusion so far is therefore that the determinist position relevant for our time is soft materialism and emergence. It is natural to speak of interactionist dualism from the libertarian side. Ostensibly, we could speak of a materialism in which the mental aspect affects the material and is influenced by it. But we saw that these are two side of the same coin in the materialist world, or the thing itself and its traits; traits do not influence the thing (because they are not something other that is not it, which can influence it), but rather an auxiliary thereto.111 A libertarianism which accepts a situation where the spirit affects the material and vice versa must be dualist. We are thus left to decide between the following two options: Soft materialism, meaning emergence (or epiphenomenalism, which as far as we are concerned represents the same approach) — the occurrences are fundamentally bodily, and the mental is but the collective traits of the body or the brain. Here the material affects the spirit, but the opposite is not the case. Does that necessarily lead to determinism? In my opinion, absolutely yes. An in-depth discussion in the philosophical and scientific literature is taking place on this question, and the following two chapters (and also chapter 13) are devoted to it. Interactionalist dualism — An approach of mutual influence, from spirit to matter and vice

versa, which leads to the libertarian picture described in previous chapters. If indeed, as we will see in the following chapters, that physicalism does not allow for free will, the only way to hold a libertarian position is the interactionist way. The presentation and clarification of this debate are the goals of the present section of the book. I must therefore show that there is no possibility of being a materialist-physicalist while at the same time believing in free will, meaning that free will is necessarily based on dualism. What I must show is that there are no gaps in physicalism, and that is the subject of the next two chapters. 107 See above in the intermezzo in chapter three, in the discussion of Rav Hisdai Crescas’ semi-deterministic approach. 108 Known as “Darwin’s bulldog” due to his energetic defense of the theory of evolution. 109 Cf. the relevant quotes from some of his books and articles in Plantinga’s article referenced on ??? [need to define page numbers first]. 110 As a result of the presentation of the three components of the causal relationship we discussed in chapter five, in the case described here, m1 is prior in time to b2 and is a logical condition thereto but it is not its cause. 111 To my surprise, I discovered a proposal in this direction in Gazzaniga’s book mentioned above.

Chapter Nine GAPS IN PHYSICALISM: CHAOS112 Probability does not come in place of causality and real determination, but only in place of human ignorance. Pierre-Simon Laplace Every year more people are killed by pigs than by sharks. This shows just how bad our danger assessment is. Bruce Schneier In the previous chapter, I presented the various ways in viewing the relationship between the mental and the physical, and bottom line, we are left with a conflict between emergent materialism and interactionist dualism. The latter is a clearly libertarian position, but there are quite a few discussions regarding the former in the philosophical and scientific literature: Can it fit with libertarianism, or does it necessitate determinism? Meaning: Can a person believe that the whole world is physics, while at the same time believe in freedom of will? Many of the libertarian thinkers in our time — physicists, mathematicians, and philosophers — believe that these positions are indeed compatible, and they therefore search for a possible way out for their position within physicalism. This is usually done by pointing to non-deterministic islands or holes in our physics, meaning non-causal physical processes into which a person’s freedom of will can “enter.” In this and the following chapter, I took upon myself an ambitious and far from simple task: To demonstrate in clear and accessible language (i.e., without getting into the complex theoretical details involved, including shortening and simplifying which I hope I will be forgiven), that this is simply not possible.

Gaps in Physical Causality To succeed in introducing any kind of freedom in a physicalist framework, we must find holes, or gaps, in the causality of the laws of physics. In the oldest formulation of the laws of physics, that which preceded the twentieth century and was espoused by people like Galileo and Newton, this does not seem to be at all possible. In this picture of physics, the entire physical reality is controlled by clear laws which determine the dynamics of every particle and every physical entity. Mechanics controls the movement of all bodies with mass, and if we add the rules of electromagnetism, we will end up with more extensive causality, but still one without any gaps. The Newtonian Laws of mechanics and Maxwell’s equations of electromagnetism paint an entirely deterministic picture; no-one disputes that no gaps existed before the twentieth century. The circumstances at any given moment unequivocally determine what will happen next, whether it be due to causal influence (as we interpret the laws of physics) or simply a logical and temporal chain of events (as David Hume claimed). It is no wonder that Laplace113 declared that if he had all the data (i.e., the starting conditions at any given time) and a powerful enough means of calculation, he could describe the history of the world from beginning to end. But this simple, rigid picture starts to show cracks in the twentieth century, when two important theories emerge which undermine the airtight determinism of physics. These are quantum theory and chaos theory. As I already mentioned, quite a few thinkers, philosophers and scientists alike, tried and are still trying to use these theories to introduce freedom of will into the physical system, thus allowing us to be both physicalists and libertarians. In this chapter, I aim to briefly describe the theory of chaos, with quantum theory being reserved for next chapter.114 I intend to show that there are no relevant free gaps in the laws of physics as we understand them today, including those discovered in the twentieth century. This means that the libertarian must abandon physicalism. That does not mean he must abandon the laws of physics themselves; Rather, he must assume that they can be bypassed (as I described in chapter six, he must assume that will is irregular in that it can create physical phenomena on a mental basis but without a physical cause).

The Three Body Problem The roots of this theory lie in the work of French mathematician and physicist Henri Poincare,115 on the threshold of the twentieth century. In 1900, Poincare tried to solve what physicists dealing in mechanics called the three body problem. We already noted the law of gravitation, which states that every two bodies with mass apply gravitational force on each other. According to Newton’s Second Law, which we have mentioned quite often, there is a relationship between force and acceleration. Thus, the gravitational force applied by every massive body on another causes each of them to accelerate, changing their speed. Now let us add another massive body into the mix, a third in number. It also applies gravitational force to the other two and is acted upon in kind. Whenever a body changes its location, the distance between the two bodies changes, thus also changing the force applied between them. This in turn changes the acceleration which changes the location which changes the force, and so on and so forth. It’s easy to see the complex interdependence of the movements and locations on each other. How to describe the movement paths of these three bodies, each under each other’s influence? Ostensibly, this is a relatively simple problem in mechanics. After all, it only deals with three bodies, while there are situations in physics which involve thousands of bodies, if not more. Moreover, this is an important and very practical problem, as it has many applications in the world around us. For instance, the movement of Earth, the sun, and the moon can be roughly described as a three body problem (the other stars are very distant, and their effect on the relationship between these three is negligible). But it turns out that the three body problem is a very difficult mathematical problem. We know of no complete, explicit solution for it, aside from unique situations (there is a hypothesis that its solution cannot be explicitly expressed in mathematical terms, with a function for the location of each body at any given moment). The greatness of Poincare, as a mathematician who believed in intuition, was in his ability to develop various methods to study the characteristics of solutions, without knowing the solutions themselves or arriving at an explicit formulation of one. Poincare discovered that in such a system, there are non-periodic paths which neither grow in distant nor more proximate to a fixed point (which is reminiscent, and not coincidentally, of the unmodelable chains we encountered above in chapter seven). This was the opening point of what became known eighty years later as “chaos.”

The Butterfly Effect The equations of the three bodies are complex, as the variables which appear in them mutually influence each other. The mathematical problem is entirely defined when the starting conditions, the locations and speeds of the three bodies at a specific point in time, are added to it. The equations of motion describe the development of locations and speeds from that moment onward. An explicit solution of these differential equations which fulfills the starting conditions provides us with the value of each quantity, such as location or speed, at any point in time we wish. This means that the future of the three bodies is unequivocally determined by the past. This is a simple deterministic picture. But what do we do when we cannot find an explicit solution for these equations, in the form of a function which describes the values of the given quantity at any point in time? Is there a hint here of deviation from determinism? In contrast to what might be thought, the determinism is expressed by the equations, not the solutions. The very existence of such equations tells us that the conduct here is deterministic, as the equations describe how any given situation determines the situation in the following moments. This is true even when we have no mathematical way of explicitly describing the solutions. But what do we do when we find no explicit solution to these equations? For this purpose, we use what is called digitation, or discretization, of the equations and the solution process. This means that in order to simplify the solution, we assume that time is not a continuous axis, but a collection of isolated points which appear one after the other with some spacing between them, then building the values of the relevant quantities at every point in time based on their values at previous points in time. It is clear that the smaller the gap between the points in the digitation (i.e., the higher the resolution), the closer our solution will be to the real one. At starting point t0, the values of these quantities are given, these being the starting conditions, and the equations use these starting values to calculate the quantities’ values at the following times (t1, t2, t3 … t1,000,000 and so on). Such an action requires repeating many calculations one after the other, with us considering the results from previous stages when conducting calculations at each succeeding stage. The more precise we wish to be, the higher the resolution we must use (i.e., a denser division of time sections), and this will of course require more calculations. It should certainly come as no surprise to discover that significant progress in chaos theory required the use of computers. Only they can conduct such a large number of calculations in a reasonable time frame. If

a person carries out these calculations, he will have to invest his entire life in doing so. Computers in our time can do calculations in ways considered to be science fiction ten years ago, and certainly forty years ago. But in the 1960s, scientists already had (from most people’s point of view nowadays) primitive computers. The next stage in the development in chaos theory was the work of Edward Lorenz, a mathematician involved in meteorology. Lorenz constructed climate simulations, in which he calculated all sorts of meteorological quantities as a function of time based on various differential equations which underwent digitation, meaning he assumed the presence of isolated points in time. In 1961, Lorenz took a collection of data points which came out of his computer in one of his simulations. He got these results on the assumption that the starting conditions, the values in time t0, were given. Based on them, the computer calculated the remaining values at all points in time — for example, from t1 to t1,000. Afterwards, Lorenz decided to run the computer again, only this time he gave the computer the values at time t437 as the starting conditions, expecting to see the calculation results for the following times. He was of course convinced that the system would give him the values of times t438 and onward which would entirely fit the values he had from the previous simulation for the same times. After all, these were differential equations, and as we’ve already explained, in such equations the quantities at a given time are determined according to their values at previous times. If so, Lorenz thought to himself, we can save some computing time and start the repetition of the simulation from the middle, and the results shouldn’t be any different. But to his surprise, it turned out that the series of values he received was a different series. Moreover, these values gradually drew away from the values of the original series the later the point in time he looked at. This looked mathematically impossible. One of the known and proven traits of a differential equation is that its solution is not dependent on the point where you start the calculation process; after all, the solution for such an equation given the starting conditions is unique. For instance, there is a differential equation which describes the dependence of location on time. Let’s assume that we start the calculation from time t0 with location x=0, and the solution gives us the value of the location as x=8 at time t5, and x=17 at time t10. What will happen if we start the calculation from time t5 and location value x=8? The simple mathematical answer is that five time units later, at t10, we will get the location value x=17. The result is not dependent on the point in time where we started the calculation. So how can we explain the difference in Lorenz’s two calculations? Lorenz understood that the difference derives from the fact that in the second calculation, he introduced starting conditions, the values of the quantities at time t437, which were rougher data values. In other words, he listed the times and locations up to three places after the decimal point, while in the first calculation, he used the precise result, which was precisely five points after the decimal point. Let us assume that the value resulting from the first simulation in the computer for time t5 was x=8.47538. Lorenz then took that value as the starting condition for the second calculation, except he put the value at x=8.475. Ostensibly, this tiny difference shouldn’t change anything significantly. The results of the first calculation should provide us with more precise results, but the second calculation should produce results which will be identical to the precise calculation until three places after the decimal point. If the first calculation produced x=9.42926 at time t100, then the second calculation should produce x=9.429 at that point. But it turned out that the second calculation produced an entirely different result. The distance between the solutions grew, without direct relation to the differences in the resolution of the calculation (= the number of digits after the decimal point). This phenomenon was later called the “butterfly effect.” The significance of this effect is that if we have a given situation described by particular differential equations, a slight deviation in the starting values of the quantities which they describe can sometimes create a very large deviation later on. Thus, it is said that sometimes the movement of a butterfly’s wings in China can lead to a tsunami in America, a small change leading to a drastic effect at a sufficiently large distance in terms of time or location. This is a simplistic and inaccurate description of chaos, as we know that if we divert the angle of a laser beam, if we look very far from the source, this deviation will be clearly apparent in the great distance between the two beams. This is not a chaotic phenomenon, as the deviation at a great distance is in direct proportion to the initial deviation. However, in chaotic phenomena, the deviation a long time from now is not in simple proportion to the initial deviation; even the direction of the deviation is undefined. In chaotic systems, the solutions do not flee to very high values, but are instead limited to a specific area of values. Therefore, even at far-away times, an infinite deviation between the two solutions would not develop (see the image of the Lorenz Attractor later in the chapter). It is therefore inaccurate to say that the deviations of the solution, based on a small change of the starting conditions, then increase infinitely. It is more

correct to say that there is no direct or simple connection between the deviation at the start and the deviation at the end. A small change in the beginning can create entirely different phenomena at much later times. As we will see, this is precisely the characteristic which makes prediction and forecasting in these systems difficult. To sharpen the matter further, let us think of someone standing on the second floor of a building, throwing rocks downward. Now let us compare the throwing of a rock downward when it immediately left his right hand stretched straight out, to a similar rock throwing by his left hand also stretched straight out. The results, meaning the locations of the two rocks on the ground, will not be significantly different from each other. The space below is almost identical to the space above. If we move the left hand more to the left, the deviation below will more or less increase accordingly. In such a case, the dependence of the deviation of the result on the initial deviation is very simple, and it is therefore clear that for every tossing of an individual rock, if we know where it came from — we can with a high degree of certainty predict where it will fall. This means that we have the ability to predict the future based on the present in this process, and we are therefore not dealing with a chaotic phenomenon. But what will happen if that person throws two identical pieces of paper in the same manner? Will there be a connection between the space between the hands above and the distance they will land from each other on the ground? It is true that the space below doesn’t have to be very large, but the relation between it and the space above is weak, if not nonexistent. It is obvious that in an experiment in which a piece of paper is thrown down, anyone who tries to predict where it will fall is doing something nonsensical; he has no chance to make such a prediction. This is an example of the chaotic phenomenon I described above, and its consequences for predictions or forecasts.116 In sum, chaotic phenomena are related to systems in which the results are very sensitive to the starting conditions. Even if we only change the starting conditions a little, the results may be drastically different. As we saw, this trait prevents us from predicting the future conduct of that system.

The Consequences of Predictive Capability Some have written that the twentieth century was the century of physics, thanks to three theories: The theory of relativity, quantum theory, and chaos theory. But what is the significance of chaos theory? Why is it considered such a major innovation in the world of classical physics? The innovation of chaos theory primarily has to do with predictive capability — the ability to predict the value of a physical quantity at a later time, given its value at a given earlier time. This of course ties into determinism and causality, as these are predictable causal processes; when there is a causal relationship between A and B, then given A, we can predict that B will occur. What happens in chaos theory is that our predictive capability is mortally wounded, and sometimes simply vanishes. It’s important to remember that in any event, we can never really know the location of anybody with perfect precision. Any numerical data point is dependent on the precision of our measurement tools and their resolution. A ruler which measures length in centimeters cannot provide us with the location of an electron. Even the location of a pin or a needle which is given in terms of a ruler measuring in those units is only a weak value. But even when a body like a soccer ball is measured by such a ruler, it is still not located with perfect precision. However, in a non-chaotic system, the deviation at the end will be in some proportion to the deviation of the initial measurement, and we call such a situation predictable. We do not expect a perfect prediction, but rather a prediction where we have control over the degree of error involved. Meanwhile, in a chaotic system, one whose conduct is highly sensitive based on the starting conditions, a small imprecision in the initial location is small and negligible, but when it comes to later times we have a real problem. We can never say something about the location of the body at any later time, even if we measured its location at the beginning of the process. Take the rock throwing example we presented above. If we know where the rock went on its way with a reasonable amount of precision, we have the possibility of predicting its future location on the ground, at a level of precision proportionate to the precision of the measurement of its location upon being thrown from above. But in a chaotic system, like in the case of the pieces of paper, there is no possibility of measuring where the paper will fall based on its initial location above. Similarly, anyone who measures air movements in China in general, will not be able to notice that there was some small butterfly there which also moved the air a little in a very specific area in that giant country. But if the measurer did not notice it, then his prediction might miss a tornado in Europe, as what caused it was precisely that butterfly’s wing movements. It’s hard to speak of a climate prediction which misses a tornado as falling within the limits of error. It would be more precise to say that it is not a prediction at all. In this sense, there is harm here to the conception of causality. The conception of physical causality is

that given a starting condition, we can calculate, and also predict, the situation at any time in the future. This was the Laplacian optimism described above. We see from the chaos theory that this optimism has no foundation. Even if we know the initial situation — unless to the point of infinite precision, something we cannot achieve and probably never will — we cannot really predict the future. So what do we do in such situations aside from pointing to the existence of wild phenomena lacking meaning or physical order? After all, the role of physics is not just to point to the mess but also make sense of it. Indeed, chaos theory does not just point to the phenomenon, it also allows us to describe and understand these phenomena. We find different characteristics which characterize chaotic systems, meaning characteristics of the equations and situations which cause the chaotic behavior of the solutions. In addition, we also find characteristics of chaotic behavior itself, meaning characteristics of the solutions themselves. For instance, just to take things down to earth: The solutions of a dynamic set of equations are described on a coordinate system which is called the phase space (see, also, the second part of chapter five); for instance, a coordinate system in which the x-axis describes the location and the y-axis describes the speed (or the momentum: mass times speed). At every point in time, the system has a location and a speed, and therefore at a particular time, the system is “at” a particular point in phase space. The behavior of such a system over the time line is effectively some line or path in the phase space. The graphs in the second part of chapter five offer a simple example for phase space which describes the movement of a local particle under the influence of a simple force. One of the most important and fundamental traits of the solutions of a chaotic system, meaning the paths which the solutions draw in the phase space, is that they usually end in “strange attractors,” these being the paths that the solutions are attracted to from a number of different starting points. In the end, they all reach the same area and act in a very similar manner (see the image below). As we already noted above, the solutions at the later times do not become very distant from one another, but what characterizes a chaotic system is the lack of a connection between what happens at the beginning and what happens at later times. The two paths which begin at points close to each other at the beginning of the process may reach very different points at the end, so that it’s very hard to predict the end of the path based on knowledge of its beginning.

Lorenz Attractor. — This picture is one of the most well-known in chaos history. It describes the paths of the solutions for the Lorenz equations to the phase space of the climate system, where pressure and temperature are described at every point in time. We can see that the various paths act in a very similar manner — they all create a butterfly form (unrelated to the butterfly effect, of course).

Another way to deal with chaotic phenomena is use of tools of probability. If the system is acting in an erratic and unpredictable manner, what can be done is to search for its statistical traits. For instance, where is the average location? Or, what will be the location area if the starting point is in a particular area of the phase space, belonging to a range of initial locations and speeds? Much like the example of the throwing of the piece of paper, there is no point to try to predict where each such piece will fall. It is hopeless, as there is no simple connection between its final location and its starting point, as opposed to the example of the rock. But we can perhaps try and characterize the distribution of the locations of many pieces of paper which will be thrown downward — how many will fall in area X and how many in area Y and so on. Thus, within a system which appears physical and rigid, and of course also entirely causal, we have a distribution of results, as though it were a random process. This is a way to look at the problem in predicting chaotic systems from a different angle and see the degree of similarity between chaos theory and randomness.

Between the Chaotic gap in Physicalism and Freedom The first to understand chaos to be a new and unique field in physics and mathematics was a group of students from the University of Santa Cruz. Doyne Farmer, one of the first chaos theorists, a member of the Santa Cruz group, described his feelings when he and his friends started to understand that they had a theory with far-reaching significance for physics in their hands:117 On a philosophical level, it struck me as an operational way to define free will in a way that allowed you to reconcile free will with determinism. The system is deterministic, but you can’t tell what it’s going to do next … Here was one coin with two sides. Here was order, with randomness emerging, and then one step further away was randomness with its own underlying order. Farmer claimed that this gap in physicalism, which allows us to find randomness and phenomena which cannot be predicted within rigid physics, allows us an opening into understanding the phenomenon of free will. Ostensibly, this is precisely the gap we were looking for in physicalism. After all, we have here cases in which the present situation does not absolutely determine what will happen in the future; we can never know the future, at least so long as we are not equipped with utopian measuring tools providing perfectly precise data. As we said more than once, freedom of will is also expressed in our inability to predict human decisions. It is thus very tempting to introduce free will into this gap, in a manner which does not harm the principle of causality (we are also thus exempt from making use of the Lex Specialis rule we described in chapter six). But despite the excitement and fascinating options which chaos ostensibly present us, the connection Farmer made between randomness and free will is problematic, to say the least. As we’ve already shown, freedom of choice is not randomness but another sort of mechanism. Already at the beginning of chapter two we saw that throwing a die or tossing a coin is a phenomenon which contains nothing random, yet we still address it with tools of probability and statistics. Moreover, data from chaotic systems are not even really random; As we have seen, they only have characteristics similar to random data. The distinction between random data and chaotic data is a lively and fairly complex field of research, but it is clear to all those who study it that they are two different kinds of phenomena. It may be that Farmer was not talking about actual randomness. He may have simply meant to say that because of the inability to predict, we cannot speak of causality, and there is therefore room for free will even within a physicalist picture. But even this lenient interpretation reveals an error, if not a misunderstanding of the philosophical issue of freedom of will. When he said these things, Farmer was just twenty four years old and starting out. We can therefore forgive such blatant philosophical errors. But it’s harder to forgive more mature and experienced philosophers, physicists, and mathematicians, who keep repeating the same mistake, until today. So where’s the mistake? The basic mistake is conceptual. From chapter one, we pointed to the lack of predictability not being equivalent to the lack of causality, just as predictability is not equivalent to causality. We also saw that if there is a synchronous relationship between two systems (this synchronicity including the logical and temporal component of causality, but not the physical component), a look at one provides me with the possibility of predicting the conduct of the other. Prediction is possible here even though there is no causal relationship between them (see the astrology example in chapter five). At the same time, the fact that I cannot predict something primarily says something about me (see the first quote at the beginning of the chapter), but there is no justification in concluding that prediction is impossible in principle. Moreover, we also saw in chapter seven that even where there is a data series which cannot be predicted, this does not mean it is not modelable; we may only be dealing with a lack of information and not an inability to determine its pattern in principle. This is precisely the situation when it comes to chaos, as what I lack is only information (on the initial values — if I knew them precisely, I could predict the future in full). More concretely, there is a misunderstanding of the concept of free will and the libertarian picture as a whole. Say we proved that human beings act chaotically, and that given the details of his present situation at some level of precision, we have no practical way of predicting his actions in the future. Does that mean that his actions are not determined by causal factors or by his present state? Of course not. After all, even the fall of a piece of paper is an entirely deterministic process. It is controlled by the law of gravity and Newtonian mechanics, taking into consideration winds, air density, initial speed, the shape of the piece of paper, and so on. In other words, the ultimate location of the piece of paper on the ground is unequivocally

determined by the equations of motion based on initial data. The fact that we cannot predict its final location does not point to the indeterminacy of the physical process itself; the physical data entirely determine the final resting place of the piece of paper on the ground. It’s just that this determination involves a very complex calculation, and it is therefore difficult for us to discover it. Is anyone willing to say that paper, or the wind that moves it, has free will? Libertarianism makes an argument which is much more far-reaching than that of unpredictability: Human behavior is not at all determined by circumstances. This is not a problem of the observer who cannot predict the behavior because of its complexity or for some external reason, but a function of the observed system itself. The libertarian doesn’t argue that I don’t know how to predict, or that it’s hard to predict, but that it’s impossible to predict, because the future is not at all determined in terms of the present. In other words, given a detailed state like the will of a human soul and its environment, he still has a number of paths before him in terms of action, as opposed to the piece of paper. In this libertarian picture, not only is there no explicit expression of the human being’s course of action, but we do not have any equations which define it (let us be reminded that it is the equations and not the solutions which define causal determinism). To sum up, in a chaotic system, no matter how wild, given certain initial circumstances, the result is unique. It will be one and no other. The present entirely determines the future, such that we have no possibility of getting two different results for the same initial conditions. But due to the inaccuracy of our tools and difficulties in calculation, we cannot know it, but the more precisely we measure the starting conditions, we will know the future with greater accuracy for a longer period of time. The natural conclusion is that a chaotic system is an entirely deterministic one, as it is entirely causally determined. In other words, chaos and free will are unrelated. Chaos is not a true gap in physicalism, but a space created in our knowledge because of our limitations. Free behavior, by contrast, is behavior which cannot be predicted even by an omniscient factor, with infinite capabilities and complete figures regarding the future (we saw this above at the end of part one of chapter four even regarding God Himself). This inability is not due to the observing person but the observed system itself. To sharpen the point even further: Is someone willing to impose moral responsibility on a piece of paper, or the wind, just because their behavior can’t be predicted? That is the meaning of the libertarian approach, and it is clear that chaos, with all its idiosyncrasies, cannot serve as a basis for it. One last example will get this point across. A local particle is on the top of a hill with a round peak, precisely at the uppermost point, like in the following drawing:

This is known in physics as unstable equilibrium. In principle, so long as no force acts upon it, the particle will continue to stay there (theoretically, if it is local and is precisely at the mathematically highest point of the hill), but any force, no matter how small, will knock it down one of the sides. There is no possibility to know where it will fall in advance, as any gentle wind can move it, and we do not know the weather conditions at that resolution (remember, Lorenz dealt with weather and climate problems, and he found that they are chaotic by nature). Such a situation is chaotic in the essential sense, as there is a strong dependence here on the starting conditions, such that a tiny difference in force operating on the body leads to an entirely different result. This is therefore a situation where we have no ability to predict, because we cannot know where the particle will fall just by looking at it. But this is only due to the fact that we do not know the wind directions in the surrounding environment and their force. Does that mean that the particle’s movement is not determined by the circumstances? Of course it is. What knocks it down is a physical force which is applied

by the wind, which also unequivocally determines where and how it will fall, and without physical force, it will remain there forever. It’s true that we can’t know what force will be applied and where the particle will fall, but as stated above this is not a fundamental gap in physics, but at most a gap in our knowledge. Can we introduce free will into such an arena in this situation? Does the chaotic nature of the situation allow us to say that the particle fell to the right because it “decided” to, making a willful decision lacking a force acting upon it? This is what it would mean to introduce freedom of will into physics via chaos theory. If the particle fell to the right because it decided to, that means that this fall took place without a force which caused it, entirely contradicting the laws of physics, as we already saw. This means that, as already mentioned, chaos does not really create a significant space within physics, and certainly not one into which we can allow free will to be introduced into the framework of physics. Our interim conclusion is that even after chaos theory, we are left with physical causality without any gaps. Therefore, at least in terms of chaos theory, free will can still not dwell together with physicalism. Libertarianism will not be built on this basis.

Probability and Randomness: Between Epistemology and Ontology So, we rejected the proposal to use chaos to introduce freedom of will into the physicalist picture, but this discussion of chaos was not fruitless; it helped put the problem into greater focus. We saw that libertarianism espouses ontological indeterminacy, an argument stating that the system in and of itself is not determined by present conditions. In the libertarian context, the problem of prediction, which indeed exists regarding free human behavior, is a result of the system itself having the trait of indeterminacy and not just a lack of information on the side of the observer. We can determine it as ontological indeterminacy, meaning the indeterminacy is of reality itself: Given the conditions of the present, there are still a number of possible futures. By contrast, chaos offers us a mechanism of epistemological indeterminacy (epistemology being the theory of knowledge and justified belief), meaning that the system itself is entirely fixed; it’s just that a person cannot recognize this or arrive at this information. There are no possible futures here; it is only the observing individual who thinks that a number of different futures are possible. The heart of the problem, then, is that there was no ontological gap in any of the laws of physics we have described so far, including as regards chaos theory. They were rigidly causal, and the present circumstances entirely determine the future results. We also saw that chaos did not change this, as it at most offered an epistemological gap. Now we must ask the question: Is this really characteristic of all the physics we know of? Is it true to say that there is no ontological gap in our physics? Are all the gaps in physics solely the result of information gaps by us and not in reality itself? Over the course of the twentieth century, it turned out that there is indeed an ontological gap in the physical world itself, as well, and it can be found in quantum theory, which we will deal with in the next chapter. 112 An extensive and still accessible introduction to chaos is James Gleick, Chaos, Penguin, 2008. 113 Important French thinker and scientist who worked at the end of the eighteenth century and the beginning of the nineteenth, 114 Historically, quantum theory formed before chaos theory (although its roots precede it), but methodologically, the order of discussion in the following two chapters will be the opposite. The reason is that chaos belongs to the field of classical physics; it was a late discovery which illuminated causality in classical physics itself in a slightly different light. Quantum theory, by contrast, has already replaced classic physics, and the conceptual revolution in its wake (alongside the theory of relativity which came a little beforehand) is the new physics. This is therefore a later stage of the discussion, at least methodologically. 115 Poincare is described by many mathematicians as the last of the universalists, since he was the last to make real contributions in all branches of mathematics (as well as a number of fields in physics). He did all this alongside his lifelong work as a mining engineer and supervisor of mining work. Poincare was also a philosopher who espoused the use of intuition and had a unique perspective in the philosophy of mathematics, according to which mathematics is not analytical. It’s worth noting that Poincare also intervened at least twice in the Dreyfus Affair, and came out against the absurd scientific claims made there. 116 This is just an example demonstrating the essence of chaos in an accessible manner to those who aren’t familiar with it, and therefore conceptual and terminological accuracy is less important here. If we use a more precise formulation, the throwing of the paper is not in itself a chaotic phenomenon, as the equations which control it are fairly simple and one-directional (the air affects the paper but the paper has almost no reciprocal effect on the air). What is chaotic here is the weather; the winds, air density, the temperature — all these and more are traits of the weather affecting the paper on its way down. We have already mentioned that weather phenomena tend to be chaotic. 117 Gleik, Chaos.

Chapter Ten GAPS IN PHYSICALISM: QUANTUM THEORY Of all the theories presented in the twentieth century, quantum theory is the stupidest. Some say that the only point in its favor is that it is undeniably true. - Michio Kaku, Hyperspace In the previous chapter, we pointed to the fact that chaos theory does not open gaps in physicalism. The gap is only by us and our ability to predict a future event, not the physical world itself. By contrast, quantum theory definitely does so. Quantum theory is therefore the last hope of the materialist to develop a libertarian approach within physicalism. It’s important for me to repeat an important point I’ve made before: Ostensibly, the closing of all the physical gaps is a death blow to libertarianism, as it sharpens and amplifies the contradiction between it and the principle of causality and physicalism. But this is a mistake. I aim to reject the possibility of a libertarian picture based on materialism and show that it is impossible to think of a world which is entirely made up of physics but which also contains free will. However, this does not mean that libertarianism is untrue, just that it must be established on a dualist basis. This will be the conclusion of the current chapter, but we’re going to have to sweat a bit before we get there.

Prelude: The Strangeness of Quantum Theory Sir Arthur Eddington of Cambridge, seen by some as the first astrophysicist, was among the first to espouse the theory of relativity and one of its greatest advocates and exponents. He was once told that there are only three people in the world who understand the theory of relativity, to which he immediately replied: “Who is the third?” (Aside from him and Einstein, of course). A few decades later another physicist, Richard Feynman,118 was told that no-one seems to understand Einstein’s theory of relativity, and he is said to have responded roughly as follows: “Quite a few people understand the theory of relativity, but I don’t know anyone who understands quantum theory,” adding: “You don’t understand quantum theory, at most you get used to it.” This is what Michio Kaku appears to have meant in the quote opening this chapter. What is quantum theory? What is so strange and embarrassing, or even stupid, about it? Is it nevertheless still true in spite of all that? Why, if at all, does it relate to determinism? I will now try to briefly and simply answer these questions as much as possible. The description I will provide here for quantum theory is far from being exhaustive even when it comes to the basics (primarily since it is a very mathematical theory, and due to the limitations of this medium, I will entirely avoid mathematics). Moreover, I am assuming certain interpretations of it (usually the accepted ones), because the other interpretations, to the best of my understanding, do not change the primary conclusions which will arise from the discussion here, and indeed may even strengthen them.

The Double Slit Experiment As Richard Feynman liked to say, the best way to explain quantum theory is via the results of the double slit experiment, which has meanwhile become a classic experiment in the history of science. The most wellknown experiment was conducted in the twentieth century in reference to electrons, and it had revolutionary consequences which led to quantum theory. But its origins lie as an experiment at the beginning of the nineteenth century in the field of optics. We will therefore begin our description from there. Since Newton, a debate had taken place between physicists regarding the nature of light. Some, like Newton, argued that it is composed of particles, and some saw it as a wave (the Huygens–Fresnel theory). In 1801, 74 years after Newton’s death, Thomas Young conducted the first double slit experiment to decide this issue. In the experiment setup Young used there is a clear difference between a wave and a ray of particles, and precisely because of this, it was useful in deciding the debate on light. This is precisely the reason that it will be considered incredibly useful in the next century regarding a similar debate, and this

time regarding the nature of particles such as electrons (there, too, the question is whether they have a wave-like nature or a particle-like nature). To understand this, we will first examine the simpler case involving a single slit. The image below describes an origin (the cylinder on the right), which sends a ray of particles or a wave (the thick dotted line) in the direction of a partition which contains a slit. Behind the partition is a screen (described by a double line) which is sensitive to being hit by a wave or particles.

Figure 1: The Experiment Setup of the Single Slit

The graph on the left describes the experiment in the case of the single slit. Every point on the vertical axis of the graph represents the parallel point on the screen. The height of the point on the horizontal axis represents the amount of particles or the strength of the wave resulting at that point on the screen. We can see that the largest amount results precisely in front of the slit, with the amount decreasing on both sides of it. So far, this is all true for both the beam of particles and the wave beam. If there is a small difference, it is at the margins. In the case of the wave, the strength at the sides is created by the deviation of the rays which did not arrive at a straight angle, as well as an optical phenomenon called diffraction, according to which the light changes its direction of advance a little around the edge of the partition, meaning it deviates from the straight line. In the experiment of the particle beam, the strengths at the sides are formed from the particles which did not arrive at a straight angle, and also from brushing up against the edge of the slit which diverts the particles which hit it. We can see these two as parallel phenomena, and it will therefore be hard to distinguish a particle beam and a wave in a single slit experiment. How, then, can we nevertheless distinguish between them? Young proposed a double slit experiment. To understand his proposal, we must become familiar with another wave phenomenon: Interference. When two beams of particles move in space, the total number of particles at any point is the sum of the numbers of particles from both beams at that point. By contrast, when two light waves move in space, the overall power of the light at any point is not a simple sum of the power of the two waves at the point under discussion. The power at every point in space will go up or down as a result of the interference effect between the waves. In certain places, the power of the two waves cancel each other out, and in other places they strengthen each other (this is dependent on what is called the “phase difference” between them). Young’s experiment setup was based on this phenomenon, and it is different than the figure above only in that the partition now contains two slits, and not just one. Now we must discuss the two cases (waves and particles) separately. If the source emits a beam of particles, like electrons, tennis balls, or elephants, some of them will go through each of the two slits. The amount of particles around every slit will be distributed like in the graph of the single slit (see the graph in figure 1). Thus, in the double slit experiment, the picture on the screen will appear like the sum of pictures of the individual slits, meaning we will get a graph with two equal peaks, as described by figure 2:

Figure 2: The Double Slit Experiment With Beam Of Particles

What will happen when the source sends a wave beam? In this case, the phenomenon of interference will cause the picture to look entirely different, as can be seen in figure 3:119

Figure 3: The Double Slit Experiment With Wave Beam

In the case of the wave described in figure 3, the maximal strength appears in the center, between the two slits (where the particle picture almost zeroes out; see figure 2). On both sides of the center there are side nodes whose peaks steadily decline. Thus, the double slit experiment provides us with a sharp distinction between a particle beam and a wave beam. At the beginning of the nineteenth century, Young conducted the double slit experiment in reference to the wave of light, and the result he got was unequivocal: A wave picture emerged (figure 3). Thus was the debate between the now-dead Newton and Huygens decided, with Newton being the loser — light was a wave phenomenon. But victory over Sir Isaac Newton, one of the greatest physicists and mathematicians of all time, is not such a simple task. As we will see later on, the future had a resurgence in store for him. More than a hundred years later, from the beginning of the twentieth century, evidence and arguments

began to accumulate in favor of the picture which sees particles as waves. In 1924, a French prince named Louis de Broglie (d. 1987), in his doctorate, proposes to see the electron as a wave, and even measures the characteristics of this wave, such as its wavelength. Now the opposite debate took place: Does what we are used to seeing as a particle — say, an electron — actually have a wavy nature? To examine this question, double slit experiments were once again conducted, but this time with an electron beam. The expectation was that the picture would be like figure 2, as it was clear that electrons are particles. The results were stunning. It turned out that the emerging picture actually looked like figure 3. It therefore turned out that electrons actually have a wavy nature. Electrons thus act like light waves, thus ostensibly erasing the difference between particles and waves. But a whole series of facts stood opposite these results which pointed to electrons being particles. It ostensibly appeared that electrons are both particles and waves. It was not clear how particle beams could interfere with themselves; interference, after all, is a characteristic of waves and not of particles. A particle is located at a particular place, and a combination of two particles in the same place, as would be necessary for them to interfere with each other, simply cannot be possible. At a certain stage, the hypothesis was raised that interference is the result of the fact that we are dealing with a beam of electrons and not a single electron. In other words, different electrons disrupt each other, and the encounter between them somehow creates a picture of interference. Since physics is an empirical science, the ultimate test is that of the experiment. So, to test this proposal, the double slit experiment was repeated with an electron beam being fired from the source at a very slow, diminished rate. In such a situation, only a single electron hit the screen each time, and there was no additional electron to hit it at the same time, ensuring that interference between the two electrons would not happen. Surprisingly enough, it turned out that the picture is still a wavy one (like figure 3). The amazing conclusion was that the electron is not a tiny particle at a particular place (a sort of tiny ball), as was thought until then, but rather a wave which is present throughout the space, just like light (we of course still remember Fresnel-Huygens and Young). But the situation is even more complicated. If the electron does indeed interfere with itself, this means that the single electron is actually a wave which moves through both slits. Thus, the two parts of the electron wave (the wave of the single electron) reach the screen from different places and create a picture of interference (figure 3). This was already intolerable. How can it be that one particle moves through two slits at the same time? To examine this question, we must see which slit our electron is going through, or whether perhaps it is going through both. So, back to the lab. The experiment was conducted again, but this time a detector was placed next to one of the slits, which we will call slit A. When the electron goes through the slit, the detector picks it up and reports that the electron passed through slit A. If the detector reports nothing, that means it passed through slit B. In the double slit experiment with the presence of the detector, the amazement became total. The detectors did show which slit the electron passed through, but the screen also showed the particle-like picture, meaning the graph in figure 2. The phenomenon of the particle interfering with itself (figure 3) disappeared. This means that when there is a detector checking it, the electron goes back to behaving like a civilized electron (like a tiny tennis ball), but when no-one looks at it, it goes wild and allows itself to act like a wave.120 In the language which developed afterward, with the formation of quantum theory, the conclusion was that the electron is a quantum particle.

Einstein as Necromancer, or: Newton Comes Back From the Dead This initial description of the birth of quantum theory lacks another point — another thought revolution from the school of twentieth century physicists. In 1900, right at the same time as Poincare was laying the foundations for chaos theory, Max Planck unified the information of the radiation of a black body, explaining it in terms of isolated amounts of energy. Five years later, Albert Einstein became world famous and his name became synonymous with genius — the year of 1905 was Einstein’s year of wonders, during which he published three different articles: One discovered Brownian Movement, which is an important branch of statistical mechanics, another reveals and formulates the special theory of relativity, and a third takes Planck’s findings on black body radiation and uses them to form the foundations of quantum theory, which he would later sharply oppose with all his strength for the rest of his life, on the grounds that “God doesn’t play dice with the universe.” In this article, Einstein defines the particle traits of light, arguing that light is also made up of particles called photons. Old Newton started to roll in his grave. The continuation of the story is no less surprising than its beginning. The double slit experiment was

conducted again with a beam of light, just like Young did, but this time with a very weak ray. The results were downright inconceivable this time. In the normal experiment, the distribution of strength on the screen looked exactly like the case of the electrons without the detector next to the slit, producing the image seen in figure 3. The conclusion meantime was that light is a wave (the image of Fresnel-Huygens), just as we thought. Newton went back to his normal state of matter. But that is not the end of the story. When a light detector was placed next to one of the slits, it suddenly began to behave like a particle, resulting in the picture of figure 2 (once again producing Newton’s image, who was now rubbing his eyes and stretching). The conclusion was that light was also a collection of particles which sometimes behave as waves (if there are no detectors looking at them), just like electrons. Thus, two hundred years after his death, and a hundred years after the young experiment, Newton was back in the picture. In the end, it turned out that both he and Fresnel-Huygens were right. But like the wellknown tale of the judge who said that both sides’ argument were right, and when his wife asked how could that be, he answered: “You’re also right” — those who wonder how two contrary arguments can be right at the same time — are also right. The verdict was rendered in the lab, and the scientific community was bewildered. The common interpretation which later developed for quantum theory (known as the “Copenhagen interpretation,” in the wake of Danish physicist Niels Bohr and his school) argued that the significance of these findings is that the distinction between particles and waves has lost its sharp ontological meaning. Reality itself contains no two such creations. The microscopic physical entities sometimes behave one way and sometimes another. This distinction is drawn from our macroscopic world, but it does not really exist in the microscopic level.

Quantum Theory: A Picture of Superposition The conclusion which physicists arrived at based on this confusing picture is that so long as it is not looked at (or detected by a device), the electron is in a wave state. This situation is called superposition since it is the sum of pure particle states. A particle state is one in which the electron acts like a tiny tennis ball. It has a defined place in space at any given moment, and if we draw a graph of its location as a function of time, we will get a defined path in space, like in the case of the tennis ball. In a pure particle state — the path of the electron goes through slit A, meaning that it moves along the path like a tennis ball from the source to the screen. But such a pure particle state emerges only when we measure the location of the electron with a detector: If the detector shows us that the electron passed through slit A, we get the first pure particle state, and if the detector shows that the electron passed through slit B, then we get the second pure particle state. So what happens if we didn’t measure the location of the electron at all (i.e., there were no detectors near the slits)? In such a case, the electron is in a superposition: Its path is the sum of paths A and B, or the sum of the two particle states. In everyday terms, this means that the electron goes through A and also through B, and effectively interferes with itself like a wave. The electron ostensibly splits and each part of it takes a different path. So does our electron take the path which goes through A and also the path which goes through B, as well as many other paths, and is therefore effectively throughout all of the space. Much like a wave, or as Israeli poet Nathan Yonathan put it: “Just like a beach.” So, when we do not measure the location of an electron, instead of talking about it like a tiny particle or tiny tennis ball, we need to talk about it like a wave. Here we are not talking about a path which defines a pinpoint location at any given time, but a phenomenon which at any given time exists and is spread out throughout all of space, like a light or sea wave. Newtonian mechanics described a particle which moved under the influence of forces. At any given moment it has a speed and a location. Newtonian theory speaks of the location of the electron (x) as a function of time (t), and its path is x(t). Quantum theory, by contrast, describes an entirely different picture. Quantum mechanics which describes the electron does not refer to a concept like a particle or its location. Erwin Schrödinger formulated quantum mechanics in terms of a wave equation, whose variable is an entire function which is defined across the entire space at every moment — what quantum theory calls a wave function of a particle. Schrödinger’s picture contains no particle and there is no specific location of a particle. Instead, there is a function which assigns some (complex) value to every point in the space, and it is usually marked as . Every location x and every moment t has receiving a particular value. The accepted interpretation is that this value describes the probability of finding the particle at location x in time t. This probability is given by squaring the absolute value of his wave function at that location: | |2 In this picture, the location of the electron is not defined in a singular manner, meaning that we cannot

say that the electron is located in place x at moment t. All we can say is that at moment t, there is some chance to find the electron at place x, and the value of this chance is the squaring of the absolute value of the wave function at this time and place. But at the same time, there is a chance to find the electron in many, many other places (at every time t there is a value of the function for every x in space). As we noted above, such a wave state is the sum of the pure particle states. This wave function can therefore be written as the sum of pure particle states, these being cases where the electron has a single, defined location at any given time. Every such situation describes the electron as a particle, but the complete picture is of a wave being built from the sum of all possible particle pictures, with every one of them having odds attached to them. This is what physicists call a state of superposition. In such a state, its being a wave or a particle is determined by the character of the wave function at that moment: When we see an electron acting like a particle (in a situation where there is a detector next to one of the slits), this means that its wave function is one pure particle state (e.g., the path going through slit A), without an addition of other pure states. In such a case, the odds of getting location X of the path of the particle are 1, and in other places, 0. By contrast, when we describe it as a wave, this means that its wave function is not composed of a single path, and a measurement of the location of the electron can produce quite a number of results which will be very different from each other.

Is Quantum Theory Causal? The Schrödinger equation is a differential one, in which the value of the quantity being described (the wave function) at any given moment is precisely defined by its values at the previous moment. It is therefore clear that the dynamic of the wave function is an entirely causal dynamic (see also the previous chapter on this). The data at a given time (the starting conditions) unequivocally determine all the values of the function at any future time. So why do people say that quantum theory does harm to the classic concepts of causality? This statement does not refer to the wave function, but to measurable quantities (or variables), such as the location of the electron or its velocity (its momentum). As we saw, these are not unequivocally fixed. If an electron starts at time t0 in a particular state in which its wave is some function , and even if this situation is a pure particle state, the dynamic of the Schrödinger equation will “disperse” this picture, and we will get a wave function in later times which describes the electron starting to be composed of a superposition which includes additional pure particle states. The particle slowly disperses and becomes a wave. This means that at later times, its location is no longer determined by the starting conditions and there is a chance to find it at other places. For instance, in the double slit experiment, even if the electron leaves the source with entirely particlelike attributes, the measurement values during its movement are not unequivocally determined by these initial circumstances. There is a chance to find it going through slit A but there is also a chance to find it going through slit B, and there is also a chance it will go through both slits at once (when it is in a superposition which includes the two pure paths). It’s important to stress again that the chance we are speaking of is not in the same sense as in the case of chaos theory. Here, the location of the particle is not at all set, while there the location is set but the observer cannot always know it. In other words, here the gap is ontological, and not epistemological like with chaos. We will expand on this point later in the chapter. In other words, the past does not determine the future in quantum theory, except in a probabilistic manner. The classic deterministic-causal picture shatters, to the great joy of the libertarian. There are other phenomena in quantum theory, which also appear arbitrary, as though they occur entirely without a cause: The emission of a photon from an atom, radioactive decay (a radioactive atom which emits radiation), the formation of pairs of particles from a vacuum, and so forth. All these are events which occur in a manner which appears entirely arbitrary, without cause (however, as we explained in chapter five, with reason — quantum theory is “sufficient reason” for these events) and even without any attempt to predict time or the location in which they take place. All these are not essentially different than what we have described so far; all these cases involve a wave function which describes the situation, and it provides the odds of the event at any time or place we wish. Thus, we see that we have found the desired freedom within the laws of physics. We saw that the meaning of freedom of will is that a person in any given situation can act in a number of different ways, meaning that the given state does not unequivocally determine his further action. In quantum theory, we get a very similar situation: Here, too, the situation at a given moment does not unequivocally determine future values. From one starting state of the particle, when we wait for some time and once again measure its location, or speed, we can get a number of different values of location or speed in our measurement. We have no way of predicting the speed or the location of the particle at future moments based on the present.

Measurement and Collapse It’s important to stress that every time we measure the location of this electron we will get one defined location. No measurement will provide us with a number of different locations for the same particle. The innovation of quantum theory is that even if the location of the particle at a given moment is a given, it is still not clear in advance which of these locations will show up in the measurement we conduct later. Such a measurement can result in one of a number of different results for its location, and the odds of getting such a location are determined by the aforementioned values of the wave function. Another interesting conclusion from the double slit experiment is that if we measured the location of the particle and we found that it is located in a particular place, then it has entered a pure particle state in which it has a perfectly defined location at any given point. Therefore, when we placed a detector next to one of the slits, and we discovered that the particle goes through it, or through the other slit, then there is no longer a picture of interference. In such a state, a measurement was conducted for the location of the particle and it was found that the particle passed through only one of the slits, and not both of them at once. Thus, according to the common interpretation of quantum theory, the measurement affects the state of the particle. If we measured its location, we got some result in that measurement (that it passes through A or B), which in itself is not determined in advance. But after the measurement, it turned out that the state of the particle is a pure particle state (as we measured a single, defined location). Therefore the particle changes states during the measurement, and instead of being a wave going through both slits at once, it suddenly becomes a particle which goes through only one of them. Quantum theory states at the same time that the measurement of speed can transfer it from a particle state to a wave state. Before the measurement, the wave function is composed of a superposition of pure states where measurement can provide one of them in the form of odds, and after the measurement, the particle collapses to one of them; it transforms from the complete superposition to one of the pure states making up this combination.121

The Meaning of Superposition: The Difference Between Epistemology and Ontology We saw that so long as we didn’t measure the location of the particle, it is in a superposition of a number of location states. In any pure state of location — the particle is at a particular state in the space at any point in time (meaning that this represents a normal path of a pinpoint particle; a state of a tennis ball). But our real particle is in a state which is a combination of a number of such pure states (superposition). The meaning of such a state is that until the moment of measurement, the particle has no defined location. Only when we measure its location, we will get one defined result. We saw above that this is not necessarily one particular location determined by the starting conditions. We are dealing with a number of possible results for this measurement, and the odds of getting it at particular location x are determined by the value of the wave function at that place. The accepted interpretation of quantum theory is that this wave function does not represent our lack of knowledge of reality. When the wave function describes the possibility of getting a number of different location values, this does not just indicate a lack of information regarding the particle on the part of the physicist, but rather ontological uncertainty, meaning uncertainty characterizing the particle itself. There are hypotheses and proposals regarding hidden variables which determine these results, interpretations which try to leave these gaps in the epistemological realm, and if they are correct then quantum theory has nothing strange about it in this respect. There may be no unequivocal proof which refutes all these proposals, but a large portion of them has been refuted, and the common approach among physicists is that there are no such hidden variables. Here we will also assume that no such variables exist, not just because most physicists believe so, but because if such variables exist, then our discussion of quantum theory becomes entirely irrelevant; if hidden variables control the results of a future measurement of the location of the particle — this means that quantum processes are entirely deterministic, and in such a case, quantum theory also fails to provide a place for freedom within the physicalist picture, thus ending our interest in the matter. We will therefore continue the discussion on the “better” assumption for the physicalist libertarian, according to which there are not hidden variables, and that quantum theory does indeed rip open an ontological gap in physical causality. What is the meaning of the above statement, according to which the wave function does not describe our lack of knowledge about the particle, but rather freedom in the measurable values of the particle itself? This is the heart of our discussion. Here, quantum theory deviates entirely from other fields and laws in physics. To understand this, we will once again return to the question of epistemology and ontology.

There are also situations in classic physics in which we use probabilistic and statistical tools to describe a system of which we lack knowledge. For instance, the area known as statistical mechanics, which is the basis for thermodynamics (or heat theory), describes a state of a multi-particle system like gas in a container. Gas is nothing more than an accumulation of particles, every one of which has a location and speed at every given moment. We are talking about an enormous number of particles, and there is therefore no chance that we will possess complete information on this system in the form of knowing the speed and location of each particle. The solution is to handle this system with probabilistic tools. We describe how many of the particles are at every speed or location (or energy), and this turns out to be sufficient to describe the global characteristics of the system in the form of temperature, pressure, volume, and so on.122 It turns out that we also have a function in this case which describes the distribution of states and speeds and the like. Here, too, we do not know the locations and speeds of every particle, and we only have statistical knowledge on the collection of particles. Here, too, there is a function of distribution which describes the odds of finding a particle at any given speed or location. This is the same when it comes to tossing a die, whose results we describe in terms of probability, since we do not know what the result will be. We saw this in the previous chapter regarding the totality of chaotic processes, such as tossing a piece of paper from the top of a building. What, then, is the difference between such a situation and quantum uncertainty? Why is the distribution described by the wave function in quantum theory essentially different than the function which describes the distribution of the speeds and locations in statistical mechanics, for instance? The difference between these two situations brings us back to the distinction between epistemology and ontology. In classical physics, the use of probabilistic tools does not derive from randomness which exists in the system itself, but rather our lack of knowledge regarding that system, precisely as we saw regarding the tossing of a die or a coin, the throwing of a piece of paper from the top of a building or the rolling of a particle from the top of a round hill). In the quantum system, by contrast, the particle itself has no defined speed or location. It itself is located in a number of locations or at a number of speed values at the same time. This means that the wave function does not describe our lack of information, in which we simply do not know the precise location or speed of the particle. The argument is instead that the particle itself does not have one defined location, only a distribution of odds to be in different locations. We can put it differently like so: We cannot define a function of distribution of classical statistical mechanics on a collection of one electron. The single electron has a particular location or speed, which we may not know. Only in the case of a collection of electrons can we speak of a distribution which determines how many of the particles will be in any location or at any speed, and so on. But in quantum mechanics, we can also do this with a single electron. The single particle also has no specific location or speed, but only various possibilities at different odds. We see this when a picture of interference emerges in the double slit experiment even when conducted on a single electron, as we saw above. Such a think simply cannot be in classical physics, as the single electron has a single path, even if it is not known to us. A classical particle cannot interfere with itself. Thus, in classical statistical mechanics the distribution describes the gaps in our information on factual reality, not reality itself. This is an epistemological gap. By contrast, the quantum distribution describes the factual state in the world itself, meaning that this is an ontological gap within physics; physics itself is not defined all the way. Quantum theory is thus different than chaos theory which we described in the previous chapter. In quantum theory, the state of the system itself is not at all determined by the starting conditions. In effect, it can reach a number of different values of measurable quantities at future times from precisely the same starting conditions, something which cannot happen in the case of chaos.

Fifth Intermezzo: Ontological and Epistemological Doubts Let us sharpen the distinction between ontological and epistemological indeterminacy with a Talmudic example. Despite the bewilderment and misunderstanding caused by the quantum description, it turns out that the commentators of the Talmud use this same distinction. They do so in the legal sphere and not the physical sphere, of course, and as we will now see, it is much easier to understand and sense this confusing distinction when it comes to law. This is the reason for this intermezzo, as aside from additional demonstration, it may aid the bewildered reader to understand it in the physical context as well. Halacha deals a lot with situations of doubt. For instance, if I found a piece of meat tossed on the floor in the center of town, there is room for doubt as to whether it is kosher meat or not. This is a state of doubt, and it is decided according to rules for deciding doubts. One of the indications which may help in deciding

the doubt is a survey of all the stores selling meat in town. If most of them sell kosher meat, then I can eat the meat I found with no legal problem. This is an example of one of the rules for deciding doubts: The rule of following the majority. In the Mishna (Bavli, Yevamot 118b), we find another case of doubt: If a man betrothed one of five women and he does not know which of them he has betrothed, and each states, ‘he has betrothed me. He gives a letter of divorce to every one of them, and, leaving the kethubah123 among them, withdraws; so R. Tarfon. R. Akiva, however, said: this is not a way that would take one out of the power of transgression, unless one gives to each of them both a letter of divorce and her kethubah. If a man robbed one of five persons and does not know which of them he has robbed, and each one states, ‘He has robbed me,’ he leaves the [amount of] the robbery among them and withdraws; so R. Tarfon. R. Akiva, however, stated: This is not a way that would lead one out of the power of sin, unless one pays [the full amount of the robbery] to every one [of the persons involved]. We have a situation where a man betrothed a particular woman and does not remember which of the five he betrothed. Such a situation can emerge when a man sends a messenger to choose a wife for his son from a particular town, like the Biblical case of Eliezer, slave of Abraham whose master asked him to seek out a wife for his son Isaac in Haran, and the messenger betrothed a woman and died. Now the groom-tobe doesn’t know who his wife is. In such a case, all these women are considered to possibly be his wife or there is a doubt that they are his wife. The second case in this Mishna deals with someone who stole but does not remember from whom. The situation here is very similar to the first case. The Tanaaim, the Jewish commentators of the Mishnaic era, were divided on the question of whether he can place the object among the claimants and leave, or whether he must return the object to all of them to “play it safe.” All these situations are similar; the factual case is singular and fixed — it’s just that we don’t know what it is. Either the piece of meat is kosher or not, the woman is betrothed or not, the object belongs to so-and-so or not. The lack of knowledge only exists regarding the individual and not the world itself. This, then, is an epistemological doubt and not an ontological one. By contrast in the case of “betrothal not given to consummation” (Kiddushin 51a-b), a discussion is held regarding a person who came to a father and gave him money to betroth one of his two daughters, without specifying which one. Ostensibly, this is also a normal case of doubt, as it is not known which of them was betrothed to him. But the discussion in the Talmud in this case leads to our attitude here being different than the two aforementioned cases. Due to the complexity of the discussion we will not go into details, sufficing here with the fact of the Talmud itself distinguishing between the two cases. I will leap from here to a profound and original Talmudic scholar who lived and operated in Lithuania in the beginning of the twentieth century — Rabbi Shimon Shkop, RSS, the head of the yeshiva in Grodno. His Talmudic thinking involves complex philosophical considerations.124 His primary work is Sha’arei Yosher or The Gates of the Straight, a monumental work on the laws of deciding cases of doubt. In this book, as in his books of innovations on the Talmud, there are a number of profound insights which can be translated into the language of modern logic, and they contain a great deal of innovation for modern logical research as well. Some of his interpretations solve problems in deontic logic, non-deductive reasoning, logic of time, the philosophy of science, legal theory, and more. Here we will deal with one of his insights, which regards the relationship between ontology and epistemology, and indirectly — the difference between quantum theory and the rest of physics. It’s interesting to note that this is the same time period in which quantum theory was created within physics. I assume that the RSS didn’t even hear about it, and certainly was unfamiliar with it in depth. But the zeitgeist seems to have led him to create a similar distinction in the Talmudic context, and it turns out that it is there that the distinction is a great deal easier to digest, and can even clear up some of the mystery surrounding quantum theory. The goal of the RSS is to explain why the first case we brought, a person who betrothed one of five wives and does not remember which of them, is essentially different than the case of someone who betrothed one of two sisters without defining which one. We will now try to study his words. I will bring them as written, and I will explain them step by step for those unfamiliar with Talmudic-halachic language.125 In Sha’arei Yosher, Section 3, Chapter 22, RSS writes as follows: It would seem in my humble opinion that when he betrothed one of five women…and when he betrothed with a cent in his pocket, there is here a deep study in the manner in which the law applies

to every one of them. In the case of a man who betrothed five women, and in the case of a man who betrothed a woman based on one of the cents he had in his pocket, the question of halachic decision is deep and requires abstract study: As here in the betrother of one of five women, where the betrother and betrothed didn’t discern which of them will be dedicated to betrothal, how is it possible to say that in actual reality a woman was singled out for betrothal? For what reason and cause was a woman singled out from Heaven who would be more betrothed and singled out than the other women? But it appears that it is more reasonable to say that in all such cases, even in actual reality, no woman was singled out for betrothal and that in such a case, it is not relevant to say if Elijah comes [in some cases of inability to decide cases, Rabbinic sources set it aside to be decided in Messianic times by Elijah the prophet], and those who say that everyone is forbidden [to others to be married] due to the doubt is not like other doubts, that the teaching of doubt is lest it is certain in reality, and here the matter is that the entire application of the law is because of a reason causing it. When a man betroths one woman of five and does not define who it is (meaning not that he forgot who he betrothed, but that he did not define any specific woman he was betrothing), there is no possibility to say that there is one woman who is truly betrothed to him, only that he doesn’t know which one. After all, Heaven did not assign him any particular woman to be betrothed to him. Obviously such a case is entirely symmetrical, meaning that the status of all the women is identical. This means that we are dealing with an ontological doubt and not a lack of information on the part of any person, which is why even God Himself doesn’t know who is betrothed to him. This, compared to a case in which a man betrothed a particular woman among a group but simply doesn’t remember whom. There we are dealing with a reality where it is certain that there is one woman who is betrothed (he would say: God knows who she is), and the problem is only lack of information on the part of the man, meaning this is an epistemological doubt. And the reason for the law of application of betrothal is the action of betrothal, the giving of the money and the statement and since he gave the money and said to one of these five women ‘Be betrothed to me,’ as by this it is appropriate that one of them will by law be a man’s woman [i.e., forbidden to other men], due to this every one of them is forbidden due to the act, and not because we are doubtful regarding her being the betrothed more than the other four remaining women and so it is in all such cases… What creates this situation is the giving of the betrothal money (according to Jewish halacha, this is the action which creates a state of betrothal between a man and a woman). After the money was given, betrothal legally applies, but there is no woman for whom the betrothal applies more than the others. Therefore every one of the women is forbidden to marry someone else, as there is a legal possibility that she is this man’s woman. Here we also saw above in the case of the Mishna that it is an ordinary, epistemological doubt. Therefore, our case is very similar to a halachic doubt, but here the doubt is ontological and not epistemological. In all such cases, the theory of doubt is not lest it is a certain case, as in such a case there is no certain case even in true reality, but regarding conduct in our case we must act in this with every women by the law of doubt due to the reason causing the law that it will be one of them and it is not determined who is the one and all of them are worthy of being the one. In such a case, there aren’t really laws of doubt, as gaps in reality itself is not doubt. A doubt is a situation where reality itself is perfectly defined but where a person lacks information, as we do not know who is truly betrothed to him; but in this case, all of reality is known to the man, and he does not lack any information. Reality itself is what is not determined and unequivocal. Because of this, the law here is like the law of doubt, as even before Heaven it was not revealed that

one is certainly forbidden and the rest are permitted as in true reality as well they are all equal and there is no difference between this one and that one. So, the law in these cases is like that of an ordinary or epistemological doubt, even though here even Heaven does not know which woman he is actually betrothed to. Here true reality itself is not determined, rather than just our knowledge of reality being lacking, but this also creates a lack of information and therefore the law is like in ordinary cases of doubt. So far, the similarity between the epistemological and ontological doubt. But there is also a difference between them: And according to this in such a case it is not relevant to say that the law of annulment by majority applies, as only where the minority is part of the majority did the Torah say to follow the majority. But in such a case where everyone is equal in true reality as well, there is no minority or majority here, as all are forbidden due to the reason causing that one is worthy of being forbidden and it is not clear who is the one as one of them goes out and one of them comes up. Here he brings the halachic consequence of this distinction. Halacha has a law of annulment by majority, meaning that if there is one woman who is married to him and she is in a group of a number of women who are not married to him, halacha allows us to treat all of them as not being married to him. This is an epistemological doubt, where there is one woman who is betrothed to him but we don’t know who. And in such a case, he can treat all of them as though they were not married, because the majority overcomes the minority and the minority is annulled by the majority. The halachic consequence is that all of them would be permitted to marry someone else.126 But in the case of someone who betrothed an unspecified woman among five women, creating an ontological doubt, there is no majority or minority. In this case, all the women have the same legal status, and there is no difference between them, and there is therefore no case of annulment by majority. There is no majority to overcome the minority, and there is therefore a legal question regarding every one of them, and none of them may be married. This is the halachic difference between an ontological doubt and an epistemological doubt.

The Halachic Case of “Breira” [Sorting]127 There is a whole range of Talmudic issues where we can see the same phenomenon. For instance, when someone wants to drink wine from the barrel, halacha obligates him to set aside part of that wine as a tithe for the Jewish priest and Levite before he drinks, and so long as he hasn’t, he may not drink the wine (it’s called “tevel” in such a case). Let’s assume that there’s a person who says: The last liter left in the barrel will be a tithe which I will give to the priest. He now wants to drink the wine on the assumption that there is one liter which is designated as a tithe. He may not be able to specifically point to the liter, but it’s there. There is a debate in the Talmud as to whether he may now drink the wine before actually separating the tithe: Is he considered to have defined a liter of the wine as a tithe (In Talmudic terms: “There is sorting,” meaning that it could be sorted retroactively), or perhaps we cannot say that there is a particular liter of wine (In Talmudic parlance: “There is no sorting”), and it is therefore forbidden to drink the wine before separating the tithe? What is the status of the barrel now, before he drinks? There are opinions among the medieval Jewish commentators on the Talmud known as the Rishonim, according to which there is no specific liter of wine within the barrel which constitutes a tithe. It’s not that we don’t know where the liter is, but rather that such a liter simply does not exist, as the future cannot determine the past. It is therefore forbidden to drink the wine because of the prohibition of consuming tevel, since the tithe has not yet been separated. But according to another approach, there is an undefined reality here in which one of the liters is a tithe, but it is not a defined liter but simply one of the liters in the barrel. Such a situation is effectively one of ontological doubt, in which there is a liter of wine which is a tithe in the barrel, but this liter is not defined; it’s not that we don’t know what the liter is, but that it is not a concrete liter in reality itself. As the RSS would put it: It isn’t even known in Heaven. Just as we saw in the case of betrothal. According to this method, the probation to drink the wine is not based on the prohibition of consuming tevel, as a tithe has actually been separated; the prohibition is based on the risk that he might consume the tithe, which is meant for the priest and whose consumption is a serious religious infraction in and of itself.

Interim Summary: From Halacha to Physics The halachic situations I described deal not in physical facts but normative ones. This is the reason that

situations involving ontological doubt, which so bewilder physicists, can be more easily understood in that framework. We legally/halachically define a situation as a case of ontological doubt and this causes no contradiction or bewilderment. To the contrary, it seems obvious. Who doesn’t understand that when a man betroths one of five women without defining which one he betrothed, then there is no concrete woman in reality who is betrothed (and not just that we don’t know who it is)? The betrothed woman is a superposition of the five women: Each one of them has an equal chance of having been betrothed (with the odds here expressing ontological and not epistemological uncertainty). The wave function here is defined for the women and not the space. But mathematically, it is precisely the same phenomenon. This is the quantum doubt, which is an ontological one, as it does not just involve a lack of information among human beings. So why is the case of quantum physics so bewildering? It is physically hard to understand how it can be that there is one electron which isn’t in a particular place, only that there is some chance that it is in any one of a few different places. The electron by nature (at least when conceived as a tiny tennis ball), needs to have one defined location; if it is here, then it is not there, and vice versa. By contrast, the status of the woman as being married is a normative determination and it is subject to no physical limitations. There is therefore more room there for the definition which is a superposition of betrothed and available at the same time. Therefore, a halachic analysis can contribute to understanding the significance of an ontological doubt. Precisely because this analysis does not involve shattering our most basic intuitions. In halacha, a state of ontological doubt is entirely understandable, and we can therefore learn about the physics case from the halachic one, or at least better demonstrate the idea of ontological doubt.128

*** The Uncertainty Principle One of the consequences of the picture described thus far is Heisenberg’s uncertainty principle. The description of the electron as a wave, a collection of particle states, and its description as a particle are descriptions which belong to different worlds. When the electron has a wave description, we cannot speak of its location in space, but rather in a number of locations. There is no meaning to the statement “the wave is on location so-and-so.” A wave is a concept which covers the whole space and not any specific point thereof. Therefore when we use the wave description the electron has speed but not location. Quantum physics states at the same time, when the electron is described as a particle, that we can only speak of its location but not its speed or acceleration. Thus, when I know the speed of the particle, which only happens when it is in a pure wave state, there is no speaking of its location, and when I know its location, which happens when it’s in a pure particle state, we cannot speak of its speed. We can even determine the indeterminacy of a particle via mathematical formulation, and this is what is known in quantum theory as the “uncertainty principle.” The first person to formulate this principle, in 1927, was German physicist Werner Heisenberg, whose involvement in the nuclear weapons project of Nazi Germany serves as a source of controversy. Five years afterward, he won the Nobel Prize for his contribution to the foundations of quantum theory. In general terms, the principle states that it is impossible to determine precise values for certain pairs of measurable values at the same time, the pairs in our case being speed and location — the more the state of one of these variables in such a pair is known with more precision, the state of its partner is known to a lesser degree of precision, and may produce many more values. As noted above, according to the presently accepted understanding among physicists this is not just an epistemological argument about our state of knowledge, but an ontological argument about reality itself.

From the Quantum Micro to the Classic Macro What we still lack is a discussion of the question of why we don’t see all this in the macroscopic world. Why is the distinction between waves and particles so clear and obvious, and why does every object at that level have a clear speed and location, in contrast to the uncertainty principle? How can it be that macroscopic physics is entirely causal, while we encounter ontological freedom in the microscopic world at its base? Our experience draws on the macroscopic world, and it is therefore no wonder that we emerge thoroughly confused from the results we get from experiments in the microscopic world, such as the double slit experiment. We know from our experience that every object has a defined location, and there is no problem with it having a single, defined location along with it having a defined speed. Therefore, when we

discover that the situation is otherwise in the microscopic world, this confuses us greatly. So what is really going on along the path from the micro with its quantum behavior and the macro with its classic behavior? The simple conception of the relationship between the micro and the macro is a reductionist approach. A macroscopic body is nothing more than a collection of microscopic particles. A tennis ball is made up of atoms, which contain electrons. The human body is made up of cells which are built of molecules, and every such molecule is also made up of atoms. So if atoms are quantum entities whose physics entail inherent ontological freedom, how can it be that macroscopic bodies which are made of them act entirely differently and lack such freedom? The basis of this phenomenon is none other than the law of large numbers. If we throw a fair dice, the odds of rolling a 5 are 1/6, just like the odds of getting any other result. If we throw the die 6 times, we will get each number once on average. But in a typical series of six throws, that’s not what will actually happen. We will usually get 3 twice, 1 three times, and 6 once, or some other mix, but almost never an equal number of results for each side (a uniform distribution). What will happen if we toss the die 6,000 times? Here the mix will already be closer to the theoretical odds. We will, more or less, get a thousand results for each one of the six sides. If we throw the die six billion times, we will get results which are much closer to the average distribution, with the same number of rolls for each result. This is the law of large numbers. To bring things closer to everyday logic, think of a situation in which we have a mixture of color particles, each of which can be yellow or blue with an ontological probability of 1/2. If we look at the particle of one color and “measure” its color, we will see it as yellow or blue. When we look at ten such particles, we will see a mixture between yellow and blue — i.e., some shade of green. Yet there is still a chance that all of them will end up blue or yellow, but when there are ten particles, the odds are already quite small. There is a greater chance that eight will end up yellow and two blue, or some other mixture, and then we will get yellow with some touches of blue, or blue with a shade tending to yellow. But if we take a mixture of millions of millions of such color particles, we will always get a container of the color green, meaning an equal mixture of blue and yellow. The odds of getting a container of millions of particles in which the lion’s share are the same color is very small, and the greater the number of particles, the smaller the odds of deviation from green. The odds for blue or yellow, the two biggest deviations from green, approaches zero. This is the physical expression of the law of big numbers. The greater the number of experiments or the number of particles in the container — the closer their result will be to a theoretical average, represented by an equal number of blue and yellow. Just like when we toss the die a larger number of times, the distribution of the results will be closer to the average (1/6 for every one of the sides). This is the situation in the color container and in die tosses, and more clearly in the case of a collection of atoms which create a macroscopic body. Every such atom acts in a quantum manner. It can receive different values for its speed and location in quantum probabilities determined by its wave function. But when we look at the location of the macroscopic body which is composed of all these atoms, we will see it precisely in the average location, this being a calculated sum of the locations of all the particles which compose it, without any deviation. Unsurprisingly, it turns out that this average location in quantum theory is precisely the location determined by Newton’s classical mechanics. When I shoot at a crack in the tennis ball, which is composed of very many atoms, some of its atoms are around it and move at slightly different speeds from the speed of the large ball, but the location of the ball is the sum of the locations of the atoms which compose it, and its speed is the average speed, and it therefore appears to us like a rigid body which moves at a uniform speed. It will always move through one slit, because the odds that it won’t according to the law of large numbers is infinitesimally small. When will we succeed in noticing quantum phenomena on a macroscopic scale, in multiple particle systems? When there is a correlation between the particles for whatever reason.129 To demonstrate this, think of the color particle experiment, but now we have a situation where there is a mechanism which forces nearby particles to be similar in color, leading to a correlation between them. For instance, in a case where particle A got a yellow color, this force compels particle B and all its other neighbors to get a yellow color. Such a force can also influence the closest particle, but it can also extend its influence and reach particles which lie at great distances. The force’s range of influence effectively determines the range of correlation between the colors of the particles. If the range of this force is a millimeter, then all the particles in a millimeter’s radius around our particle will be yellow. Let’s assume that we are dealing with a color container with a diameter of one meter. If the range of the correlation between the colors of the particles is at the scale of a meter, then the color of one particle will determine the overall color of the entire container. When the first particle chosen receives a yellow or blue

color, the entire container will be correspondingly blue or yellow. In such a case, we will see the same quantum behavior when it comes to the color of the entire container, and it will be yellow or blue at a probability of 1/2, just like the single particle. But if the range of correlation is at the scale of a millimeter, then as we saw above, our container will always have an entirely uniform color of green, while containing small clumps of about a millimeter in yellow or blue, which will of course provide a green color for the whole container combined. Thus, in large bodies, quantum phenomena are “spread,” and thus average out into classic behavior. This means that when it comes to a container whose size is the scale of the range of correlation (and at a very low temperature130), and we don’t know what color it is, then so long as we haven’t measured it is in a state of superposition between blue and yellow, like all the particles within it. In other words, it is both blue and yellow at equal levels of probability. Only when we “measure” the color of the contents of the container will the wave function of the entire content “collapse” throughout the container together into the color blue or yellow, at equal levels of probability. This is quantum behavior on the part of the whole container; in this quantum field, the container will never be green but rather either yellow or blue, at equal levels of probability. By contrast, when the size of the container is far larger than the range of correlation and is at slightly higher temperatures, the color of the container will always be green, without need for measurement. This is precisely the “classic” behavior of large bodies. Thus, what determines the scale at which quantum phenomena appear is the scale of correlation between particles.131 A body whose size is to the scale of the range of correlation will behave in a quantum manner. A much larger body will be a classic body. This means that quantum behavior disappears when we reach scales which are significantly larger than the range of correlation. The ranges of quantum correlation are usually at the level of nanometers (a billionth of a meter), and are incredibly tiny compared to the sizes of everyday objects around us. Because of this, we cannot see quantum phenomena in our daily lives.

The Problem of Measurement Let’s complete the picture with another aspect of quantum theory. We saw above that measurement does something to the wave function. When we measure the location of the particle to see if it goes through slit A or slit B, the wave function changes character (in Copenhagen terms: It collapses) from wave behavior, a superposition of pure states, to particle behavior in the form of one pure state. This means that in contrast to what we thought before, observation is not a passive act, during which a person looks at the world around him and examines its nature; it turns out that the act of observation itself changes the world we are observing. This is a very surprising conclusion. One of the consequences of this approach is a very famous thought experiment known as Schrödinger’s cat. Think of a cat placed inside a sealed box. Inside the box is a device with radioactive material, which has a 50% chance to decay during the experiment. We build the system in such a way that if the material decays, poison will be emitted which will kill the cat. If the material does not decay, the cat will stay alive. At the end of the time set for the experiment, the box is opened (= a measurement is conducted), and then we know if the cat is alive or dead. But what was the state of the cat when the box was sealed? The obvious answer, at least according to the Copenhagen interpretation of quantum theory, is: “Not alive and not dead, but half this and half that.” Again, we are not dealing with a lack of information regarding the cat on our part, but the state of the cat itself. The cat is in effect in a superposition between the two states (see the image below).

This means that opening the box, this being the act of measurement, changes reality itself.132

One of the greatest mathematicians of all time, John von Neumann, and with him some of the greatest physicists of the twentieth century like Wigner, made an even more far-reaching argument. They examined the question: What is measurement? After all, every particle which moves in the world is under the influence of various physical systems. There are no particles in a completely isolated state. There are always some external systems which are interacting with the particle, and effectively “measure” it. Sometimes a light wave hits it, and sometimes it collides with a particles or other object. It itself does not “know” whether this is measurement or not. If so, how is the system called a “measuring tool” different than any other physical system? In other words, how does a quantum system know that it is being “measured” and decides it must “collapse”? The response of Wigner and von Neumann is that a measurement is defined by some “I” which is behind the measuring tool. This means that the collapse of the wave function, the change in reality, is not the result of the measuring tool itself, but rather the consciousness standing behind it. Put differently, when human consciousness observes a physical system — the system changes its state. What determines matters is not the measuring tool but the consciousness using it. This interpretation aroused quite a bit of opposition over the years, as what is the difference between human consciousness and other tools of observation? How does the quantum system feel if a human being stands behind the ampere meter measuring current? How is it possible that this interpretation still survives until today among some physicists and it is even more popular among philosophers and popular science people? In any event, as of today, experiments have been done without human consciousness, in which the results were registered on a computer without any human being observing the experiment, and the results were identical.133 This means that the quantum tricks are apparently not dependent on human consciousness. This of course brings us back to the question of collapse which we raised above. So, there are various hypotheses regarding the question of when a collapse happens and what measurement entails, but as far as I know, this matter is still open and serves as a serious challenge in reference to the Copenhagen interpretation in general.

Consequences for the Libertarian Question So far, quantum theory in a nutshell for our purposes. But how does all this tie into the question of determinism? First, regarding the aspect of the influence of consciousness on physics. We saw above that measurement causes the physical state to change. The wave function, which describes an electron with a wave character, changes after measurement and becomes a pure particle state, in the form of a function with a particle character. According to Wigner and von Neumann, our consciousness can influence physical reality, much like we saw in chapter two, where we noted that libertarianism must accept the possibility of the creation of physical phenomena from factors of consciousness. If so, this interpretation of quantum theory demonstrates a mechanism of influence by human consciousness on a physical system. According to the argument of von Neumann and Wigner — which as we said is controversial and seems to be untrue — identical physical influences without a human system behind them will not produce the same effects. Another aspect of the connection between the question of determinism and quantum physics is the existence of freedom within the physical picture. To introduce freedom of will into the physical picture, we must first locate freedom within that system. We must try and find if physics allows a situation in which one starting point can develop into a number of possible futures, which would mean that the present does not unequivocally determine the future. We ostensibly found such freedom in quantum theory, and indeed quantum theory is the only ontological gap we know of in physics. Here we found that the development over time of measurable quantities is not determined by the situation in previous times, meaning a number of possible options can develop from a single starting point. In quantum theory, then, the future is not determined by the present, and this ostensibly opens up a significant opening into which we can introduce freedom of will into physics. And indeed, we find quite a few physicists and philosophers, some of them central and important, who peg their hopes to quantum theory to introduce freedom of will and choice into the physicalist picture, thus allowing us to believe in libertarianism without harming the laws of physics.134 One of the most prominent of these is the famous mathematician Roger Penrose, who was involved primarily in the study of the general theory of relativity but also researched quantum theory. Penrose was a particularly colorful and diverse person, who won the prestigious Wolf Prize in physics with Stephen Hawking and the Eddington Medal, along with many other prizes and awards. Among other things, he served as inspiration for some of Dutch painter M.C. Escher’s drawings, when he sent him some illusions and visual paradoxes. Two of them are shown in the image below.



In Penrose’s bestselling book The Emperor’s New Mind,135 he starts his journey in search of a mechanism which would base consciousness on quantum theory. There he raises the argument that quantum theory is what allows us to believe in the existence of consciousness and freedom of will even within the physicalist picture, and he continues this train of thought in his later works. Below, we will examine these two aspects of quantum theory — the influence of consciousness on physics and ontological freedom in physics — and we will reject the relevance of both for the discussion of the question of freedom of will.

Psychophysical Influence in Quantum Theory I’ll start the discussion by addressing the influence of consciousness on physical reality via measurement, as it is simpler to explain why this argument fails as a basis for libertarianism. First, the common belief that human consciousness can successfully influence physical phenomena has no real basis. Wigner and von Neumann’s argument has amazingly survived until today, but this is primarily among philosophers and non-physicists. It is pretty clear today among physicists that there is no real difference in this context between measuring tools and human consciousness. It is indeed difficult to clearly define measurement, and there does not yet seem to be a clear answer to the question of what measurement is, but as we will see immediately, there are even experimental signs that the quantum character is unrelated to the presence of human consciousness. To understand this, let us think back to the double slit experiment, in which we place a measurement tool next to one of the slits, but without a human presence in the lab. Moreover, to ensure that human consciousness is not a factor at all, we will ensure that knowledge of the location of the particle never becomes human knowledge. We will ensure that the disk which registers the results of the measurement is burned immediately after the measurement. We will indeed measure whether the particle goes through A or through B, but no human being will ever know the results of that measurement. If we conduct such a double slit experiment, what will be the picture when the particle hits the screen According to Penrose’s proposal, the picture will be that of a wave — the electron will interfere with itself, as no human consciousness was present at the experiment. But it turns out that such an experiment has actually been conducted,136 and the results were precisely the same as one in which a human was present in the lab. In other words, the picture resulting on the screen was of a particle and not a wave. Therefore, the measuring tool next to the slit caused the wave function to collapse, not any human consciousness behind it or next to it. We can perhaps offer convoluted explanations which would “save” the von Neumann thesis, but here we are already diving deep into mysticism and moving far away from solid science. The second argument against the psychophysical thesis as allowing freedom of will is more principled. Even if consciousness could change the physical wave function of the particle or of the object which it observes, this would still not be the same thing as freedom of will. After all, we saw that the result of measurement in quantum theory is not determined by the measurer. The measurer placed a measuring tool next to slit A, but he does not control the result which emerged — that is, if the particle would go through slit A or slit B. The particle also did not control the result of course, and it could not decide whether to pass or not pass via any slit. As we saw, the result is random, and the odds of getting any one of the results are determined by the values of the wave function. Thus, the particle will do what is determined by the lottery described by the wave function, but certainly not what the observing human being wants or dictates. Thus, even if von Neumann was right, that still leaves no room for freedom of will in quantum physics. If human

will truly determined the results of the measurement, then this result would not fit the values of the wave function, and we thus return to the problem of contradicting the laws of physics and failure to introduce freedom of will into physicalism.

Does Quantum Theory Actually Contain Freedom? The second aspect which we pointed to is the ontological freedom in physics created (or described) by quantum theory. As we saw, quantum theory does indeed teach us that one starting state can develop into a number of different states, just like human free will. But the similarity between the two begins and ends here. The first problem with this equivalence is very much like what we saw in the previous section. Even if our brain acted in a quantum manner, this quantum conduct would still have random characteristics and not those of freedom of choice. Meaning, even if our brain acted in a quantum manner, this would mean that our actions occur randomly and without a cause, and not based on human judgement and decision. Once again we have confusion between randomness and freedom, and we already saw above, in chapters two, six and seven, that these are two different mechanisms. Here, too, we can say that if our will truly determined the physical outcome, then it would no longer fit the laws of quantum theory, as these state that the result needs to fit the relevant values of the wave function. We have once again failed to introduce freedom of will into physicalism, and once again this error is due to the conceptual confusion we keep encountering between randomness and choice. Beyond this, as I explained above, quantum theory is conditional based on scale and temperature, meaning the size of the system in question. When we want to explain our freedom of choice via quantum theory, we are dealing with biological systems, this being the brain. Can we assume that the brain behaves in a quantum manner? Various calculations have been done on this question, and it turns out that the range of correlations for the neural system (at room temperature) is significantly smaller than that of a single neuron. In other words, it is unlikely that even a single neuron behaves in a quantum manner, and certainly not a system of many neurons.137 At this scale and temperature, the quantum theory is already “spread,” and we are left with classic physics. This point is mentioned (with regret) already in Penrose’s The Emperor’s New Mind, but he as a physicalist searches for other mechanisms so that he can nevertheless introduce quantum theory into our brain in various esoteric ways — Penrose searches primarily for an explanation for the very existence of consciousness and mental phenomena — meantime without any real success. As noted above, even if he succeeded in his task, this would not amount to a physical explanation for the phenomenon of free will. Here it is worth mentioning the proposal of Rabbi Prof. Yehuda Levi.138 He uses a combination of chaos and quantum theory, proposing that the combination does what each one of these mechanisms cannot do on its own. His argument is that quantum phenomena do indeed occur at atomic-level small scales, but the butterfly effect of chaos teaches us that sometimes the small scale also affects the large scales, like the butterfly creating a tornado. Thus, per his system, chaos is a mechanism which allows quantum phenomena which affects larger scales. He also noted the mechanism of sliding in physics which describes a similar phenomenon. This argument seems problematic to me, as it misses the significance of the range of correlation. The classic behavior of large objects points to there being no option in quantum theory for influence from the small scale to the large scale, but there is a correlation at the range which reaches the larger scales (like in superfluids or superconductors). It is not likely that chaos can do what quantum theory itself does not allow. In addition, there is also no sign of such a mechanism in our brain. If the brain contained a chaotic structure which conducts phenomena from the small scale to the macroscopic, then we should be able to see all sorts of quantum phenomena beyond our free will which would move up from the small scale to the macroscopic scales. I know of no such observations. In any event, even if this argument was correct, I showed above that this still cannot explain freedom of will. As we saw, quantum phenomena are random, lacking cause or reason, while free choice is a different mechanism based on judgment, and having reason even though not a cause.

Between Randomness and Choice, Again Our conclusion is that even if quantum theory was relevant at these scales, freedom of will still wouldn’t be capable of being introduced into physicalism. Quantum theory does express ontological freedom within physics, but we already saw in chapter two that there are two kinds of ontological freedom: randomness and choice. Quantum theory offers us randomness within a certain distribution, while freedom of choice speaks of an entirely different kind of freedom in the form of decisions based on judgment.

Thus, even if we adopt quantum theory regarding human behavior we will have to deviate from it, as the events will be determined by our choice and not the odds determined by the wave function. But if we have to violate the laws of physics either way to adopt libertarianism, then why are we even reaching for quantum theory? A dichotomous picture in which freedom of will deviates from physics can exist in classical physics as well. This is precisely what we described in chapter six, where we saw that one can hold a libertarian position if we are willing to exclude free will from the laws of physics (creating an electric field by force of a mental cause and without the physical cause of charge sources). We don’t need gaps in physicalism or quantum theory for that. They add nothing to our discussion. To sum up, this chapter shows that quantum theory, even if we agree that it describes an ontological gap in physics, does not add anything to our discussion of determinism and freedom.

The Possibility of “Hiding” Freedom There are thinkers who refer to the role of quantum theory in a more minor manner. The argument is not that free choice fits quantum theory, but that quantum theory softens the difficulty, and effectively hides the events of free choice. When a person freely chooses, the mental acts on the physical. If we observed such a phenomenon in the framework of classical physics in which rigid causality reigns, this would appear immediately to be a blatant deviation from the laws of physics. Electrons which start to move without a physical cause would immediately catch our eye. By contrast, in the framework of quantum physics, an event of free choice is not so prominent. When a person’s choice determines the result of some action, the contradiction with the laws of physics is not so prominent. It hides within quantum randomness. This is of course possible in principle, but this provides no real solution to the philosophical problems. It only explains why we can assume that physicalism is not true even if we haven’t noticed it until now. But that doesn’t really succeed in introducing freedom of choice into physicalism without limiting it. I will note that similar arguments are made by religious thinkers, who try in this manner to explain God’s control of the world in spite of the natural causality of the laws of physics. If rain comes down because of climate conditions of one kind or another, then how can we explain that God decides if, when, and how rain will fall? Do the laws of nature control events, or God? God’s influence, according to explanations like those we presented here, hide in the quantum gaps. There, too, we are dealing with a hiding of the miracle, but it is still a miracle. Thus, there is a deviation from the laws of physics, but quantum theory tries to hide it, and this is equivalent to the approach which arises within the framework of theological determinism to the explanation proposed above in the framework of causal-scientific determinism.

Summary of the Last Two Chapters The last two chapters were meant to examine whether the libertarian approach could be introduced into the physicalist picture. The question in the background is whether one can believe in free will without limiting the application of the laws of physics, like we did in chapter six. If this was possible, then libertarianism would not require dualism; we could hold an entirely materialist view, in which the entire world is nothing but material, and is also entirely controlled by the laws of physics, but we can still accept the argument in favor of freedom of choice and will. The previous chapter showed us that chaos does not at all open an ontological gap in physics. At most, it shows an epistemological gap, and is therefore irrelevant for our discussion. The present chapter showed us that quantum theory does indeed open such an ontological gap in physics, yet at the same time it fails to solve our problem. First, quantum randomness does not operate at the relevant scale. Second, even if it did for the purposes of the brain and biology, it would at most enable randomness, not freedom of choice. The conclusion is therefore that holding a libertarian position requires some sort of limitation of physicalism and the adoption of a position that there are elements in reality which physics does not control (this is the Lex specialis argument we presented at the end of chapter six). We arrive at the conclusion that libertarianism necessarily leads to a deviation from physicalism and essentially to a sort of dualism. We saw that reliance on quantum theory to explain freedom of will derives from a materialistic position. Those who hold a dualist position have no need of quantum theory, as if we accept the possibility of deviating from the laws of physics, then we can also be libertarians within classic physics (and limit it, accepting the possibility that a force field can be created via mental events). Bottom line, the conclusion is that people who are not materialists do not need quantum theory to be libertarians. And those who are materialists — quantum theory will not help them. They are forced (in a deterministic manner…) to accept determinism. A good example of these difficulties is Stephen Hawking’s latest book, The Grand Design.139 After

describing the attempts to base freedom of choice on quantum theory and rejecting them, he arrives at the conclusion that the world is deterministic and that freedom of choice is an illusion. He of course looks at matters from a physicalist-materialist perspective. From this point of view, quantum theory does indeed not allow libertarianism, as we saw in this chapter. But does that necessarily mean that libertarianism is incorrect? Definitely not. Those who don’t assume materialism do not need to be really concerned with the lack of a physical explanation for free will. He can remain a libertarian; he must simply deviate from physicalism.

The Meaning of These Conclusions: The Identity Between the Fundamental Questions We have thus arrived, de facto, at a correlation between a position on the question of determinism and the question of materialism: Libertarianism is dualistic and materialism is deterministic.140 If we are materialists who believe that the mental is merely a trait of the physical body (the brain), then it cannot be that it acts on the physical. As we noted at the end of chapter eight, two aspects of the same thing cannot act on each other. If spirit is just the way in which we (who is this “we”?) conceive the material, then there is no way to say that it influences the material. Therefore those who accept the activity of the spiritual on the material — must see this spirituality as a separate substance, different from the material. This means that the libertarian approach requires deviating from physicalism, such that not all of reality is controlled by physics. The part which is not controlled by physics is what we call a soul (in a human being) or a spirit (in the world), and thus we arrive from libertarianism to interactionist dualism. In addition, we also arrived at the conclusion that causal determinism, as opposed to the theological variant, is fundamentally materialistic, and it sees the mental as an emergent trait of the material/bodily. Since there is nothing in the world besides the material, the conclusion is that the laws of physics dictate all. The whole world is physics, we already noted that other deterministic approaches are unlikely. So, the main stream of determinism is materialist, and the main stream of libertarianism is dualistinteractionist.141 These are the two relevant approaches claiming the crown, and we must decide between them. Until now, we have been busy arranging the various approaches alongside each other, and figured out which of them we can take with us. At this stage we are left with these two, and from now on we have to examine them one vs. the other. In the next chapter, we begin doing so. 118 Winner of the Nobel Prize in 1965 for his contribution to quantum electrodynamics, one of the developers of the atomic bomb in the Manhattan Project, and no less importantly — an amusing and friendly man with a wonderful didactic ability. 119 For the distribution of strength to be as described in graph 3, we need a certain distance between the two slits, which is dependent on the wavelength of the beam. I will not go into this level of details so as not to complicate the description. 120 The first suggestions were that the detector is doing something to the electron beam. If so, we would expect that when the electron goes through slit B, where there is no detector, the picture would be of a wave. But no: It turns out that a particle picture emerges here, too. It turned out that when there is a detector, the electron panics and becomes a particle, whether it encounters the detector or not. 121 The concept of collapse arouses quite a few difficulties, primarily philosophical but also scientific. A number of interpretations have therefore been developed to avoid it. Here I described the common interpretation, for simplicity’s sake. It’s important to stress that the other interpretations protect causality in its classic sense more strongly, and they therefore are of no benefit to the libertarian, and hence will not be discussed here. 122 The situation is in fact a lot better than that. If we had the complete picture of the speed and location of every particle, then we would actually not be able to understand a thing about the physics of the entire collection. This is defined via mediating quantities, which address and describe the statistics of the totality of gas particles (pressure, energy, entropy, and so on). See more on this in chapter thirteen in the discussion on emergence. 123 Jewish marriage contract which requires the husband to pay his wife a fixed sum upon separation and divorce. 124 Quite a bit of the analysis done by the Bar-Ilan University Talmudic logic group, whose contents appear in the Studies in Talmudic Logic Series, is based on the analysis of the Rashash for various Talmudic discussions. In this series, we base Talmudic thought on modern logic and compare it to other logical structures. 125 I gave a lecture in Hebrew at the Steinzaltz Institute where I explained this within the framework of the Talmudic discussion, and which can be viewed here: http://www.hashefa.co.il/home/artdetails.aspx?mCatID=68393&artID=9600. 126 This is just an example for demonstration’s sake. In practice there is no annulment by majority in this case, for reasons worthy of a separate discussion. 127 This subject is described at length in the fourth book of the Talmudic logic, Michael Abraham, Israel Belfer, Dov Gabbay, and Uri Schild, Temporal Logic in the Talmud, College Publications, London 2011, primarily in the fourth section. 128 In the aforementioned book on Talmudic logic, an interpretation is proposed for quantum physics which allows implementing this normative freedom within the physical context. See ibid, pp. 442-444. 129 Such as in the case of a superfluid or a superconductor. These are two unique examples in which a large, macroscopic collection of particles act in a quantum manner. 130 The description here is correct for a situation in which the microscopic laws of physics fully apply. This happens when the temperature is absolute zero. At higher temperatures, the system becomes more random, and quantum effects are spread out more quickly (meaning at smaller scales), and quantum phenomena therefore almost entirely disappear. In certain contexts, this phenomenon is called dephasing, in which the elements in the system at this scale are no longer in the same phase as the others and are not correlated, or decoherence. If so, to notice quantum phenomena, we must conduct an experiment at a very low temperature and in a smaller system. This is another reason that

we cannot view quantum phenomena at the temperatures and sizes of objects we see around us. 131 In the cases of superfluids and superconductors which I mentioned in a previous note, the range of correlation is fairly large, and we can therefore see quantum phenomena at relatively large scales. However, in most cases this requires very low temperatures. Regarding superconductors, a dramatic technological development occurred a few decades ago, as materials which maintain a large range of correlation even at much higher temperatures were discovered (yet still much lower than room temperature), meaning that their critical temperature was high. This may have very important technological ramifications, and a very intensive search is on for materials whose temperature can be close to room temperature. If they are found, this means that it will be possible to create perfect conductors (almost without energy loss), which will operate in the real world, and not just in a lab. 132 This example is apparently erroneous, and there is an explanation provided in fairly simple language in Fritz Rorlich, From Paradox to Reality, Cambridge University Press, 1989. 133 See immediately below, in the section regarding the psychophysical effect in quantum theory. 134 See, e.g., the following articles: Henry Margenau, “Quantum Mechanics, Free Will and Determinism,” The Journal of Philosophy, 64, 1967, pp. 724-725. Barry Loewer, “Freedom from Physics: Quantum Mechanics and Free Will,” Philosophical Topics, 24, 1996, pp. 91-112. 135 Roger Penrose, The Emperor’s New Mind: Concerning Computers, Minds, and The Laws of Physics, Oxford University Press, 1989. 136 See: R. H. S. Carpenter, Andrew J. Anderson, “The death of Schrödinger’s cat and of consciousness based quantum wave-function collapse,” Annales de la Fondation Louis de Broglie, 31(1), 2006. 137 See the following two articles, for instance, which also address Penrose’s claims (thanks to Dr. Ariel Furstenberg for the references): Max Tegmark, “Importance of quantum decoherence in brain processes,” Physical Review E, 61 (4), 2000. K. Hepp, “Towards the demolition of a computational quantum brain”: http://www.springerlink.com/content/q5m1367n84r415m0/ It should be noted that one can perhaps speak of quantum influence on the brain in other contexts. For instance, in the context of vision (affected by single photons) or in the context of ion channels (See S. Roy, R. & Llinás, C. R., Biologies, 332, 2009.). But this does not seem to have an effect in our area, as the move to the mental scales involves the summary of many more such processes, if it is even possible. See also: Jack A. Tuszynski (ed.), The Emerging Physics of Consciousness, Springer, Germany, 2006. The book is fairly speculative, and deals only in consciousness. So far as I saw, the only chapter dealing in will blatantly ignores its “freedom” aspect. (Cf. p. 449) 138 See his book In the Face of the Challenges of the Age (Hebrew), Olam Hasefer Hatorani, Sinai, 1993, p. 42, as well as his Judaism and Science (Hebrew), Sifriyat Beit El, 1991, at the end of chapter seven. 139 Stephen Hawking and Leonard Mlodinow, The Grand Design, Bantam, 2012. 140 The order of words I chose in this sentence is important: I did not speak of the question of whether determinism is materialist, or if dualism is libertarian. These two connections are not necessary. 141 I speak here only of the mainstream, as there are approaches on both sides which try to reject this correlation. These are opinions which are highly unusual nowadays (very few espouse them today), and they are highly unlikely in and of themselves, and some aren’t even coherent (a few were briefly described in chapter eight).

Chapter Eleven MATERIALISM: A PRIORI CONSIDERATIONS The rifle doesn’t give the hunter the bird, only the body. — Popular saying Let me return to that forlorn corner, where is that life with the joy of simplicity? — The Song of the Land (Hebrew), Yehoshua Sobol Is there a ghost in the machine? — Gilbert Ryle In the two previous chapters, I showed that it is unreasonable to assume the existence of free gaps in the framework of physicalist causality. We therefore concluded that those who hold a materialist position cannot introduce freedom of will into this framework, but at most include randomness or probabilistic causality. The conclusion from all this for our discussion is that those who espouse libertarianism must give up the absoluteness of physicalism, and accept the existence of psychophysical influence; the existence of a non-material component in human beings and the world, a component which can also affect the material world. We therefore see that there is a connection between the two problems accompanying us: The basis for determinism is materialism, and the basis for libertarianism is interactionist dualism. This is precisely the picture which we presented at the end of chapter eight, where we placed our two challengers against each other: Soft materialism, meaning emergence — events are fundamentally bodily, and the mental is no more than the collective traits of the body or our brain. In the last two chapters, I showed that since physicalism does not allow freedom of will, this necessarily leads to determinism. Interactionist dualism (interactionism) — there is matter and spirit, which influence each other. In the last two chapters, we saw that the only way to hold a libertarian position is the interactionist approach. We can now see that an attack on materialism will also be an attack on determinism, and arguments in favor of materialism will also support determinism. What we need to do now is to examine materialism, just like we did with determinism. So, we must first examine materialism at the philosophical-a priori level. In the next part, third in number, we will go on to empirically examine these positions facing each other in the context of neuroscience. In chapter six, we pointed to most of the considerations in favor of the libertarian direction, and we saw that the main reason to nevertheless hold to a determinist position was the principle of causality. This means that the dilemma regarding determinism is based on a contradiction between two basic intuitions: The intuition of libertarianism and freedom on the one hand and the intuition of causality on the other. Deciding this question is ostensibly dependent on the question which everyone needs to ask themselves: Which of the two is stronger as far as they are concerned. In the end, we saw that the principle of Lex specialis renders this dilemma unnecessary. The situation is much the same when it comes to materialism. When it comes to this issue, we once again have two basic intuitions which contradict each other: Our unmediated experience of the existence of our spirit on the one hand (see also in the prologue on the principle of the cogito), and the principle of Ockham’s Razor on the other, according to which we should prefer the simpler theory, and materialism

would appear to be the simpler theory. We must once again ask ourselves how we can decide this conflict. The first and primary part of the present chapter is devoted to this question.

The Principle of Causality Before we get to Ockham’s Razor, let me make a brief comment on the place of the principle of causality in the discussion on materialism. In chapter six, we pointed to the contradiction between libertarianism and the principle of causality. But with further thought we could show that ostensibly interactionist dualism in itself, even without assuming the existence of free will, contradicts the principle of causality. The interaction we speak of between spirit and matter means that physical processes occur due to spiritual causes (not necessarily involving will, as we do not necessarily have need of libertarianism). If so, they occur without a physical reason, which ostensibly contradicts the principle of causality in and of itself. Here we must once again turn to the discussion in chapter six. We saw there that the principle of causality is actually preserved in this situation, as it is incorrect to say that the event occurred without a cause. It has a cause, just not a physical one but a spiritual one. If so, there is no violation here of the principle of causality but at most of the laws of physics, which determine that every physical event must have a physical cause. But we also already explained this with the principle of sufficient reason and Lex specialis. There is therefore no need to repeat that discussion in detail here.

Ockham’s Razor: The Joy of Simplicity After removing the principle of causality from the stage, what we are left with when discussing materialism is the aforementioned Ockham’s Razor. This principle has a number of formulations, not all equivalent to each other. It was originally formulated by William of Ockham, a Christian philosopher who lived at the beginning of the fourteenth century. Ockham was dealing with proofs for the existence of God, and he stated that we should not assume the existence of entities beyond what the facts before us require. In other words, when we are offered a number of possible explanations for the same set of facts, we will choose the theory which includes the smallest possible number of theoretical entities (meaning, entities we did not observe directly). In recent generations, this principle has also been expanded in relation to the number of principles included in a theory, deriving the statement that in science we choose the most elegant/simple/minimalistic theory, with the smallest possible number of assumptions, the smallest number of terms, and so on. Ockham’s Razor has come under quite a bit of criticism in recent years.142 Some claim that it is a principle devoid of meaning, which simply states that we should accept the theory which appears most logical to us. Another argument against the principle is that it only expresses conservatism. After all, what does a “more elegant” theory mean? Simply the sort which fits the assumptions I presently hold. If so, the argument is that this is nothing more than a principle to be conservative, preventing us from opening up in new directions. All this was true when it came to formulations which speak of elegance, reasonableness, or simplicity, without defining what those are. But the original formulation of Ockham’s Razor refers to the number of entities in the theory, and the expansion of this concept speaks of the number of principles, and it seems to me that the two critiques fail when it comes to these formulations. It is not correct to see it as an empty principle or necessarily a conservative one. If we are offered a theory based on fewer principles or concepts, then it should replace the existing one. A common error when using the Razor is using it to support a theory which may be simpler, but which does not explain all the facts. I can reject electromagnetism on the grounds that it assumes the existence of electrical and magnetic fields, while those who do not accept this theory do not assume their existence. This is ostensibly a more cost-effective option, as it assumes fewer entities, and it is therefore supported by the Razor principle. But this is nonsense, as it is clear that the Razor principle does not operate in a vacuum. The Razor principle is a referee in a competition between theories of equal weight. If there are two theories which both explain all the facts, and one of them is more minimalist than its colleague, containing fewer entities, assumptions, or concepts, then it is preferable. But a theory which does not explain all the facts is not even allowed into the competition. It is rejected even before we appeal to the Razor principle, as the facts disqualify it a priori. From another angle we can state that the Razor principle is simply a starting point for discussion. A more cost-effective theory is principle, so long as we have found no fact which contradicts it. The burden of proof therefore lies with those who support the more wasteful theory. But if he brings a fact which supports his theory and refutes the cost-effective theory, then he will obviously have the upper hand. For instance, it’s hard to think of someone who will argue that the quantum theory is the simplest conceivable theory, yet

modern physics still prefers it to Newtonian mechanics. Why? Because it fits the facts, and Newtonian mechanics does not (see the double slit experiment described in the previous chapter). This brings us back to Kaku’s quote brought at the beginning of the previous chapter, according to which quantum theory is the stupidest theory around, with the only point in its favor being that it’s true. If we return now to the question of materialism, some have tried to examine this question against dualism in terms of Ockham’s Razor. They argue that materialism assumes that only one substance exists in the world, the material. Dualism, by contrast, assumes the existence of two substances, matter and spirit. If so, they go on to argue, by force of this fact alone it would seem that materialism should be preferred, as it contains fewer entities. Ostensibly, we therefore come to the conclusion that if we examine dualism against materialism in and of themselves, the Razor principle grants preference to materialism. But all this is on condition that materialism does indeed explain all the facts. We can use the metaphor of the “joy of simplicity” which appeared in Sobol’s song which we quoted at the beginning of the chapter to say that simplicity is a criterion only when it brings joy, meaning correctly describes all the facts. The simplicity of the ostrich which ignores part of the facts just for simplicity’s sake would bring us no joy.

The Question of Subjectivity: Qualia One of the central problems in materialism is that it cannot explain mental phenomena. Such a sweeping statement requires a definition of the term “explanation,” but here I will suffice with an intuitive understanding of this term. An explanation should add to our understanding regarding the process or the phenomenon which we are explaining. As of now, there is no physical-materialist explanation for cognitive or emotional phenomena such as memory, will, thinking, love, and recognition. There is an ever expanding amount of scientific knowledge on the physical factors which accompany these phenomena, and perhaps even engender them. These or those neurons in the brain, or these or those genes in our genome, are responsible for various traits and also mental phenomena. But it’s clear that these factors cannot be identified with the mental phenomena themselves. The physical phenomena at most cause the mental events, or merely accompany them, but they are clearly not the mental phenomena themselves (see more on this in chapter eight regarding Taylor’s argument on the identity between me and my body). When I say that there is no explanation, I mean that we have no understanding regarding this question of causation: How does a collection of molecules, formed in some way, create the possibility to think, experience, desire, understand, love, or hate? We can point more and more to the physical processes which occur in these collections of molecules known as the human brain and which accompany these phenomena, but this is not an explanation for them themselves. In this chapter, I will assume for the sake of argument that there is indeed no materialist explanation for mental phenomena, and thus to clarify the role of the Razor principle. In the third part of the book we will see in further detail why we lack such an explanation. At the beginning of chapter eight, we already presented a number of demonstrations of our subjective dimension known as qualia, and the necessity of distinguishing it from the objective external world. We must make use of it again in the discussion here. In the first chapter of my book God Plays Dice, I mentioned that Bertrand Russell and Moore after him pointed to the fact that the answer to the question “What is the color yellow?” is not “An electromagnetic wave with a length of 565-590 nanometers.” The color yellow is a mental phenomenon which exists only in our consciousness, and not in the world itself. The physical phenomenon which exists in the objective world is an electromagnetic wave at those wave lengths, and it is what causes the mental phenomenon of the conception of the color yellow. An example of Russell and Moore’s argument from a different angle is the thought experiment proposed by Frank Jackson, known as Mary’s Room:143 Marry is a brilliant scientist. For some reason, she is forced to study the world from a room which is in black and white through a television screen without any colors at all. She specializes in the neurophysiology of sight, and receives all the physical information conceivable regarding what happens by us when we see tomatoes, the sky, and our use of the terms red and blue, and so on. She discovers, for instance, precisely which wave lengths from the sky awaken the cornea, and precisely how the central nervous system shrinks the vocal chords and transfers air from the lungs to express the sentence ‘The sky is blue.’ […] What will happen when Mary is freed from her black and white room or when she is given a color television screen? Will she learn something or not? In other words, Mary is a scientist who knows everything there is to know about colors from a scientific point of view, but she has never experienced actually seeing color. The question Jackson asks is: The

moment she first sees color, will she learn something new? This experiment is a more live presentation of the following logical argument: (A) Mary acquired all her physical knowledge regarding the human conception of color before she left the black and white room; (B) Upon leaving the room and first experiencing viewing color, she acquires (some sort of) new information. Conclusion: There was some knowledge regarding the human conception of color which Mary did not know before she left the room. Therefore, not all knowledge is physical knowledge. Jackson used another example in his article: Fred can see a color unfamiliar to ordinary people. We may want to know what color Fred experiences when he looks at an object with this color. Regardless of how much knowledge we accumulate regarding Fred’s conception of colors, we can never know which color he is experiencing, much like Mary who didn’t know the experience of seeing color herself. In both cases, subject A has no access to certain knowledge in the possession of subject B. A cannot know if B is experiencing some trait within his consciousness. This knowledge of subject B is barred to A since A is not experiencing this trait himself. This subjective dimension of our consciousness, meaning what exists only in our consciousness and not in the world itself — that which is beyond the physicality of things — is the qualia, which we already encountered in chapter eight. There, too, we encountered a number of demonstrations of this subjectivity, like the problem called the “philosophers’ chestnut” (Is what I call red also what another person calls red?). We also saw there that sound and light are mental phenomena which exist only in our consciousness. Both have physical causes — electromagnetic or acoustic waves — but the identification between the physical cause and the mental cause it engenders is but a conceptual error. Richard Swinburne makes a similar argument when it comes to the feeling of pain.144 He describes a scientist who researches the phenomenon of pain: He observes peoples experiencing pain with all the tools he can muster, and in the end he has complete knowledge of the physical aspects of the phenomenon of pain. We can now ask if that researcher understands what pain is. In other words: When the feeling of pain develops by the researcher, will he discover something more, something new, beyond what he already knew? Of course he will. The conclusion is therefore that we should not identify mental states, like pain, or mental actions, like memory or decision-making, with any states of our neurons in the brain. The same is true when it comes to love, being a cheapskate, hatred, friendship, lust, sadness, and all other emotions. All these have unique physical brain states which accompany them, and perhaps even engender them. But these emotions and mental states are not to be identified with the physical states which accompany them. The article of Shimon Marom, which we mentioned in the introduction, refers to this point: And it is after all a category error to identify human adventurism with a particular combination of genes, or explain love in terms of the neural network places in the brain stem. It is a category error, even if we succeed in the far future to demonstrate that well-defined damage to the neural network harms our ability to love. This is like looking for the essence of the term ‘travel’ in the engine of the car a person drives. We can take it apart and put it back together, analyze and understand every screw and spring in the engine of the car; moreover, there are parts in the engine of the car which if we harm them, travel via the car will become impossible. But there is no connection whatsoever between understanding the action of the engine and understanding the concept of ‘travel.’ The argument that qualia cannot be understood in material terms requires a study of the phenomenon of emergence, and we will therefore return to this issue in our discussion of emergence in the next part, primarily in chapter thirteen. In any event, if we do become convinced that there is no materialist-physical explanation for mental phenomena, this will pull the rug from under the argument in favor of materialism based on Ockham’s Razor. As we noted above, a theory which does not explain all the facts cannot enter the competition to be refereed by this principle.

The Question of Exclusivity There is another common error involving the use of the Razor principle. Let us assume that we have before us a collection of many and varied phenomena for which we are trying to find an explanation. For instance, the collection of all the physical phenomena. Physics today proposes an overall explanation for the material world, based on four fundamental forces: Strong, weak, electromagnetic, and gravitational. We can of course conceive of a much simpler explanation: There is a demon named omnipotent, which engenders all these phenomena. This is ostensibly a simpler explanation, to be preferred to the scientific explanation. It includes just one entity, while the scientific explanation includes four entities. Meanwhile, both theories explain all the facts. This is the case regarding the collection of historical phenomena. There, too, we seek out different explanations, and we will not be set at ease with the argument that there is some demon

named Histocratus which engenders all these phenomena. We certainly would not accept such an alternative. The reason is not just that this theory is unscientific, as it does not provide us with predictions with are empirically testable. This explanation seems less convincing also for other reasons. We have additional considerations, deriving from accumulated experience or our a priori logic, regarding the reasonableness of such explanations. For instance, we don’t think it’s reasonable to believe in demons for various reasons, and we must take these into account. This means that simplicity, expressed by Ockham’s Razor, is just one of the criteria for reasonableness, but there are additional criteria which we use in coming to determine which explanation is more reasonable. Our conclusion regarding the question which explanation is more reasonable is the result of weighing all these considerations at the same time. We should not arrive at conclusions solely based on considerations of simplicity. The opposite question now emerges, of course: Is the explanation that there is something spiritual in a human being which is responsible for the collection of mental phenomena not like the two honorable demons we dealt with above? After all, here too we are proposing a metaphysical explanation for a collection of phenomena via an entity which no-one has observed. Why, then, do I argue that it is actually dualism which is the better explanation in this case? In fact, the comparison here is lacking. The hypothesis of the existence of the soul does not have to do with an additional entity which is the cause of mental phenomena. The soul is the entity in which they occur. The mental phenomena in and of themselves clearly attest to the existence of another, non-material substance within us. It is therefore not like the argument regarding the demon engendering historical or physical phenomena. Let’s use an example to sharpen the point. Let us assume there are two cups in front of us: One has water and the other appears to have red water. The natural conclusion is that the second cup contains something besides water (concentrate). Is the alternative explanation that both cups contain water but the second cup has water which itself created the red color (i.e., the color emerged from the water), a preferable explanation? Although it has need of one less entity (the concentrate), we will obviously not accept this explanation. The concentrate is not an entity which creates the red color. The red color is the color of the concentrate. Moreover, even if there was room for such a comparison, dualism would still be preferable for reasons unrelated to Ockham’s Razor and other considerations of reasonableness. These considerations alone might make us prefer materialism. The problem is that here we have no alternative theory. In the case of physics or history we had two alternative explanations, and we therefore made use of considerations of reasonableness and the Razor principle. But if what I said above is correct, and that materialism provides us with no explanation for mental phenomena, then determinism isn’t an option. In essence, in this case we have no competition between two theories. The facts have had their say, and therefore considerations of reasonableness are almost irrelevant.

The Significance of the Razor Principle for the Materialism Discussion Bottom line, the question of simplicity is indeed an a priori one. We can immediately state that the number of entities in the materialist picture is smaller than in the dualist one. But in our context, the relevance of this question remains open. If we do indeed reach a situation where both theories explain all the facts, and if there is indeed no other consideration regarding the reasonableness of these theories, then materialism will have the advantage of simplicity. Let’s use another example to sharpen this point. A man stands before a table. He sees it, and he therefore decides that a table is indeed standing before him. Now his skeptical friend comes and claims that it is an illusion. He is imagining a table, but there is no such thing in reality. During the resulting debate, the friend claims that his theory is simpler. He does not assume the existence of the table, while the man is assuming it. If so, the idealist theory, which denies the very existence of an external world, is more cost-effective, and is therefore to be preferred based on the Razor principle. Would we accept such an argument? Obviously many of us would summarily reject it. Some would do so because this “cost-effective” theory does not explain all the facts, such as the fact that there is at least the image of a table. Others would argue that this sounds unreasonable to us, as we know of no mechanism which creates such illusions, or that we have no reason to believe the existence of such a mechanism. Materialism is effectively making an argument similar to that of the skeptical friend. It raises the Razor principle to support the argument that an exclusively material picture of the world is simpler than a picture which sees it as also containing a spiritual component. It effectively argues that if we have no proof of the existence of a spiritual component, then we must deny its existence and consequently adopt materialism.

But can one not say the same about idealism? Just as I argued regarding idealism, here, too, I experience my psyche in an unmediated manner, and certainly no less clearly than if I could view it. An argument like the Razor principle cannot convince me to give up what I consider to be such clear and fundamental conceptions. Here’s an example I once heard from a “repenter” of Chabad (Rabbi Yitzchak Cahan): Two people entered a sealed room, without windows or entrances, and see that it is entirely empty. Afterwards, they leave, lock the only door of the room, and sit in front of the entrance such that it is impossible to enter the room without being noticed. After an hour of sitting, one of them enters the room and reports to his friend with shock that the room is full of furniture. The friend chuckles and says he’s saying nonsense, as he has clear logical proof that the room is empty: The room was empty and no-one entered it, which means that it is still empty. Should the friend who entered the room dismiss what he saw with his own eyes by virtue of that “proof”? It turns out that many of us would not. Our mental phenomena are the object of very intimate and unequivocal observation, no less than what we see,145 and therefore “proofs,” and certainly dubious arguments like the Razor principle, should not cause us to dismiss their existence. If so, we cannot decide the question of the existence of the soul based on the Razor principle. It remains in the air and dependent on figuring out the scientific-philosophical questions: Does material explain all the facts? Is its explanation reasonable or convincing? These questions will be discussed at length in the next part of the book.

Materialism and Morality Another a priori problem in the materialist picture is the question of morality. In chapter three, we dealt with the question of moral responsibility in light of determinism. We saw there that in the deterministic picture, we cannot understand the imposition of moral responsibility on a person for their actions, as moral responsibility is only behavior which derives from autonomous choice. Therefore, a creature which acts in a deterministic manner cannot be judged as moral or immoral. We saw there that in a deterministic framework, we cannot condemn immoral behavior and also cannot expect that people bear moral responsibility. Here we will look at morality from a different angle. As we saw there, one of the Kantian assumptions is that morality is not just a form of behavior. We cannot derive a person’s morality or immorality just by observing his actions and inactions. His motivation and motives matter, as well. Moral behavior is only behavior based on a commitment to the moral imperative (what Kant would call the categorical imperative) or moral norms. But a norm, moral or other, certainly in its Kantian sense, cannot exist in the materialist world. The norm is not made up of material, and it exists in no specific point in the world.146 After all, this world only contains physical facts. Atoms and molecules move from here to there, and are in this or that state. So what is a moral norm in such a world? How is it even possible to explain the existence of the moral category in a materialist-atheist world? And what meaning can the question of whether these physical facts (the distribution of speeds and locations of collections of particles) display criminality or immoral behavior even have? Beyond the very existence of norms, we must ask who they are aimed at. A moral norm does not appeal to collections of molecules and does not deal with them. A collection of atoms is not subject to rules of right and wrong. We can at most say that another collection of molecules will destroy it to protect itself from it. But we cannot say that a collection of molecules condemns another collection of molecules for acting immorally. How can we speak of moral norms and judgments when both judge and defendant are nothing more than collections of molecules? How would we address the argument that “This collection of carbon and hydrogen is a cruel murderer, lacking humanity”?147 These questions do not have, and do not seem to have, a reasonable answer. Here may be the place to go back and quote Rickin’s description which was brought in the introduction to the book: Homo sapiens have no ‘natural’ rights, just as grasshoppers have no such rights. Human beings use what Harari calls an ‘imagined order,’ not because it is ‘correct,’ but because if we believe in it, we can create an order which will protect us. Because of this, Voltaire wished not to tell his servant that there is no God, so that he would not murder him in his sleep. The subject of Rickin’s review, Yuvan Noah Harari, wrote similar things in his book A Short History of Humanity (Hebrew; English version Sapiens): According to the science of biology, human beings were not ‘created’ but developed in an evolutionary

process […] There is no such thing as ‘rights’ in nature. There are only traits. […] And what are the traits which developed in the evolutionary process? Among the traits human beings and other animals have we can certainly include the trait ‘life.’ But ‘freedom,’ there is no such word in the biological language. Like ‘equality,’ like ‘rights,’ and like ‘limited liability company,’ ‘freedom’ is also a concept which human beings invented and which exists only in the language and imagination of human beings. From a biological perspective, there isn’t much significance to the statement that a person who lives under a democratic government is ‘free’ and that a person who lives under a tyrannical regime is ‘not free.’ And what of ‘happiness’? […] Most studies of animals recognize only the existence of pleasures, which are definable and measurable in an objective manner. On this issue there is already quite a large number of conceptual errors, some of which have been pointed out by Tomer Persico.148 But Harari is absolutely right about one thing: In a materialist picture, values — as opposed to other global traits, as Persico rightly pointed out — are indeed imagined.

The Naturalist Fallacy Usually, when a materialist is asked why we should act morally, he will give answers of a naturalist bent. For instance, where does the moral feeling come from? Evolution. What benefit does it bring? It makes people feel good. People give themselves a moral account of the moral commitment they feel, and when they find no explanation for feelings which also exist among them involving moral commitment and critique, they rationalize it from a naturalistic perspective and base them on facts. But here they suffer from two main fallacies. The first is the fallacy of Plantinga which we described in the second part of chapter four — evolution doesn’t really explain the creation of the moral sense. But the main fallacy is entirely different: Such an explanation always commits the naturalistic fallacy of Hume and Moore, according to which facts can never be a sufficient basis for justifying a norm. What do I mean? Morality, what ought to be, cannot be based solely on facts, these being what is. For instance, when I ask a normative question: Why shouldn’t I hit Reuven? There can be factual answers to these questions. For instance: Because it hurts him, or because I have a moral sense. But factual answers to these questions are never sufficient. To present a valid argument whose conclusion is that Reuven should not be hit, we must add an additional assumption, which is not a purely factual one, like the following argument: Assumption A (fact): When someone hits Reuven, it hurts him. Assumption B (relation between fact and value): An action which hurts others should not be done. Conclusion (norm or value): People should not hit Reuven. Now instead of questioning the conclusion of why we shouldn’t hit Reuven, we will question assumption B (which includes the moral content): Why shouldn’t we do something which hurts others? Here we clearly cannot respond with the factual argument “Because it hurts him.” The same is true of an argument which will contain the factual argument that we have a moral sense: Assumption A (fact): When Reuven is hit, it hurts him. Assumption B (relation between fact and value149): The moral sense in me tells me not to do an action which harms others. Conclusion (norm or value): Reuven should not be hit. I will ask the same thing here: Why should I follow my moral sense? Its existence within me is but a fact, and the commitment to follow it is a norm. We must add an Assumption C: The moral sense obligates me. When asked why this is so, the answer cannot be “Because I have a moral sense within me.” There is therefore a fundamental gap between the factual sphere and the ethical sphere. Those who base morality on facts, regardless of the details of their argument, are committing the naturalistic fallacy. Therefore, the argument that we should be moral because evolution created us this way is an error. If evolution created a moral sense in us, this is a fact. The question of why we should act based on this fact is already a normative one, and we cannot base the answer to a normative question solely on facts. We can ask this another way: Is there justification for condemning Shimon for not acting based on his moral sense? Alternatively, is there justification for condemning me if I don’t feel like acting based on my

moral sense? Must I follow every sense or emotion I feel? And if I feel a sense to murder, or go to the beach, do I have to follow that, too? What, then, is the difference between the moral sense and these feelings? The only possible answer, on the assumption that we accept the possibility of moral condemnation and judgment, is that the moral sense is correct, or valid, in contrast to other senses. But what defines its validity? This should be something beyond the bare fact that it exists within us. We will therefore always require an anchor outside of ourselves, and really an anchor outside the material world in general. This is precisely the argument of Kant: Moral commitment is not related to the moral sense we have. This is a consequence of obedience to the imperative; the sense is at most an aid which helps us identify and fulfill that imperative. Therefore, the attempt to base the moral imperative and moral judgment on facts is nothing more than a fallacy. But in the world of the materialist there is nothing more than facts, and physical-chemical ones at that. Norms cannot exist there, and therefore neither can morality. It’s important for me to repeatedly stress that I do not mean that people with a materialist outlook do not act morally. Many certainly do, and seemingly no less than those who are dualists. My argument is more fundamental: Either their ethics are inconsistent — they do act based on commitment to morality, even though such norms cannot exist and/or obligate according to their system — or they only act in a manner which is comfortable for their surroundings in the practical sense, but which should not be seen as moral, as it is not based on obedience to a moral imperative. A person who holds such views is not necessarily less moral; my argument is that his conception of morality is inconsistent.

The Significance of Morality in the Discussion on Materialism In chapter three, when we examined determinism at the a priori level, we dealt with the question of moral responsibility. We saw there that one cannot impose moral responsibility on a person for their actions in a deterministic picture. Here we deal with a different but adjacent question: What does morality even signify? We did so from the angle of materialism rather than determinism. We saw here that it is hard to accept the existence of moral norms in a materialistic picture, and certainly when it comes to their validity for human beings. The conclusion from both these discussions is that the deterministic approach, which as we saw is based on materialism, is entirely contrary to our moral intuitions. Materialism does not allow for the existence of valid moral norms, and it also does not allow the imposing of moral responsibility on people. The materialist-determinist position in general therefore stands in opposition to moral intuition and the concepts which derived from it. This is of course not an argument which refutes materialism or determinism.150 It only forces those who hold such positions to look in a mirror which reflects the heavy price they must pay for holding these views back at them. It’s important for me to once again stress what I already wrote in previous chapters: I am not speaking of a moral cost of materialism — I don’t believe that such a cost exists — but an intellectual cost. Even if a terrible moral cost was involved, if determinism is correct, we must recognize this. Approaches regarding facts — and the questions of materialism and determinism are factual queries — must be examined solely in terms of true or false, not good and bad or cost and benefit. Morality is another example of a very basic intuition that we have, and which the determinist-materialist picture forces us to see as an illusion. Beyond this, we saw that there is no reasonable explanation for the creation of this illusion according to this picture, as Plantinga showed that one can ensure unharmful conduct even without creating illusions of judgments, norms, and the like. Thus, those who hold such views must forgo their most basic intuitions of morality, and seriously consider whether the evidence which forces him to do so — this being a particular interpretation of the principle of causality like we saw in chapter six — does indeed compel him in this direction. In my opinion, this is certainly not the case. At the beginning of the chapter, in the discussion on Ockham’s Razor, I mentioned the fact that the materialist-determinist picture does not explain the creation of the mental in general. This is a more general argument, and as we saw, it stands at the basis of the decision based on reasonableness (like the Razor principle), and even precedes it. The next natural step is to examine the question itself: Does materialism truly not provide an explanation for mental phenomena? This brings us to questions which are scientific at heart, and which we will deal with in the next part of the book, which takes us from the philosophical and a priori to the scientific and the empirical. 142 See the following and its sources: Dieter Gernert, “Ockham’s Razor and its Improper Use,” Cognitive Systems, 7(2), 2009, pp. 133-138. I’d

like to thank Dr. Vladimir Weinstein for an interesting discussion we had on this subject in the wake of my book God Plays Dice. 143 See his articles: Frank Jackson, “Epiphenomenal Qualia,” Philosophical Quarterly, 32, 1982, pp. 127–136; Frank Jackson, “What Mary Didn’t Know,” Journal of Philosophy, 83, 1986, pp. 291–295, as well as the Wikipedia entry on the “Knowledge Argument.” (My thanks to my son Nachman who brought this example to my attention). 144 See his book: R. Swinburne, The Evolution of the Soul, Clarendon Press, Oxford, 1986, revised edition, 1997., and the “Dualism” entry in the Stanford University Encyclopedia of Philosophy. For a brief presentation of his arguments along with those of Jackson on Mary’s Room, see the lecture series classes.uleth.ca, in the following chapter: Philosophy of Mind I- Swinburne, Jackson, Papineau.doc. An interesting discussion which includes him and where some arguments on this issue are raised can be found on YouTube under the heading: “Is there a ghost in the machine?” 145 Here is the place to mention what we saw in the prologue, in which Descartes’ cogito argument leads to the existence of the soul being more fundamental and obvious than the existence of the body. Similarly, clear and obvious internal feelings can be no less certain than vision. 146 A question might be raised regarding the argument here. The laws of nature are also not material bodies, yet there are still materialists who believe in their existence (others, who have a nominalist approach, really see the laws as structures which exist only in our consciousness). But this is not the same as the question of moral norms. One can conceive the laws of nature as traits of things or entities, but they are not entities in and of themselves. By contrast, a moral norm is not a trait of something but rather of a moral ideal, if we view it as a sort of object; see more on the ontological status of the ideal in the third book of my quartet, Human as Hay. 147 This is reminiscent of Richard Taylor’s arguments against materialism, which we presented in chapter eight. 148 See the website The God Loop (Hebrew), the article “A note on the conception of man and morality in Sapiens by Yuval Noah Harari, 26.11.2012. His criticism primarily focuses on the question of emergence which will be discussed below in chapter thirteen. In his article, Persico points out another error by Harari, who presents a naturalist ethics, basing morals on facts. This of course fits his materialist approach, but is entirely out of sync with our ethical intuitions. As I noted, this passage contains additional errors, such as the assumption that pleasures can be tested in a quantitative and objective manner; pleasure is part of the qualia and can thus not be tested in any objective manner. See also Daniel Algum’s book Conception and Psychophysics (Hebrew), Broadcast University, Defense Ministry, 1986. By the way, to the best of my understanding, the study on the behavior of bees and their considerations regarding pleasures mentioned above towards the end of chapter two, suffers the same fallacy. They assumed that the degree of pleasure is proportional to the degree of stimulation (the amount of percentage of sugar). 149 In another respect, it is a pure fact: I have a feeling that such and such. 150 I believe we should entirely oppose the adoption of worldviews based on costs or benefits, moral or otherwise, which result from them. The true must not be subordinated to the beneficial; that is pragmatism, which I consider to be an absurd position.

Part Three Freedom of Will and Neuroscience In the first part of the book, we presented both ideas opposite each other: Determinism VS libertarianism. In the second part, we tied the question of determinism to the question of materialism, and presented the two possible pictures in a fuller sense: Determinist materialism on the one hand, and libertarianist dualism on the other. We also saw various philosophical considerations in favor and against these two pictures. These considerations were raised throughout the long history of discussion on these issues. But as I already noted at the beginning of the book, the discussion has reawakened in recent years primarily in the wake of new findings and conceptions which arise in modern neuroscience. It would appear that the discussion has gone from the field of philosophy to the field of empirical science. Moreover, as I noted, it would appear that research institutes are arriving at a consensus — one I consider worrying — among researchers, in favor of the materialist-determinist picture. The present part is devoted to an examination of these questions, this time at the scientific level. We will examine these two pictures in light of various findings and common assumptions in neuroscience. I will not be able to get into all the details here, but I will present the primary considerations and findings. My aim here is to show that there is no real basis for the emerging consensus in favor of deterministic materialism, at least not at our present state of knowledge. We will see that at least at the present stage of knowledge, the question remains a philosophical one, just as it was understood during the millennia of discussion and inquiry into the matter. The question remains open scientifically, and the time at which it will be decided scientifically, if it ever occurs, is still a long way off. In chapter twelve, we will present an entirely general picture of neuroscience to provide the reader with the basic principles of the field. In chapter thirteen, we will examine a central idea raised by materialists to explain mental phenomena: Emergence. This is related to a developing field in modern physics known as complexity, and we will therefore have no choice but to touch on this issue somewhat. In the following chapter, we will discuss the significance of the Libet experiment on which modern determinists rely, and we will see that the accepted interpretations of these experiments are rather hasty. In chapter fifteen, we will discuss the connection between the brain and mental phenomena in a more general way, among other things in light of fascinating phenomena which appear in damaged brains. All throughout this part we will see that the determinists offer interpretations which are certainly not necessary, and that for now, the present findings do not support their position.

Chapter Twelve Neuroscience The goal of this chapter is to present a schematic description, albeit sufficient for our purposes, of what we know about neuroscience as of today. We will see what the brain is and how it works, its relationship with other organs, and the processes which it activates and engenders, and how it does so. This schematic description is very similar to a functional description of a computer. Every component in the computer is wonderfully complex and its physics is very complicated, with each component containing innumerable particles, but we would describe the manner of its activity with a functional block diagram — units for processing, memory, input-output, and more — which are related to each other in various ways. This would be sufficient to understand the activity of the computer. The same is true when it comes to the brain. The description of the activity of the neural cell or synapse could fill thousands of articles and books. But all this is unimportant for our principled functional discussion. I will therefore describe the brain primarily in a functional manner here, and I will almost entirely avoid the small details. In places where I do so, this will only be to the extent that this is needed to understand the brain’s general and principled functioning and its consequences for our discussion.

A Brief History of the Brain Thinking and mental activity of human beings have been a mystery for thousands of years until our own day. The role of the brain in the creation of mental and cognitive functions may seem to many of us today to be obvious, but it is far from being so. In fact, in the beginning, not many understood the central role of the brain in mental events and in the creation of our conduct. When Egyptians were mummified, for instance, the brain was removed from the dead body and thrown out, unlike other organs like the heart, stomach, or lungs, which the Pharaohs kept next to the dead man so that he could continue to use them after he was brought back from the dead. Some have concluded from this that the ancient Egyptians believed that the brain is an unimportant organ. Precisely against this background, it’s interesting that already in the fifth century BC, famous Greek doctor Hippocrates (father of Western medicine, to whom the famous oath is attributed) said/wrote: Man must know that from the brain, and only from the brain, awaken our pleasures, our happiness and our laughter as well as our sadness, our pain, our suffering and our tears, and especially — through it we think, see, hear and distinguish between good and evil. The fundamental change in the medical world’s conception of the brain begins, according to some historians, in the period of ancient Rome, when doctors began to notice influences and changes in the behavior of people whose brains were injured due to gladiator duels. Still, in ancient sources such as Galen, the three — the brain, the heart, and the liver — were seen as the managing backbone of the body. Today, the centrality of the brain is clear to every child, but over the years scientific study of the brain moved slowly, primarily due to technological and conceptual problems. Only towards the end of the twentieth century was a focused research effort made in this direction, which now takes place under the heading of “neurosciences.” This heading encompasses a concentrated multi-disciplinary effort whose aim is to study the human brain and its various functions. These studies include psychologists and psychiatrists, neurologists, biologists and people from the life sciences, mathematicians, computer scientists, physicists, engineers, cybernetics and systems researchers, chemists, and others. All of these use advanced imaging technology and tools of various kinds to observe the brain and analyze its activity, and build sophisticated brain simulations to understand these processes in greater depth. In recent years, brain studies centers have been quickly erected throughout the world, and an enormous sum of money is invested in them. Progress in this field is indeed advancing very rapidly. In fact, it’s hard to speak of the “latest findings” in this field, when these are changing all the time. The field is incredibly dynamic, and sometimes even some of the most fundamental conceptions change with the advance of research. In recent years, we have even seen various practical applications of these findings, primarily in the fields of medicine, and it should come as no surprise that hopes regarding this field are stratospheric.

The Brain: A General Map Today we know that the brain is the primary organ in the nervous system of all developed animals. It coordinates the actions of the various bodily systems, and carries our activities like decision making, thinking, wanting, memory, and learning, and engenders various emotions. At the same time, it is also responsible for receiving information from the outside via the senses, in order to use it to manage the activities of the human being and form his insights.

Image 1: The Human Brain

One might get the impression from this description that it is a sort of computer which includes input-output functions, memory, a processing system, and so on. Indeed, various models for the brain draw inspiration from the workings of the computer, and today there is even influence in the opposite direction, with computers operating based on models derived from what we’ve learned about the human brain. In terms of the physical description, the human brain is composed of three primary structures: The cerebrum, which takes up most of the space in the skull; the brain stem, which is at the bottom of the brain and is connected to the spine; and the cerebellum, which sits underneath the cerebrum and behind the brain stem. Along with the spine, this is the central nervous system, whose orders are sent to the peripheral nervous system and through it to the various organs. The cerebrum is the part which is important for our purposes. It is responsible for our cognitive activities, and we will therefore focus on it.

The Hemispheres The most basic division of the cerebrum is into two hemispheres or semi-circles, a right one and a left one, which are connected via a region of white matter known as the corpus callosum.151 We know today that there is a cross-relation between the hemisphere and the body: The right hemisphere controls the left part of the body — the left hand, left foot, and so on, and vice versa. This cross-relation also works in the opposite direction, of course: Data coming from the right eye or ear will be processed in the left hemisphere, and vice versa. A common opinion is that there is a division between these two hemispheres for various functions. The left side is responsible for linear thinking such as mathematics and logic, analysis and systematic organization of data, speech, and so on — in short, all analytical and systematic functions. The right hemisphere, meanwhile, handles creative and intuitive thinking, parallel processing, holistic thinking, artistic creation, and so on. There is also a common opinion that there are clear differences between men and women in this matter, with women being stronger on the left and men on the right (contrary to the stereotype).152 The crossing of the sides is a clear and agreed upon fact. But all the other distinctions I noted are not yet agreed to by all researchers. Moreover, we’ve known for some time that even the various functions of the two hemispheres are fairly dynamic, and that one hemisphere in the brain can be taught to replace its neighbor, at least partially. Still, let us demonstrate an interesting phenomenon which points to a difference between hemispheres.153 Look at the two pictures in image 2 and try to determine without thinking too much: Which is the happier?

Image 2: Physical Expression of Happiness

It turns out that at least 80% of the population point to the picture on the right as expressing more happiness. This is a strange result, as there is no difference between the two pictures besides them being reversed. Why, then, is there such a strong tendency in favor of the smile on the left? A common explanation for this lies in the difference we noted between the right and left hemispheres: Meaning, the crossing between sides of the body and hemispheres. First, we need to take ourselves into account as observers. The part in charge of identifying emotions in our brain is on the right, and therefore what determines our conceptions regarding the emotions of those we are facing is what is absorbed in our left eye. The picture on the right is therefore considered to be expressing more happiness, as there the smile is seen by our left eye (as we are standing in front of him). We can of course ask if this is true as far as the smiling person is concerned. If the guy on the right turned out not to be happier, we would probably learn this from experience and correct our understanding. Is the guy simply influenced by how we are viewing him? First, it is certainly possible that he is, as a person expresses emotions in ways that are affected by how those who surround him conceive him; when you want to express something, you must consider the observers who pick up on these expressions. But in this case, it does not appear that it is necessary to go that far, as the left part of his face is controlled by his right hemisphere. If so, his mood is indeed expressed by what the right hemisphere expresses. It just so happens that his angle and my angle as an observer align with each other (obviously this is true only if we are standing in front of each other; if we had met via a mirror, the situation would be different).154

The Lobes Let us continue with our description of a general map of the brain. The cerebral cortex of each one of the hemispheres is divided into three lobes, as can be seen in image 3: The frontal lobe, located in the front, the parietal lobe, which is at the top on the brain and which is responsible for receiving information from the senses in order to create a spatial picture. The temporal lobe, which is on the side, is responsible for processing visual and audial information and identifying faces. The fourth lobe is the occipital lobe, and it is responsible for processing visual output from the visual system.

Image 3: The Brain’s Division into Lobes

It’s important to note that this breakdown between lobes is very rough, and most brain activities involve a

number of regions which can be present in different lobes.

Brain Activity at the Neural Level: The Neural Network Biologically, the brain tissue is divided into two main parts: gray matter containing nerve cells and white matter mainly containing axons which connect the grey regions. The nervous cell or neuron is the basic unit of the brain. The brain is primarily a dense and extended network of neurons, and they are interspersed with short fibers called dendrites. The axon is a long fiber which leads the electrical information created in the cell onward to other neurons. The dendrites are fibers which enter the neuron and bring it information from the axons of other cells. The connection between neurons is effectively a connection between the axon of the originating cell to the dendrite of the recipient cell. This is a very complex chemical and electrical process which is done in the region of interfacing between the two neurons, called the synapse. This connection takes place via chemical materials known as neurotransmitters. As noted, there is an asymmetry between axons and dendrites, as each cell has one axon which conducts the information within the cell outward, but there are many dendrites, as it is connected to many axons (each one via another dendrite) which brings it information from other neurons). If so, every axon of every neuron can connect to many other neurons, whose number can range from a few dozen to tens of thousands. A normal nerve in our body is but a collection of hundreds or thousands of axons from various neurons. The dendrites are short and local, with each of them leading from a long axon which arrives from the originating cell to another region, and every dendrite takes the information to one of the cells in that region, but the axon can be very long; there are axons which reach the length of roughly a meter (the axon of the thigh neuron reaches from the base of the spine to the toe of the foot. It can be up to a few meters in a giraffe, for instance). The information passing between the neurons and along these fibers is a series of electric pulses, each of which is called action potential, or “spike” in the accepted neuroscientific jargon. The action potential is created in the neuron by a collection of inputs which it receives from the axons of the neurons it is connected to. The neuron absorbs all the information which reaches it from its dendrites, sums them up in some way, and according to some criterion regarding the result it “decides” if and when to create its own spike (see image 4).

Image 4: Schematic of a Neural Network. In the network structure in this example, axons of six from the neurons (thick line) enter the dendrites (thin line) of our neuron, marked in grey. The axons are long, and the dendrites are short and are present only around their cell. The other neurons, some closer and some further away, have no connection to our neuron. The squares drawn at all connecting points between dendrites and axon represent the synapses. Every neuron of the six sends electric signals through its axon to our neuron, and our neuron sums up all the signals and decides according to some criterion whether to send out its own signal. This signal will flow onward via the axon of our cell (for demonstration purposes, I marked it with a double line, but it is no different than all the other axons in the network).

The spike created in the cell moves on through the axon of the cell and is sent to all cells connected to it. It joins information from other axons which enter these sells to be summed up in every one and create its own spikes in every such cell. These spikes will also be sent in the axons of the recipient cells, meaning they will move further along the network (and sometimes even return to the original cell). The original cell itself continues its activity all the time; at every moment it does or doesn’t create its own spike, according to the new information coming into it from its dendrites at that moment. It therefore continues to create more and more spikes, sending them in the same manner, as well. Thus do such spikes leave and enter

every neuron in all directions along the neural network. Every individual spike in this series has a uniform structure, determined by the physics and chemistry of the neuron. This means that every one of the spikes lasts the same length of time and at the same level of intensity. The differences between types of information and activity of the different neurons are determined not according to the duration and intensity of the spike, which are fixed, but by the rate of spikes in a series, and effectively by the structure of the series along the timeline. Obviously, the structure of the series over time in every axon of every neuron is different. Even when it comes to a different action by the same axon (sometimes it participates in the identification of Grandma and sometimes of auntie), there is a different series of spikes which enter and leave it. Examples of two different series of spikes are presented in image 5. If we present such a series with the numbers 0 or 1, which represent spikes along the timeline, with 0 meaning that there is no action potential at this time and 1 meaning there is, we will once again see the similarity to the activity of a computer.155

Image 5: Different Series of Action Potentials Leaving Various Neurons. The bottom axis is time. The vertical axis is the strength of electric voltage. We can see that every spike in each series has a singular structure, and the difference between the activities of the different cells is only in the structure over the time line (i.e., the locations and gaps between spikes).

Ultimately, the information or series of spikes reaches the target cell, a cell which is located in one of the organs of the body, and which is responsible for activating that organ in various ways, according to the information it receives from the brain. Every series of spikes instructs this cell, and through it the related organ, to act in a different manner. Thus does the brain manage and determine the activity of organs.

Input and Output To round out the picture, we need to describe another aspect. We saw how the brain activates the organs, giving them instructions when and how to act. This is an action of issuing instructions from the brain outward (output). But there is also the entry of information into the brain itself (input). Where does it come from? The senses, of course. For that, every one of our five senses has receptors, responsible for taking the information received from the relevant sensory organ — the eye, the ear, the nose, the mouth, the hand, and so on — and translating it into an electric signal. This signal flows to the brain and reaches the neurons, and thus begins their activity which reacts to the information received. If we return to the example I brought at the beginning of chapter six, the visual information that someone lifted his hand (to strike me) is translated by the receptors into an electric signal which flows into my brain.156 There, it reaches the neurons responsible for the absorption of information from the visual sense, and thus the information received in the form of a particular series of spikes is transferred to the part responsible for processing visual information. The product of this processing is another series of spikes which reaches the decision-making portion. There electric activity awakens, whose result is a series of spikes which ultimately flow to the target cell in the hand, which translates this information to deliver an order to the hand to lift itself and strike the assailant or defend against his blow. The brain can of course send another set of spikes, like one telling my legs to run.157 To sum up, we know of three kinds of neurons: 1. Sensory neurons: These transfer impulses picked up by the receptors to the central nervous system. The receptors pick up the physical/chemical changes and translate them into

electrical signals or spikes which go through the sensory neurons located in the sensory organs, muscles, skin, and joints to the brain. 2. Motoric neurons: Bear signals from the brain and the spine to the muscles and the glands and activate them. 3. Interneurons: “Connecting” neurons, which receive signals from sensory neurons and send impulses to other interneurons or motoric neurons. It’s important to note that this network is not static; it is always changing. The action potentials flowing along the network can change it. Thus are new connections or synapses formed and other existing ones destroyed, strengthened or weakened, and there is also the creation of new neurons (primarily at a young age). The process of destruction of cells happens all the time, and an adult human being experiences the destruction of tens of thousands of cells! As we will see immediately, the dynamism of the brain structure has great importance in understanding processes of learning and memory, and it is related to the plasticity of the brain mentioned already in chapter one.

Actions of the Neural Network Let me begin with two fundamental points: First, the neural network at a given moment is defined by its structure, meaning by the connections between neurons (which neuron is connected to which neuron). The structure of the network can be represented in a giant table, with columns and rows registering all the neurons, and with each box marking if there is or isn’t a relation between two neurons in specific rows and columns. Therefore, if any relation is destroyed or created, a different network is then formed. Second, the information is stored in the neural network through its structure. Every network structure means a certain collection of information such as a familiar face, known skills, scientific knowledge, memories of books we read, and so on. When the network has a different structure, this means that it contains different information. Let us demonstrate this with the action of identifying a face. When we meet some familiar person, we identify him. This act of identification is based on a comparison between the information which comes from the senses — in this case, the visual senses — and the information already stored in our neural network. Thus, the facial identification requires first the study and remembering of the face when making someone’s acquaintance, and afterward, in the next meeting, the conducting of a comparison between the visual information before us and the visual information stored in our memory, thus allowing us to recognize the person. Let us look at image 4 above, and for the purposes of our discussion assume that it is describing a neuron participating in facial identification. The sensory information comes to it from the axons which enter it, it is summed up within it in some form, and after it arrives at the conclusions of the sum it determines whether to issue a particular series of spikes. The significance of this series of spikes is that I have identified Grandma standing in front of me in the flesh based on the look of her facial expressions. What will happen if part of Grandma’s face is hidden by the branch of a tree? How can I nevertheless identify her? We can say in a simplistic manner that our neuron is “programmed” in such a way that even if only five of the six neurons connected to it send the “right” series representing visual information from Grandma’s face, it still issues the right series of identification and the partial identification is also made. But how does it know to issue the series of neurons which points to identification? Our neuron also needs to compare the information which reaches it to information stored on the network. There is no information there in the sense we may think of it like the memory of a computer, as the neural structure itself is the information in question. When there is a network structure which takes the visual signals of Grandma and causes the issuing of a series of spikes which point to her identification, this means that this network structure contains the visual information of Grandma’s face. In this sense, as I said above, the network structure is the information stored within it. Here we must note that there is an essential difference between the act of learning and the act of identification. When we recognize a new face, we are in effect storing some new information, and thus doing an act of learning. The fruits of that learning are stored in the memory, and we are thus remembering. What does this mean? As we saw above, it is clear that the network structure has changed. Learning or remembering mean a change in the brain state, as now it has information it did not have

before, and the structure of the network thus necessarily changed. As I explained above, the visual learning of Grandma’s face means the creation of a network with a structure which causes the visual information coming from that particular face to lead the relevant neuron to issue a series of spikes of identification, sending us the message “Warning! Grandma’s in front of us.” In any event, the conclusion is that learning is carried out through the creation and destruction of synapses, or the changing (strengthening or weakening) in the strength of connections they create.158 The above has been a schematic description of the actions of learning or remembering. By contrast, the action which is done in the second stage, in which we identify the face whose visual information is already stored in the network, should not change a thing about our brain, as no new information is entering it. What is happening now is an action which uses existing information. The system needs to do a comparison between the sensory input and the information already existing in the network, and determine if it is indeed seeing Grandma. This is a fundamentally different action than learning and remembering, as it does not involve the destruction or creation of connections between neurons. In light of what we said so far, we can easily see what sort of actions will require a change in the network and which kinds of actions will require use of the existing network. This distinction is not that clear-cut, of course, as using existing mathematical knowledge to solve a new problem also involves a dimension of learning, and the identification of a person standing before us is not just a comparison to existing information but also the adding of new information, such as the information that this person is now here. Thus far, we described some of the basic functions of the brain. We saw that the brain makes a decision to make a mechanical action (in the top part of the frontal lobe), and transfers it from there via a path known as the corticospinal tract via the spine to the motor neurons, which are connected via synapses to muscle cells which turn the signal into a mechanical reaction. We also spoke of sense, which receives information from the receptors in the senses and processes them in the brain, and after which there is a more advanced processing done in the regions known as associative regions (as the processing is in comparison to other information). It’s interesting to note that the picture which reaches us from the senses is absorbed in the same manner as a picture which emerges from our imagination. Here we are speaking only of the processing, and not the part related to the sensory receptors.159 We focused on the activities of study and memory. The main activity there was the strengthening and weakening of certain synapses, as well as the creation of elimination of other synapses. Part of the learning is in response to stimulation (classical conditioning), and another part is learning based on the reaction of the environment to our behavior (operant conditioning). But the basic actions of the brain do not just include these kinds of action but also many other kinds of activities. For instance, emotional activities take place in the brain in the limbic system (although today we know that there is no particular place where it is located). Strong emotional actions, such as a reaction of fear, awaken the sympathetic nervous system, and it is expressed in a rise in heartrate, blood pressure, and the excretion of adrenaline. Feelings processed in the pleasure center create a sort of “learning,” as the human being understands that this or that action pleasures him, and due to the motivation to repeat the action, he changes the network in a way which will encourage it. In the frontal part of the frontal lobe, there are mechanisms focused on regulating and blocking emotional reactions which are too strong, such as anxiety or aggression. The identification and creation of facial expressions are also related to the creation of emotions. In other words, there is a fairly strong connection between the sensory systems and the systems responsible for processing and creating emotions. Another type of basic action of the brain which is being extensively researched is the use of language and verbal skills such as coding processes and use of grammatical rules, and it is today accepted among researchers that there are a number of regions in the brain which work together in these actions.

Example: Moral Dilemmas To get even closer to our subject, let’s get into a brief description of studies done on brain activity in the context of moral dilemmas There is a lively debate between philosophers as to whether moral conduct is rational or emotional. The use of modern imaging technology shows that at least at the neural level, both sides are right. Moral conduct mixes these two centers of brain activity. For instance, in the experiments conducted by Josh Greene, he presented test subjects with moral dilemmas while they lied down in a device scanning their brain, and followed their brain activity while they were experiencing these dilemmas and deciding them.160 In one of the experiments, he presented the following dilemma: A group of people with a baby are hiding

in a bunker from the enemy, and the baby is starting to cry. The emotional instinct is that we mustn’t kill the baby in an attempt to silence him, but rational thinking could justify killing him, as if not then everyone (baby included) will die.161 The brain scan showed that this conflict led to an increase in the activity of the frontal lobe, whose tasks include decision-making, long-term planning, and cost-benefit calculations. The more the decision referred to the general good, the greater the activity. Additionally, people whose lobe was damaged revealed themselves to think in more utilitarian terms, and they had no emotional conflicts. Greene reports162 the results of a test involving two classic situations, trolley problems, which were often used even before the neurosciences: 1. The double track problem: In image 6 there is a situation where the engineer is advancing along the tracks and is about to run over five people unaware of the danger. I, a bystander, can press a button and change the track so that he only hits one person. It turns out that many feel that the dilemma of whether to push the button or not is relatively simple, as there is a fairly easy argument in favor of switching the track.

Image 6: The Double Track Dilemma

2. The bridge problem: A fairly similar problem is described in image 6, but here there is only one track, and the possibility of saving the people on the tracks is only by pushing a fat man standing on the bridge so that he stops the train with his body.

Image 7: The Fat Man on the Bridge Dilemma

In both cases, we are dealing with the alternative of one person against five, and we would ostensibly expect the attitude of the test subjects in both cases to be the same.163 But it turns out that there are differences between the cases. In the first case, most of the people made the decision almost immediately, and there does not seem to have been a conflict; most people would switch the track.164 By contrast, in the second case involving the bridge, the decision was usually slower, as there is a conflict, and in the end very few people made the decision to push the fat man off the bridge. The accepted explanation is that emotion affected cold calculation in the second case, as I myself must push a man down to his death (like in the case of the baby described above), while the first case involves a more neutral action. The brain scans in Greene’s studies show that the first case involved only the DLPFC region (the dorsolateral prefrontal cortex), where cost-benefit calculations and decisions are made, while in the second case there is a dilemma against the more emotional DMPFC region. The dilemma between the intellect and emotion is decided by a region called the ACC/MSA. I will note that researchers tend to assume as obvious that these are two equal cases at the rational level, and that the difference between them is solely emotional. But of course it may be that there is also a difference at the logical level, as well: There is a prohibition on actively murdering someone, and this is different from a situation involving a passive act of not rescuing someone. The assumption at the basis of this approach is that there is not just a teleological prohibition to murder, in which the prohibition is judged only according to the results — i.e., how many people died — but also a deontological prohibition, meaning a prohibition applying to the very act itself. This is primarily relevant for the general philosophical question: Do people decide moral dilemmas based on logical or emotional considerations? Greene’s study showed that in the second case, emotions were involved in the form of neural activity in the DMPFC regions. In another study by Greene and Paxton,165 the researchers tried to see what makes people demonstrate fairness and integrity, even in cases where they might lose out from it: Is it a policy related to resisting temptation, or perhaps the lack of temptation?166 To test this, he placed a group of test subjects in a brain scanner, and asked them to predict the results of a coin toss, with a correct guess earning them a cash prize. From another group, he asked them to report if they predicted correctly, and he paid them according to correct predictions according to their report. It turns out that the second group had a high number of correct predictions (above 50% of statistically expected results), meaning that many of them lied to win the money. It turns out that among those who lied, there was an increase in the limbic activity in their prefrontal cortex. The intensity of the activity reflected the amount of lies. By contrast, no such increase was observed among the fair subjects, and the conclusion is that they didn’t feel the temptation to lie, as they did not feel a conflict. Moral dilemmas are of course only an example of the decision-making process. We could describe any process of decision-making in a similar manner, as a combination of brain activity in a number of regions (sometimes it represents a conflict) ending with a decision, which for neuroscience is nothing more than the mechanical integration of all brain activity which produced electrical signals or series of spikes to the organs to carry out an act.

Summary In this chapter I provided a schematic description of the manner in which the brain operates. We saw that in light of the knowledge accumulated until today, it is fairly clear that most if not all of our mental actions can be described in terms of various neural activities, meaning in physical and bio-chemical terms. This picture arouses a number of questions, and effectively takes the philosophical questions we have dealt with so far and gives them greater focus. I have already noted that some argue that neuroscience effectively takes these philosophical questions and makes them scientific ones. We must now return to the question of the relationship between body and soul, otherwise known as the psychophysical problem and the question of materialism. We must also now reexamine the question of determinism: Do we indeed choose freely, or is that choice a deterministic physical action? To sharpen the matter further, the question is how is all this physical-chemical description related to the mental level? What is the connection between the electrical activity we described here and moral decisionmaking, emotional experiences, or the identification of Grandma? We also have the question of whether or not the mental has an independent existence and whether we truly make decisions, experience, and feel, or perhaps these are only physical and entirely mechanical acts done in our brain? Perhaps our consciousness and feelings are byproducts or epiphenomena lacking any practical significance? All these interrelated questions will be dealt with below.

151 See chapter fifteen on the split brain phenomenon, in which this section is cut, dividing the two hemispheres from each other. 152 A fascinating example of the role of the two hemispheres is given by Dr. Jill Bolte Taylor, a neurologist and brain researcher, who herself suffered a stroke which damaged the left hemisphere of her brain. She describes her experiences and how she conceived herself in this state with complete consciousness; See her TED lecture: https://www.ted.com/talks/jill_bolte_taylor_s_powerful_stroke_of_insight 153 See the post here: orenmada.net/archives/59904 (Hebrew) for the pictures in this text. 154 Although the mirror itself also reverses the picture. See the discussion above in chapter five. 155 There is one prominent difference: In computers, the series of numbers is important. There is no significance there to the gaps between the different numbers and the timeline plays no part. By contrast, in the brain the form of the series along the timeline is also important, not just the series of numbers in itself. 156 The phrase “in order to” is inaccurate here. The visual information does not contain analyses of intentions and goals of what stands before me. It is simply visual information, meaning some picture. Only the analysis done with that information in the brain produces the conclusion that we are dealing with the lifting of a hand in order to strike me, leading to the decision of how I must act as a result. For the materialists among us, even this meaning of “in order to” is incorrect, of course. 157 This is the materialist-determinist description of the process. Later, I will distinguish between it and the libertarian description. 158 I will note that this is also true when it comes to learning a new technique or skill, and not just learning new information, as learning a skill is nothing more than storing information. Such an action therefore involves changes in the structure of the network, although this is of course done in different regions in the brain. 159 Studies are being done right now in which researchers are trying to reconstruct a picture from a brain state, and even dynamic pictures (meaning a movie, not a static picture). For a report on an interesting study on this subject: http://www.youtube.com/watch? feature=player_embedded&v=nsjDnYxJ0bo 160 For a description of Greene’s experiments, see, Joshua D. Greene, R. Brian Sommerville, Leigh E. Nystrom, John M. Darley, Jonathan D. Cohen, “An fMRI Investigation of Emotional Engagement in Moral Judgment,” Science, 293, September 14, 2001, pp. 2105-2108. 161 I used the term “can justify” and not “justifies,” as the assumption that this is indeed the logical conclusion is not necessary. In halacha, for instance, deciding such a case is not simple. The question is whether the baby has a status of “pursuer” [i.e., is an active threat to someone else’s life], or whether it is simply a factor and the pursuer is the enemy. If the second possibility is correct, then some believe then there is no warrant to kill the baby, as it would be the murder of a third person to be saved (saving himself with the life of his friend). On the prohibition on murder (of those who are not pursuers) there is an absolute duty: be killed and do not kill. I might add that in some of my articles, I disagreed with this Halachic stance. 162 In the above mentioned article. 163 The two track experiment was proposed by philosopher Phillipa Foot in 1967, and the experiment of the fat man on the bridge is an expansion done by Judith Thompson, and both have since been endlessly discussed in different variations. For instance, the dilemma of the bridge in the case that the fat man is the one who tied the people to the tracks, is called the “dilemma of the fat villain.” When it comes to halacha in that case, see the two opinions in the Minhat Hinuch, Mitzvah 440, article 1, and Or Sameach, on Hilchot Rotzeach, 1.8. For a general halachic survey of the quantitative question in life and death laws, see J. David Bleich, “Sacrificing the Few to Save the Many,” Tradition 43 (1), Spring, 2010, pp. 78-86. 164 It’s interesting to see the live “experiment done by Dr. Andy Thompson during his conference at the AAI atheists conference in 2009. See: http://www.youtube.com/watch?v=jnXmDaI8IEo&feature=related from 16:00 onward. 165 Published in the PNAS, 2009. 166 He calls the first possibility the hypothesis of will and the second the hypothesis of grace. I am not sure I agree with these definitions and will therefore not use them here.

Chapter Thirteen BRAIN-PSYCHE RELATIONS: EMERGENCE A silent, soft voice — …And I heard a voice coming within the silence … And the actual voice is not heard. (Rashi’s commentary on 1 Kings 19:12) Is there silence at the end of the storm, there is light hidden in the happiness that was hidden!” — Yehiel Moher, “In light of the memories” In chapter eleven, I argued that materialism does not provide an explanation for mental phenomena. What do materialists say in response? Many of them find the desired materialist explanation in the picture proposed in the previous chapter — a physical-chemical description of the totality of brain activities. In effect, they translate the mental world into physical-chemical terms. This is ostensibly a purely materialistic description, as the mental aspect has no part in it. So how is the mental related to this, if at all? Is there even such a thing as a mental layer? Does the picture arising from neuroscience mean there is no mental activity and that everything is material in the form of physics and chemistry? With this simplified formulation, the answer is pretty clear: It’s not reasonable, as we experience mental activity in an unmediated manner. We don’t experience electric currents but desires and yearnings, loves and hatreds, thoughts and emotions. We already saw that identifying all this with neural mechanisms is a conceptual error. What, then, is the relationship between this activity, which is so vividly and clearly present to us, and the physical-chemical description we presented last chapter? This chapter of the book examines how the concept of emergence tries to deal with this important question. This term, which is occupying more and more space in discussions on determinism, primarily in the context of the neurosciences, was originally an umbrella term for a collection of phenomena handled by a developing field in physics and mathematics known as the study of complexity. There are quite a few researchers from this field who question the accepted assumption that all the phenomena in our world can be reduced to the laws of microscopic physics (meaning, in the terms we used in chapter seven, that physics is modelable). The field of complexity deals with complex phenomena which emerge in some way from the laws of microscopic physics, but which cannot be based on them in a simple manner (see our discussion of unmodelable chains in chapter seven). This chapter can thus be seen as a continuation of chapters nine and ten. These two chapters dealt with the possible existence of gaps in physics based on looking at two central fields in today’s physics (chaos and quantum theory), while the present chapter asks if there is another form of gap, even if all the laws and fields of physics are hermetically sealed. Many materialists claim that it may be that mental phenomena emerge from within physics, even if the laws of microscopic physics do not allow for cracks to hide behind, as we saw. We will examine that argument in this chapter.

Reminder: The Two Alternatives In the wake of the discussion in part two, we arrived at the conclusion that there are two significant alternatives, one of which we must choose: Interactionist dualism or determinist materialism: Interactionist dualism sees the human being as a combination of two substances, matter and spirit, which interact with each other. The mental occurs in the spirit, and it is related to the totality of its traits and its characteristics, while the physical is an issue of matter. The assumption is that there is mutual influence between the two. A wound hurts me, and this is influence from the physical to the mental. By contrast, depression leads to physical weakness, and sometimes even an organic disease. This is influence from the mental to the physical. The question which arises here is how this picture fits with what we described in the previous chapter. Is the dualist doomed to reject the findings of neuroscience?

Determinist materialism, by contrast, sees the human being as an entity composed solely of matter. The mental is but a byproduct of the material, not a separate substance. As far as he is concerned, it is clear that the description given in the previous chapter contains all the relevant information for understanding the human psyche, even if it is sometimes difficult to extract from there. Here the opposite question obviously emerges: How does the picture emerging from the neurosciences sit with the subjective manner in which we experience the mental? What is the nature of the relationship between the neural and the mental? The present chapter is no more than a more detailed description of these two models, equipped with what we know from neuroscience. In this sense, we are to an extent repeating chapter eight, except that there we dealt with the issue from the philosophical and a priori perspective, while here we are dealing with the issue from the perspective of neuroscience. The important question accompanying us throughout the book, and especially from this chapter onward, is whether what we know from the neurosciences changes anything in the picture we have described so far. From here on out, I will try to show that as opposed to the repeated claims of neuroscientists, neuroscience does not add anything tangible to the discussion; it at most focuses it a little more. It was and remains a matter of philosophy. This does not mean that this will always be so, but this is the situation as of right now.

A Common Error: Freedom of the Gaps There is a common fallacy among creationists in the debate on belief in God VS evolution, where they base belief in God on gaps in scientific knowledge. Their basic argument is that if evolution does not explain everything, then there must be a God. Some call this the “God of the gaps.” In the third chapter of my book God Plays Dice, I explained that this is a fallacy because God is not part of the scientific map and is not supposed to provide explanations in places science has yet to reach. After all, based on such an argument, God was much greater a century ago, and He has been shrinking the more our scientific knowledge expands. By the way, this mistake does not just characterize creationists, who try to prove God’s existence based on deficiencies in scientific knowledge, but also the other side of the debate — atheists and neoDarwinists, who reject His existence based on the accumulation of scientific knowledge. I explained there that gaps in scientific knowledge and scientific understanding need to be filled with scientific tools, not metaphysical hypotheses in either direction. A similar phenomenon can be seen when discussing free will. Some wish to prove the existence of free will based on the fact that our scientific knowledge of the brain is not complete. One could call this the “freedom of the gaps” argument. This is obviously also a fallacy. Deficiencies in scientific knowledge should be filled by research and not spiritual hypotheses of any sort (even if I think they are correct). The psyche and the soul are not supposed to serve as scientific explanations. It’s important to understand that the fallacy once again emerges on both sides of the debate. Just as the libertarians and dualists cannot prove freedom and the existence of the psyche based on a lack of scientific knowledge, so determinists and materialists cannot refute this argument based on scientific knowledge and results. Only if the gap appears unbridgeable in principle, or that it appears to be unfillable by scientific tools based on what we know today, is there perhaps room for such arguments. As I explained already in chapter eleven, we should not identify the physical-chemical activity with mental events which accompany it. The flow of electrons and blood in the brain in regions which operate while identifying Grandma, purely physical processes both, are not the actual identification of Grandma, a mental process. Thus, neuroscience does not at all deal with the explanation of mental phenomena, and there is therefore no competition over the same area of discussion. At least as of now, there is no indication that we will ever have a materialist explanation for all these phenomena in the future. The argument that such an explanation will emerge is at most an expression of hopes and dreams, not facts or scientific findings. Take note, we are not talking about the lack of an explanation or scientific knowledge, but rather the complete lack of a conceptual foundation for such an explanation. It’s simply not the same field. This is like the argument that we so far do not have a literary explanation for chemical processes, but this does not mean that we will have no such explanation in the future; as of now, these are two entirely different and separate fields. To demonstrate this, I will use the famous example of John Searle from his book Mind, Brains, and Science.167 Searle proposes a thought experiment which he calls the Chinese Room.168 Imagine a person

who only speaks English, sitting in a room with two windows, an input window and an output window. From one window, the input one, he is sent questions in Chinese, and he is supposed to send answers in Chinese through the second, output window. He has an innumerable inventory of Chinese characters, and since he does not understand a word in Chinese, he simply chooses and composes them randomly to form his answer, and then pushes it out the output window. Let’s assume that this person has an infinite amount of time, and every time he gives an incorrect answer, he gets an electric shock. When he answers with a good answer which is to the point, he gets a candy. After enough time, a situation may emerge in which he has learned by a process of elimination how to give a proper answer to any question he is asked.169 Now Searle asks: Does this person actually know Chinese? I believe anyone with intelligence would not answer in the affirmative.170 Searle explains that this example distinguishes between the semantics or meaning and the syntax or form of the conversation. Syntactically, this man understands Chinese. His manner of action and that of a person who speaks Chinese is identical — they produce the same sentences. But when it comes to semantics, it is clear that he understands not a word. His physical actions may appear like he is having a conversation, but they are not accompanied by the mental events of understanding. The same is true of the relation between the neural level and the mental level — even if the neural always accompanies the mental, it is incorrect to identify the two with each other. Electric currents are not understanding, and correlation is not identity. By the way, it is interesting to note that Searle’s experiment now changes as far as neuroscience is concerned. After all, the measuring tools used today in such a situation would certainly identify that this person is not activating those regions which a real Chinese speaker would use while producing the same sentences. He can for instance write a sentence which expresses a strong emotion but of course not feel that at all.171 But all this of course does not change a thing when it comes to the distinction itself between correlation and identity, which remains correct. This is another demonstration of the fallacy in identifying the physical or neural activity with the mental activity which accompanies it. Therefore, at least in our time, there is no real basis that the neurosciences will ever be able to provide a neural explanation for mental phenomena. One can always hope, of course, and one need not necessarily have a real basis for hopes, but arguments based on hope and nothing more cannot decide such questions.

The Cerebral and the Mental: What Is Their Relationship? Still, there is room to ask whether the relationship between this neural activity and mental activity is like the relationship we described between the electromagnetic wave (physics) and light (qualia). Is this a relationship of cause and effect? If so, then it is still a case of determinism. If cerebral activity is the cause creating the mental events, then they are still processes entirely controlled by the laws of physics, as the nervous system is a physical system, and we already saw that there are no gaps in physics which allow for freedom of choice. Moreover, in chapter five we saw that the cause is a sufficient condition for the effect, meaning that given a specific brain state, the mental state is necessarily derived therefrom. Based on the assumption of determinism, there cannot be two mental states which come from the same brain state. It is true that in such a state we can still speak of dualism, meaning the existence of two substances of matter and spirit, but the influence between them is now one-way, from matter to spirit. Such a relationship between the cerebral and the mental expresses an entirely deterministic dualism, what we called in chapter eight epiphenomenalism. Another possibility is that these are two sides of the same coin, which we called double aspect theory in chapter eight. Neural activity is a physical phenomenon, and the mental is simply the manner in which we conceive these physical events, like in the case with the electromagnetic wave and the light. This is once again a deterministic picture, as one of the sides of this coin is pure physics, and the other is but its expression in the consciousness of the observer. This agreement is almost entirely equivalent to the previous one, and is therefore problematic for the very same reason. In chapter eight, we also pointed to the lack of definition of the “we” who observe these events. In what consciousness are these cerebral events being seen as mental events? Who is “conceiving”? We still seem to be implicitly assuming another, non-physical substance, if not in the conceived object, then at least among the conceiving subject. A third possibility is that the mental is but the global traits of the material system itself. There is no second, non-material substance in that case. Wanting, thinking, love, and hate are but traits of the collection of molecules which constitute our brain, just like the trait of mechanical elasticity or plasticity of the brain, which is also a global trait of this collection of molecules. This is the approach of emergence, which speaks of traits of a complex system which emerge at the collective level, without them actually existing among the items which compose this system. This is a very common approach among

neuroscientists today, and it also obviously assumes determinism. These questions return us to the psychophysical problem, but now the discussion is not a priori but a posteriori, conducted in light of the new findings of the neurosciences.

Reminder: The Neural-Mental Axis As we said in the previous chapter, at the same time as physical-chemical events occur, a different sort of events occur by us: Mental events. When the electrons flow within the axons, from the sensory receptors to the center which does the initial processing in the brain, a picture of a woman is also created in our consciousness at the same time. When the initial processing center sends the product which comes out of it to the second processing center, which is more complex and sophisticated, our identification of Grandma is also created by us at the same time. In the third stage, an order is sent from the brain to the motoric organs to move towards her and give her a kiss, what we experience as love and a desire for contact. Alternatively, we have an experience of fear which is created within our minds, accompanied by a motoric order to the organs to run quickly. In both cases, the cerebral processing ends in that the motoric action is done by the organs. The same is true when it comes to decision-making and deciding moral dilemmas. At the same time as the complex processes occur, which we just described, activity takes place in the mental part of our mind which we usually call moral deciding, or decision-making. When we say “moral decision” we do not mean a collection of electrical currents but a mental activity which accompanies these currents. The question we are dealing with here is the question of the relationship between these two levels: The neural and the mental, or physics and qualia. The dualist answer to this question was fleshed out in chapter eight. We pointed to the mental being the collection of traits of the spiritual-psychological part of the person, and that the connection between these and the material dimension is two-directional; there is mutual interaction between these two components of the person. The conclusion from the discussion in part two of the book is that one cannot speak of libertarianism without dualism and the assumption of mutual influence. We saw that this means that physical phenomena can be caused by mental events — will as a mental phenomenon can move electrons, without any preceding physical cause. The alternative is materialism, but here we will deal with the sober variant, which does not deny the mental or confuse correlation with identity, but simply bases the mental on the material. It’s very much like what is known in philosophical discussion as epiphenomenalism. Epiphenomena are an auxiliary phenomenon, meaning that the mental is an appendage to the material. This usually means that there is a spiritual substance, except that it is operated by the material and does not act upon it. But in recent years, the approach which entirely forgoes the psychological dimension as another substance has become more and more popular; an approach according to which the mental is a collection of collective traits of the complex material micro, and the mechanisms which create it are called emergence.

Emergence John Searle, in his book Mind, Brain, and Science, explains the emergent picture (even though the term itself was not yet in use) with an example. His basic claim is that when we observe a complex system, there are cases in which traits which do not at all exist at the microscopic level suddenly emerge at the macroscopic one. He brings as an example the trait of liquidity. A water molecule is not liquid. In fact, all the traits related to states of matter such as solidity, liquidity, and gaseousness cannot characterize a single molecule or atom. These are the traits of a collection of molecules, and they describe the nature of the relationship between the molecules in the collection. A collection of water molecules can be liquid in the form of water, gas in the form of steam, or solid in the form of ice. A single molecule is none of these things, as these traits are not at all defined at the level of the individual molecule. In such a situation, we say that liquidity is a trait which emerges at the macroscopic or global level and does not at all exist at the microscopic level. Searle’s argument is that mental events — such as wanting something, identifying someone, or love towards someone — as well as our psychological skills such as the ability to want, think, love, or identify, or traits liked generosity or its opposite do not exist within a different substance called a soul, but are rather traits which emerge from the material collective itself. The collection of molecules which creates our brain is graced with mental macroscopic traits. In other words, when we connect an identical collection of molecules in the same manner and clone the human being, it will be a living, mentally active human being for all intents and purposes. It will not just be us who think that he has mental traits, but the collection itself will see itself as such. This means that human beings have nothing beyond the collection of molecules, and

mental events are nothing more than traits of that collection of molecules, nothing more. We already saw that a trait of something is not an addition to it — the color of a table is not a separate entity from the table; the only entity is the table, and it is characterized by the color brown. According to the emergent approach, the mental events are also a trait of the material collection and not an additional entity. If there is some collection of water molecules at room temperature in some vessel, a liquid will form in this vessel, and there is no need to add a soul or any other substance for this to happen. This is the case, argue those who support the emergent approach, with the collection of neural molecules which are gathered in the form of a brain, which creates wanting, generosity, religious faith, bravery, scientific thinking, and so on. Here, too, there is no need to assume the existence of an additional substance in the form of a soul. This is the idea of emergence in a nutshell. Its power lies in its ostensibly providing an explanation for the appearance of the mental without denying its existence or confusing correlation and identity, and without having need of the metaphysical assumption of the existence of another substance. Truly, this is the realization of the dream of Razor creator William of Ockham. We will better understand the fundamental significance of the idea of emergence if we make a distinction between the various levels of integration at which one can refer to the same system. We can, for instance, describe a computer in terms of the movement of electrons within it, and this is the physical level, which is the most basic (there are of course even more basic levels, such as those of the elementary particles which make up the electrons and the atoms). The level above this is the electronic level, meaning the level of the transistors (or logical gates or circuits, which are additional, higher levels of integration, within the electronic level). Above this is the functional level, meaning the blocks which describe the various functions of the computer such as input/output, processing, memory, and the like. Which of these descriptions is the most correct? Obviously they are all correct to the same degree, but they are done at different levels of integration. However, a description at the physical and electronic level will not help us when coming to understand the activity of the computer. Without the functional explanation, it will at most be a collection of fields and particles which move in all sorts of directions. Without the macro picture, meaning the description in broad brush strokes, we will not be able to understand what a computer is. It’s important to understand that the laws which govern a computer’s behavior at each level are entirely different: In physics, we speak in terms of field theory and elementary particles, and then in terms of the laws of physics (the laws of mechanics and electromagnetism). At the electronic level, we think in terms of the influence of the laws of electronics and electric circuits. In the language of logic gates there are the laws of logic. At the functional level, the language is that of hardware and software engineers, and so on. This means that the laws at the micro level are not simply expressed at the macro level. It’s hard to know what we will get at the macro level based on the laws at the micro level. The type of behavior at the macro level, and even the terms which the macro language is composed of, are entirely different. What is particularly interesting with this idea in the neural-mental context, aside from the fact that it offers a more reasonable platform for materialism, is that it’s not entirely clear to those who deal in the field that emergence necessarily leads to determinism. There are quite a few thinkers and scientists who take this idea one step further, arguing that the idea of emergence not only allows us to accept mental phenomena within the materialist picture of the world, but it also allows us to even hold to libertarianism alongside a physicalist picture. Their argument is that just as the mental emerges from the collection of material, so does freedom emerge from the laws of physics. True, at the microscopic level, there are no free gaps within physics, but at the macro level new behavior emerges, operating according to new and entirely different laws, which may also contain elements of freedom. This is another attempt, the third in number (after chaos theory and quantum theory, which were discussed in the previous part), to introduce freedom into a picture which is effectively physicalist. The reader will certainly not be surprised if I continue my McCarthyist ways in this chapter and hunt this witch once again. I will demonstrate below that at least based on our conceptions today, this attempt fails, and does not seem like it even has the potential to succeed. There is no possibility to show freedom within a physicalist picture, even if one enlists the idea of emergence for that purpose. My argument now is that the conclusion that a libertarian must be a dualist remains with us for the duration. To understand this better, we will have to go a little deeper into another field in physics — statistical mechanics.

Statistical Mechanics in a Nutshell The idea of emergence was imported to philosophy from science, where they’ve been dealing with the idea for quite some time, albeit not always explicitly. I already noted that in the physics of recent years, emergence is primarily discussed in the context of the study of complex traits. But there is another, older field of physics where the main subject is emergence and its characteristics, and this is statistical

mechanics. This field deals with the relationship between the micro and the macro, and especially the emergence of traits at the macro level from the details belonging to the micro (and for whom there is sometimes not meaning for these traits, as in the case of liquidity we encountered above). Statistical mechanics was effectively born as a microscopic explanation of the field of thermodynamics, which deals with the traits of large systems at a temperature different than absolute zero. Traditional thermodynamics deals with traits like states of matter, pressure, and energy. In order to do so, other characteristics of these systems are defined such as entropy, enthalpy, and others, and thermodynamics describes the relationship between these traits. Statistical mechanics shows how these macroscopic traits emerge from the microscopic structure of the system. For instance, take a system of ideal gas. Ideal gas is a collection of atoms with no interaction between them, all of which are trapped in a container with a given volume at some fixed temperature. Even before they thought to invent statistical mechanics, thermodynamics already noted that if we are given the volume V and the temperature T, we can know the pressure within the gas represented by P, which is also the pressure which will act on the walls of the vessel (ideal gas equation: PV = nRT).172 Now comes statistical mechanics and explains this equation. It refers to gas as though it was a collection of ideal atoms, without any force acting on them, analyzes their microscopic traits — speed, location, energy, and so on — and effectively determines the distribution of these traits (how many atoms will be at each level of speed or energy). Based on this, statistical mechanics defines and derives the macroscopic traits of gas, arriving at the ideal gas equation. This is the explanation of the macroscopic trait or thermodynamics of gas in terms of its microscopic traits (the Newtonian laws which govern the movement of the individual particles). It’s important to understand that at the microscopic level, only the regular laws of physics or Newtonian laws apply. The concepts which apply there are speed, location, acceleration, energy, and others. There is no room for concepts such as temperature or entropy. These traits only emerge at the macroscopic level, much like liquidity in the example above, while there is no significance for traits like speed or location at the macroscopic levels. Gas as a collection has no location or speed. Thus, statistical mechanics is not just a tool for analyzing the state of a collection of atoms and deriving some traits of the collection. What is more important is that it also defines the macroscopic variables (sometimes called intermediate concepts) such as entropy, pressure, and enthalpy in terms of the microscopic traits of all the atoms. Thus is pressure derived from the sum of all the forces applied by every one of the atoms on the walls of the vessel. But it is not just pressure which is derived in this way. It turns out that a trait like temperature, which has no meaning at the microscopic level, is also a trait which can be defined and calculated through a sum of other traits of the collection of atoms. Thus does statistical mechanics base thermodynamics, which deals with the macroscopic system — ideal gas, in our case — on the traits of the microscopic system (in our case, a collection of atoms). This is, in effect, an emergent explanation, which shows the emergence of macroscopic traits from within the microscopic picture where they do not exist at all. The example of liquidity is a good example of statistical mechanics, as states of matter, and the movement between them (phase transitions), are a very important research area in this field of statistical mechanics. A liquid physical state is effectively the state of this collection of atoms at temperatures which are not too low. When we reduce the temperature, the physical state can change and the collection of atoms becomes solid. There is a third physical state at higher temperatures — gas. The move to it is also a phase transition, and it also expresses a change in the distribution of the microscopic traits of the collection of atoms. Again: Nothing definite happens to the individual atom in these phase transitions. It does not “know” what temperature is. As far as it is concerned, there are forces which act on it, speeds and accelerations, and in effect — Newtonian laws. What changes are the macroscopic traits of the entire collection (the state of matter). Just as an imprecise example, a lower percentage of atoms are at high speeds in a solid state. Thus do the sums of the microscopic traits such as speed change which determine the macroscopic traits such as states of matter. The trait of liquidity can also be explained by the simple laws of physics. There are different kinds of atoms in nature (which all appear in the well-known periodic table of Mendeleev). Surrounding every such atom is a force field, which means that if another atom encounters it, it will feel the force acting on it. Obviously, the other atom will also apply its force field to the original atom. These forces cause attraction or repulsion between the atoms, and they may thus end up clinging to each other in different and various ways, and such stable bonds (collections of atoms connected to each other in a rigid connection) are called molecules. The character of the molecule is of course dependent on the traits of the atoms which compose it.

By the very nature of things, each molecule or atoms is surrounded by a force field (in the case of the molecule, it is the collection of forces of the atoms which compose it), and anyone who encounters it feels it. The scale of distances between the molecules ( which is bigger than the scale of distances between individual atoms which compose the molecule) which is temperature dependent, determines whether a collection of molecules is liquid, solid, or gas. The same is true of a collection of atoms. In this manner, the traits of atoms, and from them the traits of molecules, effectively determine the macroscopic traits of the collection. If we know that the force field surrounding molecule type X is constructed in such a way that it traps another type X molecule at a regular distance but with a different orientation, then it is clear that such a collection of molecules can create liquid. By contrast, a force field which traps the other molecule in one specific place, fixes the collection of molecules to create solid matter (solid materials can be different from one another, in their crystal forms. These are dependent on the traits of the atoms or molecules which create them).

Consequences for the Neural-Mental Relationship: The Conceptual Aspect and the Scientific-Computational Aspect The argument of emergence, when applied to the relationship between the neural and the mental, effectively says that mental events and our psychological traits are nothing more than the global traits of the physical-chemical system called the brain. In other words, if we build a brain which is biologically identical to ours, it will produce the mental phenomena we call consciousness and psyche. For this, we must go back and distinguish between the global traits of the brain as material and the mental events and traits themselves. Searle’s argument, that memory is a global trait of the neural network in the brain, can be interpreted at two different levels. At the first level, the neural activity which accompanies memory is a global trait of the neural network. Any memory, or collection of memories to be more precise, is represented in the brain by a specific global neural state. This is apparently a correct answer, and today we have a pretty good understanding of how it works, how the information is stored — learning, by the process we described in the last chapter — and which neural state fits all the information stored within it. At the second level, memory itself, as a mental event (meaning what we experience as remembering, the qualia, not the physical-neural events which accompany it), is a global trait of the neural network. This argument lacks any physical basis, and is effectively nothing more than a conceptual error, as I will now explain in further detail. We saw in the example of liquidity, or ideal gas, that we have a systematic and splendidly constructed scientific description of the move from the microscopic to the macroscopic, and this at two levels: (A) The computational aspect — via an organized calculation of these traits for a particular collection (hydrogen atoms or H20 molecules of water and so on); (B) The conceptual aspect — the definition of the macroscopic traits in terms of their microscopic traits (i.e., the pressure of the gas is the sum of the forces acting on the walls, with the conceptual connection between the force and the pressure being obvious — the force created by the atoms hitting the walls creates a phenomenon of pressure on the walls; liquidity is the clear result of the character of force fields of isolated molecules, and here too the conceptual connection is clear; temperature is some sum of the energies of the molecules or atoms). These two aspects, both the mental connection and the calculational connection, do not at all exist at the neural-mental level. Start from aspect A, the computational aspect. We have no way at the neural-mental level to calculate a mental state from the sum of neural network states. Again, we can in principle calculate a neural state which is accompanied by a mental state of depression via such a sum. But the mental state of depression itself cannot be calculated in this manner. This is like the attempt to calculate the formation of the experience of the color yellow in terms of the traits of the electromagnetic wave which creates it (plus the traits of the neural network which processes this visual information). As of now, there is no way to do it. We simply say that the color yellow was created when an electromagnetic wave of such and such wavelength hit our cornea. There is no description here of the process during which this is created from that or an explanation for the process, but only the establishment of a correlation between the two. We do not say that it is a trait of the wave hitting the cornea; the relation between them is one of cause and effect, and they are therefore two different things. To say that they are identical is therefore a conceptual error. Just for comparison’s sake, the wavelength of the yellow color is a physical trait of the wave itself. We can measure it with tools and see the wavelength of the field reaching us. We are therefore not dealing here with cause and effect but with a trait (wavelength) of some subject bearing the trait (the wave itself). Noone will say that the wave is the cause of its wavelength, just as a person is not the cause of his height. Height is a trait of people, and the same is true of the wavelength of a wave. Is this also the relationship between the neural and the mental? Here it is pretty clear that we have a relationship of cause and effect,

like in the case of experiencing the color yellow and the wave which created it (and not like between the wave and its wavelength). But if the relationship here is causal, then the natural conclusion is that these are two different things, that one is the cause of the other, and not the relationship between an entity and its trait. (It’s worth remembering Taylor’s argument in chapter eight regarding statements like “My body is sad” or “My body believes that it’s nighttime now.”). At the neural-mental level, by contrast, apparently to avoid dualism, we are for some reason told to identify between cause and effect and see the relationship between the neural and the mental not as a relationship between cause and effect but one in which the mental is a trait which emerges from the neural. So far, we dealt with aspect A, the ability to calculate the mental from the neural, and saw that this is impossible. Now we will go on to aspect B, the definition of terms. We saw that in statistical mechanics, we accept the emergent explanation because it gives us a way to not just effectively calculate but also define the macroscopic traits based on the microscopic traits. Temperature is some calculated sum of energies of all the particles. The same is true of pressure calculated as the sum of all those forces, and other macroscopic traits as well. Even if we didn’t know how to effectively calculate temperature or pressure, the conceptual connection between the micro and the macro is clear. Everyone understands that pressure is some sum of the forces applied by the particles on the walls of the vessel, or that liquidity is the result of forms of force fields between molecules. The same is true of temperature, a concept which simply does not exist at the microscopic level, which can be defined and calculated in terms of the physical traits of particles, just like states of matter or other values of intermediate concepts. By contrast, we do not even have the beginning of a direction to define and certainly not calculate macroscopic mental traits at the neural-mental level in terms of traits of neurons or currents in the neural networks which connect between them. These are two entirely different conceptual worlds. A mental state of depression is not electric currents, and we do not even have a conceptual system which hints at the possibility of translating the psychological conceptual system in one way or another into physical terms. How do we define depression as the sum of the traits of neurons or the electric currents between them? Again, the problem is not just that we don’t know how to calculate it. We dealt with this when we discussed aspect A above. Here we are dealing with the very definition of traits themselves. There is no way to define depression in terms of neural states and electric currents, and therefore there is certainly no hope to do a calculation which would produce the mental state of depression as a result. If so, the comparison between the emergence of states of matter or intermediate concepts from the microscopic traits of particles and the emergence of mental states from the physical traits of the neural network, is baseless, both scientifically (we have no way to calculate it) and conceptually (we have no way to define the conceptual connection between the micro and the macro). Like we said before, it’s like trying to propose a literary explanation for chemical processes. These are two different and separate fields, and the argument which identifies between them or connects the two to each other is both speculative and lacking any real basis. I found just such a point explained clearly in Shimon Marom’s aforementioned article in Odyssey 6: Something has happened in recent years; the picture has been overturned. Ostensibly, room has been created to discuss consciousness, emotion, individuality, and similar concepts in the framework of the proper discourse of neuroscience. These are not the retirement musings of prominent scientists or philosophers, but research subjects which occupy the long term projects of active researchers, who promise that the day is not far off when the insights of neuroscience will turn our spirit into an open book. Just how justified is this optimism of neuroscience in light of the mission which it has taken upon itself — to understand the origin of spirit from matter? What is the real status of the neurosciences? Not the declared, interest-driven, economic, which we (yes, also the present author) direct towards the ears of potential donors, government officials who are in charge of research budgets, or science journalists thirsty for colorful and catchy formulations? I fear that to the extent that it is directed to understanding the origin of the spirit from the material, neuroscience has failed the most basic test, a test which any field of knowledge called science should meet. The neuroscience of our day lacks agreed upon criteria for determining the relevance of its findings; this science possesses no tools, with which it can declare a finding relevant or irrelevant to understanding the origin of spirit from matter. Elsewhere in the article he writes:

It seems to me that the answer to the question of whether science in general, and the neurosciences in particular, really dreams of the ‘origin of spirit from matter,’ is — yes, certainly. It’s hard not to see the way in which the all-powerful neurosciences, reflected in the dream of understanding of the origin of spirit from matter, is joined at the hip with alchemy — a medieval science which combined chemistry, medicine, spiritualism, mysticism, and astrology under one roof; a science which was driven by the desire to turn simple and cheap metals into valuable gold.

A Similar Yet Different Example: The Problem of Vitalism The question of vitalism, which has accompanied biology and precedes the modern neurosciences, is similar: Does the appearance of life force us to assume the existence of another component beyond physics and chemistry which would allow us to describe the biological traits of a living being? Despite the similarity, this is a different problem than what we are dealing with here. A vitalist does not necessarily assume the existence of a soul. Vitalism is a scientific claim, according to which living matter is different than inanimate matter. The vitalist claims that living matter contains an additional component, which is described not by physics or chemistry but the laws of biology. But this can be an entirely material component, and it is not necessarily related to phenomena like experiences, emotions, thinking, or will. We are not dealing here with the mental but with the appearance of biological life. Even a living creature without any mental phenomena (like a plant) can awaken the question of vitalism. The common approach today in this context is that of reductionism, according to which vitalism is scientifically unnecessary, as we can base all the phenomena of life on chemistry and physics; we will at most have need of emergent processes which create the biological traits from within the microscopic laws of physics. In this sense, biology is a derivative science. But there is still a serious debate between philosophers and life science researchers on the question of whether we can effectively derive biological traits and phenomena from physics and chemistry. The reductionists of course say that it’s possible. According to them, just as chemistry is nothing more than the macroscopic traits of complex physical systems, which can be defined and calculated by applying the laws of physics to the particles which compose them and with the sum of all the particles, so too can biology be derived and calculated based on the laws of chemistry and physics. They argue that biology deals with the traits of complex physicalchemical systems of organic molecules, but we can at least in principle define and perhaps calculate the laws and concepts of biology based on the laws and concepts of physics and chemistry. And what we don’t know today is nothing more than a gap in scientific knowledge, which will be filled in by further research rather than by adding metaphysical assumptions. Those who do not believe in reduction claim that one cannot necessarily calculate the laws of biology based on chemistry and physics. It’s important to understand that this does not necessarily mean that they become vitalists and conclude that the living material must contain another component beyond physics and chemistry. They can argue that it is indeed impossible to calculate the laws of biology based solely on the laws of physics and chemistry, but this is only because biology emerges from physics and chemistry in a non-trivial manner. The biological trait is a trait of the macro; it may be connected to the micro which is only physical-chemical, but it cannot be explicitly calculated from that physical-chemical micro level. The basis for these kinds of opponents of reductionism is that at least at the conceptual level we can see a connection between biology and physics and chemistry, and there is therefore no need to assume vitalism. According to them, the fact that we cannot calculate the former from the latter is solely a problem of calculation, even if an essential one, and it is not a problem involving the lack of a conceptual connection between these two levels. This is similar to an extent to what we saw in chapter nine when we discussed chaos theory. We concluded that chaos theory does not allow a real gap within physicalism, and the reason for this is that the future state (the ultimate location and speed of the object at some point in the future) is entirely defined in terms of the present data (speed and location), but we simply have a problem in calculating it. The small ball which stands on the top of the round mountain will fall to one of the sides. The result is clear and can be defined entirely in physical terms. We simply have no practical way of calculating in which direction it will fall. The position which rejects reductionism on the basis of calculation difficulty argues that this is the case in biology as well. We have a calculating problem, not a conceptual problem. The concepts of the life sciences are logically related to physical concepts, there does not seem to be an unbridgeable conceptual gap here, and it is only the calculation which is difficult and perhaps impossible. Vitalism is a third approach, in addition to reductionism and the approach which sees the inability to do a reduction as solely a calculation problem. Vitalism argues that reductionism is also wrong at the conceptual

level; per this approach, we cannot even define the appearance of life in physical-chemical terms. The vitalist argues that the fact that we cannot base biology on chemistry and physics is not just a matter of calculation problems, as new concepts are created at the macro level which have no roots in the micro. Vitalism for our purposes is an analogy. In any event, whether vitalism is correct in the context of biology or not,173 in the context of neuroscience the present situation is that we certainly cannot define and base mental concepts on physical phenomena. Moreover, there is no doubt that the problem here is not just in terms of calculation, but it is a different semantic field. We simply do not possess the language to define mental concepts in neural terms. The idea of emergence, even if it is applicable elsewhere, has no basis in the context of the brain. Since there is a conceptual and not just calculational problem in the neural-mental context, there is almost no avoiding the assumption — parallel to vitalism in its general biological context, but not necessarily dependent on it — that we contain a spiritual-psychological component aside from matter, in which mental events take place and serve as its traits. This is especially so as we saw that the relationship between the neural and the mental is one of cause and effect and not one of an entity and its trait.

On the Lack of Observational Confirmation: Between the Soul and Raspberry Juice Our conclusion so far is that an observation of the relationship between the mental and the neural does indeed point to a correlation between the two. Mental activity is always accompanied by neural activity. But it is not reasonable to see the root of this correlation in one being an entity and the other its trait. We should not confuse correlation and identity. It is more reasonable to conclude that the relationship is one of cause and effect. We therefore conclude that the mental is apparently a different substance than the material, and that the two have a relationship of interaction (we have not yet discussed whether this interaction is mutual). As we already noted, nothing can interact with itself, and cannot be its own reason. If A influences B, then A and B must be two different things. Here’s an example: When we look at a cup of raspberry juice, someone can come and say that when such a number of water molecules fills the cup it becomes red. The red color is a trait of the water (the collective of molecules). It emerges from them and hence there is no need to assume the existence of another component in the liquid in the cup (the raspberry concentrate). It’s a possibility, sure, but in light of our knowledge of water it’s also an unreasonable one. There is nothing in the water molecules which hints at them becoming red when there’s enough water molecules gathered together. The color red is not a trait of the water molecules themselves, but the result of something else inside. Therefore, even without doing a detailed chemical analysis of the juice, the more reasonable conclusion is that there is another component (the concentrate) which makes the difference. It’s true that in the case of the drink, we can do a chemical analysis and prove that there’s another component in the water, a possibility which probably does not exist in the mental context. No analysis, chemical or otherwise, will produce the existence of an additional, spiritual substance. Ostensibly we see nothing but matter. But what we are looking for is not a physical entity, so it isn’t reasonable to assume that it will be observable, at least not directly, by physical means. We may be able to see an electron start to move, and notice that this movement was not preceded by the activity of any force field. But it is very hard to use scientific tools to arrive at the conclusion that there really wasn’t a force field there. Moreover, the principle of causality also leads us a priori to the argument that every event must have a cause, and we will therefore always assume that there was some cause there as well; we just haven’t noticed it. The materialist will assume that this was a physical cause he or his tools missed, but as I explained in chapter six, a spiritual cause is also a cause, or at the very least a sufficient reason. Therefore, observation may not confirm the existence of spirit or even the interaction between it and the material, but we also don’t expect it to be able to do so, and therefore the question of observation is not so important for our discussion. Those who hold to a dualist position shouldn’t have to change it because of the lack of observational confirmation, and those who hold a materialist position also have no need of such confirmation. This argument can be relevant only for those who haven’t decided one way or the other, yet. Moreover, emergence is also an argument which cannot be confirmed by observation. Those who hold this position are not making any sort of quantitative argument that can be empirically examined, and in fact present no mechanism, even without quantitative elements, which can translate the neural into the mental. Their argument is therefore no more scientific than the dualist one. They simply state that in a certain neural state a mental phenomenon emerges, and that’s it: No explanation how it is created, no definitions, and no ideas or even the beginning of a direction of how to get from the neural to the mental.

On Strong and Weak Emergence: Is There a Materialistic Libertarianism? So far we dealt with the question of materialism and dualism. The conclusion from the discussion so far is that the materialist position has need of mechanisms of emergence, but there is no hint to such mechanisms existing, and even the assumption that they exist at the conceptual level lacks foundation and is problematic. All the examples where use is made of the concept of emergence are fundamentally different than the example of the neural-mental. Therefore, the assumption that the mental is a macroscopic trait of neural material is an unreasonable argument and at the very least one lacking any foundation. I already noted that some want to go one step further and see emergence as a theoretical tool which doesn’t just allow us to recognize mental processes, but also reconciles materialism and libertarianism. Meaning within a physicalist world, in which all is matter and the laws of physics reign supreme, there will still be room for an approach which espouses freedom of choice. We rejected this possibility when it came to chaos and quantum theory in chapters nine and ten. But now we are dealing with another attempt to do so, an attempt based on the argument that even the laws at the macro level are not derived from the laws and interactions of the micro system. According to this proposal, there may indeed be no gaps in physicalism, but the macro laws nevertheless create freedom (it emerges at the macro level). This sort of argument is known as strong emergence. By contrast, the example of liquid as opposed to the molecules which compose it, is an example of weak emergence. In this case, the connection between the laws which control the liquid and the laws which control the molecules is clear, and therefore this is weak emergence. Weak emergence refers to situations in which we can base the macro phenomena on the laws which control the micro. In this sense, as we saw in chapter ten, Newton’s laws of physics are an emergent result of quantum mechanics, as the laws which control microscopic quantities are the sum of the microscopic traits which control the particles at the micro level. The argument of materialist libertarianism is that mental phenomena, and especially the freedom, are a result of strong emergence from the material. It argues that even though the laws of physics contain no gaps at the micro level, behavior containing elements of freedom is created at the macro level. Gazzaniga proposes such a division between kinds of emergence, and the example he brings for strong emergence174 is the behavior of traffic on the road, composed of a collection of vehicles, drivers, the weather, culture, and so on. We cannot, he argues, predict something about the traffic by observing a single vehicle. We need to observe it at a different, higher level of integration. This is his example of strong emergence, meaning the relationship between the mental and the neural. This example is incorrect, of course. It merely demonstrates weak emergence. If we know the behavior of each individual vehicle, we can sum those up and know how all the traffic will behave. The fact that there is interaction between the levels of freedom in the micro (the individual cars, the people, and so on) does not mean that the macro is not some kind of sum of all of them. The behavior of the vehicles is more complicated due to the presence of the other vehicles and other factors and it complicates the calculations, but this is a technical problem. Traffic is nevertheless still the sum of movement of all the individual vehicles. If we return to the various levels of integration in the computer example presented above, we can put it this way: The laws of Newton and electromagnetism can be derived from the sum of the physics of the elementary particles and the fields. The laws of electronics (the transistors) can be derived from the laws of physics. The laws of logical description (the activity of the logic gates in the computer) can be derived from the laws of electronics. The laws of computer engineering which control its functional description can be derived from the laws of electronics and logic. The conclusion is therefore that it is correct that we are not dealing with the same laws, but it is certainly reasonable to assume, even without direct and explicit calculations, that the laws at the higher level of integration are some sum of the laws at the lower levels. This is precisely the meaning of weak emergence. This is also what happens in Gazzaniga’s traffic example, and it can therefore not be a good example for an argument on the neural-mental relationship which requires strong emergence.

Why is the Argument of Strong Emergence a Weak One? The fallacy in the traffic argument is no coincidence. Gazzaniga presented an example of weak emergence since there is no good example of strong emergence. We can show pretty easily that we do not and cannot have any indication of strong emergence in any situation we could recognize.

The argument goes roughly like this: Say there is a certain collection of phenomena A at the macro level, and another collection of phenomena B at the micro level. The argument that there is strong emergence here includes two components: (I) The A phenomena are solely the result of micro level B, without any addition in the form of an added substance or additional laws at the macro level. (II) There is no way to derive A phenomena as any sum of B phenomena. But now we must ask the obvious question: If we cannot derive A from B, then how do we know that A really does emerge only from B? How can you show that there is no additional component involved in the laws and phenomena at the macro level? After all, if you did not succeed in explaining the macro based on the micro, the assumption that it is only based on the micro is speculative by definition. Strong emergence is thus always speculative by definition; there will never be strong emergence which can be demonstrated. When it can be seen, it will simply be a case of weak emergence by definition. This argument is reminiscent of the Gödel theorems of logic, according to which there are statements in some mathematical systems which are correct yet cannot be proven within that system. The theorem itself was proven, as it is indeed unprovable within the system, but a proof does exist outside of it. But we are dealing here with a situation where there is no way, inside the system or out, to derive the macro from the micro. So on what basis can we conclude that the macro emerges from the micro without any addition? To the contrary, it is more reasonable to say that if we cannot derive the macro from the micro, then the macro must have some components which do not exist at the micro level, and precisely because of this we cannot derive it from the micro (as in the Raspberry Juice example). Thus, the argument that the mental emerges in a strong sense from the material or the neural is by definition an argument lacking foundation.175 We can of course argue it as a hypothesis without justification, but we must remember two things: (I) It is not scientific, as it can be neither confirmed nor denied; (II) It usually relies on familiar examples of emergent connections, presented as support for the possibility of the argument being true, but all the examples presented for familiar emergent phenomena cannot be the grounds for the argument of strong emergence, as all of them are necessarily demonstrations of nothing more than weak emergence. Any example of strong emergence will be speculative to the same degree, and the emergent materialist can therefore not use it to base his arguments. In a sense, these examples even lead to the opposite conclusion: Every time there is an emergent relationship, we can derive the macro from the micro, and when we can’t there is no reason to assume that the relationship between the micro and the macro has characteristics of emergence.

Back to Statistical Mechanics: Multiple Realizability We will now take a step further with statistical mechanics. Above, I described statistical mechanics as a theory which helps us explain macroscopic traits via microscopic traits. I used the example of ideal gas, which is a collection of atoms not interacting with each other and which is subject to no external force aside from the walls of the container. We saw that statistical mechanics is a physical theory which defines the macroscopic traits of the gas via the microscopic traits of the collection of atoms. For instance, the gas’ temperature or pressure reflects a certain distribution of speeds or energies of the atoms which compose it. This picture is based on a mapping between two levels of reference: The macro — the gas in the container, and the micro — the collection of free atoms in the container. What is the nature of this mapping? Is it one-valued? Does every state of particles have one macro state which fits it? Yes. But does every macro state have one micro state which fits it? Certainly not. Think of a container with three free particles lacking interaction between them. (a, b, c) For simplicity’s sake, we will divide the container into two equal regions (x, y) as described in image 1:

Image 1: A Simplified Case of Ideal Gas in a Container. The gas has three particles, and each particle can have only two locations (in two equal areas, x and y). The arrows mark the movement of each particle, with the arrow’s direction marking the direction of movement and its length symbolizing its speed).

The energy of the gas is a collection of the energies of the particles. It is clear that there are infinite microscopic possibilities which are appropriate for a given quantity of overall energy. If the sum of all the energy is 10, that can be built from three energies like (2, 5, 3), (9, 0, 1), or (1.36, 8.14, 0.5) and so on. Meaning, there are innumerable micro states of particles which fit a particular overall level of energy. We should remember that the energy state of one particle also represents infinite particle states — as energy is only dependent on its speed but not its direction. And we haven’t yet spoken about the locations of the particles, which do not at all affect their energy. Every microscopic state of the system includes data on the location and speed (size and direction) of every particle. In our simplified model, the location is one of the two regions, and the movement can be any speed and any direction. A particular pressure state is a macroscopic state. The pressure describes some force which acts on some wall. What is the appropriate microscopic state? It’s very easy to see that there are lots of such states. For instance, if we are speaking of the wall on the right, then there is a state in which particle a is in area x and is applying all its force, and the other two are in area y and are therefore applying no force on that wall, but it may be that particle b is applying all its force or particle c, or two of the particles are applying force on the wall, which adds up to the general pressure, and so on. Once again, we are dealing with an infinite number of possible microscopic states which fit any given macroscopic state. This means that there is a phenomenon of multiple realizability. If we know the macroscopic state of the system, this does not mean we know all that much about its microscopic state (where each particle is and what their speed is). The macro state can be realized in many microscopic forms. We of course know something about the distribution of locations and physical traits of the particles, but nothing specific about this or that particle. But this is not the case in the reverse: If I have complete and absolute knowledge of the microscopic state of the system, there is only one, well-defined, macroscopic state which fits it. It may be that I cannot calculate the macroscopic traits of the gas in a given micro state, but this is just a problem of calculation complexity. It is clear that the macro traits are entirely fixed and defined by the micro traits. So far, we saw that one macroscopic state can fit very many microscopic states, while at the same time every microscopic state has only one macroscopic state which fits it. This asymmetry is very important to understand for our purposes, and it effectively expresses the trait of the causal relationship. In the first part of chapter five, we saw that the cause is at least a sufficient even if perhaps not necessary condition for the effect. In our context, this means that a particular microscopic state can be the cause of the appropriate macroscopic state, but the reverse is not possible. This relationship is described in image 2 here:

Image 2: Relationship between the macroscopic state (represented by an uppercase letter) and its realizations, which are microscopic states (represented by lowercase letters). All microscopic states have precisely the same macro traits.

Ramifications for the Neural-Mental Axis The same is true when it comes to body-soul relationships. A test conducted by Prinz, Bucher, and Marder,176 involved the isolation of the entire brain system of a spiny lobster, which is a lobster with a very simple nervous system. Every neuron and synapse was entered into their model, and based on this they built a simulation which included all the possible states of the neurons and the varying degrees of relational intensity between them. They found that the number of microscopic states of the lobster’s brain — the various configurations of the neural network — is about twenty million, as opposed to only about 200,000 macroscopic options of movement. This means that this lobster only needed one or two percent out of the number of possible configurations, meaning that every macro state fits about a hundred micro states. Every one of the hundred micro states is expressed in precisely the same manner at the macroscopic level (i.e., the lobster carried out precisely the same motoric action).177 We can assume that the situation is much the same with us human beings. Thus, the relationship we pointed to in the context of statistical mechanics between the macro and the micro exists also in the brain, with the macro being the mental and the micro being the neural. The conclusion is that if we have complete knowledge of macroscopic (mental) state A, we still do not possess complete knowledge regarding the microscopic (neural) state of the brain. What we know is some traits of distribution, shared by all the appropriate microscopic states, meaning those which realize it (the microscopic states which engender the same macroscopic state). But which of them is the microscopic state which is presently operating in the brain microscopically (the state of every neuron)? We have no way of knowing this based on the macroscopic information. This means that there is a certain degree of freedom between the physical and the mental. Some wanted to derive from the freedom at the micro level that there is a possibility of freedom in a physicalist framework, involving freedom at the macro level. Is this really possible in an emergent picture? If we agree that physical systems, in this case the neural micro level (our neural network in the brain), do not allow freedom, then the microscopic state at any moment unequivocally determines the next microscopic state. This process is controlled by the laws of physics, and we already showed that physics contains no gaps. As we saw now, every microscopic state unequivocally determines the macroscopic traits, while the reverse is not the case. The dynamic dictated by the emergent approach is one controlled by neural micro states, and thus by the laws of physics and chemistry, and this in turn dictates the dynamic of the mental, as the mental is a passive appendage of the neural. But the conclusion is that there does not seem to be the possibility of influence in the opposite direction: That the macroscopic mental level creates changes at the neural microscopic one.178 Like the gas, we cannot think of a situation where the general pressure or the overall energy of the gas will change anything in the traits of the particles which compose it. The general trait is but the sum of the traits of the particles, and a change in the traits of the particles is the result of the laws of physics which act on the particles themselves. Such a change can be expressed in terms of the macro traits, but there is no possibility of influence from the macro to the micro.

The Possibility of the Mental Influencing the Neural in the Emergent Picture So far, we saw that the lack of one-valued mapping between the macro and micro works in only one direction such that every micro state has only one macro state which fits it, but every given macro state

has a huge number of micro states which would fit it (see image 2 above). The attempt to introduce the mental into the physicalist worldview therefore fails, since it requires a non-one-valued-mapping in the opposite direction: Not multiple realizability from the microscopic to the macroscopic, but multiple realizability from the (mental) macroscopic to the (neural) microscopic. If such freedom existed in the system, then one physical state would allow a number of different mental states, and there would be the possibility of arguing that the choice between them is subject to the free will of the individual. But we find no such freedom here, and we therefore conclude that the given physical state entirely determines the mental state. To understand this better, we will try and follow this process in greater detail. Let’s assume that I am presently in a state of anger, a macroscopic state. The question is whether I have a choice between the possibility of striking the person who angered me and the possibility of restraint and avoiding doing so. Such freedom ostensibly exists, as the same macro mental state of anger can fit the two different micro states. Then the choice of the individual is which of the neural realizations which fit the state of anger he prefers to implement. If so, a person in a given mental state can choose between a number of different options in reaction, and we have thus introduced freedom of will and choice into physicalism. But this is a mistake. To understand this, let us assume that we are in the initial macro state of anger. What is the micro state of our neural network? There are quite a few micro states which fit the state of anger, but I at any given moment can of course only be in one of them. The brain as a physical object is not in two different states of matter at the same time;179 it is clear that at this particular moment, every neuron in my brain has a sole, specific, and defined state. Thus, at this initial moment, we can describe my state at both levels in the following manner: Mentally, I am in state A, which neurally fits an infinite number of states which we will mark as follows: a1, a2, a3, a4 … and so on. But as we saw, in practice, my brain is only in one micro state, say a5. We will mark the state of the person with a pair of signs, with the left representing the mental (macro) state and the right the neural (micro) state: In our example, in the first moment I am in state (A, a5). What happens in the next moment? In the physicalist picture, the dynamic of the brain is controlled by the laws of physics and chemistry, meaning by the micro states; we will therefore assume that the laws of physics move the brain from micro state a5 to micro state b8, which is one of the states which fits macroscopic mental state B. Thus, the physical dynamic of the brain leads it from state A to mental state B. Now it is in the state (B,b8). Let us now assume that state B is hitting the guy I’m angry at, meaning the response of that person to the anger which he caused is hitting him. If the initial state was (A,a9), whose macroscopic expression is also anger (A), the laws of physics would take us to another micro state, say c2, whose macroscopic expression is C (restraint), meaning the state (C,c2). These are two descriptions of two dynamics which take us from a state of anger to two different responses. But the decision between them is not actually our choice; it is instead the result of the laws of physics and chemistry, in accordance with the brain (micro) state which existed by us in our moment of anger. If the micro state was a5, the laws of physics would force us to move over to micro state b8, and not any other state. They therefore force us to respond in manner B, and not in any other way. If so, that person has no more than one possible reaction in any given state, and the freedom that exists at the micro level does not therefore allow freedom of will. Hence emergence, if it was the correct description of the relationship between the micro and the macro, certainly does not get the job done — it does not allow for introducing freedom at the macro level into the physicalist picture. This result is derived from two assumptions: (I) In emergence, the micro determines the macro, and not vice versa; (II) Physicalism dictates that the micro is controlled by the laws of physics, which are determinist (i.e., deterministically set the next micro state based on the present micro state). The conclusion is that it necessarily arises from the materialist-physicalist picture that the reaction of that person to the initial state of anger is not a result of his choice, or of his free will, but a necessary result of the laws of physics. The determinist therefore argues that if this process is accompanied by a feeling of free choice — say, the person felt that he freely decided to react with violence, for instance — this is but an illusion. We once again see that physicalism dictates determinism, and emergence cannot create the desired freedom. Just like quantum theory and chaos theory, emergence does not allow the introduction of free choice into the physicalist picture. Physics remains without gaps. But how does a libertarian see this? He shouldn’t deny the findings of the neurosciences. He also believes that there is an emergent relation between the neural micro and the neural macro which stands at the basis of the mental macro. He also understands that the initial state is (A,a5), but what is the dynamic from here on out? Here we are no longer dealing with empirical findings but assumptions, The deterministphysicalist assumes that the mental does not influence the neural, and the dynamic of the neural is the

result of the laws of physics, which take him from the initial neural state and transfer it to second neural state (say, b8, which fits macroscopic state B), as I described above. By contrast, the libertarian believes that there is sometimes a possibility where the mental intervenes with the neural — not the mental as an emergent phenomenon which cannot affect it itself, but the mental as an expression of the soul which is not a physical entity and which influences the material. For the libertarian, will can move electrons. This means that in his normal conduct the human being does indeed react in an entirely mechanical way, and then the physicalist description is correct — the neural micro state defines the person’s reaction. But in the cases in which the person chooses to act otherwise, there it may be that the next micro state will not be the one dictated by the laws of physics, meaning (B,b8), but another state which the mental will decide, say (C,c2), with C being restraint (not hitting the angering person). According to him, the move from micro state a5 to state c2 is because of the intervention of our free will, even though it should move to b8 according to the laws of physics. Therefore, even though the laws of physics would lead the person to react by striking the other person, he may choose to restrain himself instead.180 To do so, he will create a force field which will move electrons into one of the microscopic states which fit C, and in our case c2. This is another look at the mechanism of the influence of will on physics.181 Once again we see that to believe in freedom of will, we must abandon physicalism and allow the mental to start a new physical chain. This means accepting the assumption that will can move electrons even without a physical cause (i.e., against the laws of physics). Only if the spiritual can affect the material (psychophysical influence) is freedom of choice an option. This is the interactionist dualism I speak of, which as we saw is the only alternative to determinist physicalism. But if we as libertarians rejected the latter, then we are necessarily left with the former.

Summary In this chapter, we examined emergence in two contexts: In reference to the question of materialism and regarding the question of determinism. We arrived at two conclusions: (I) The materialist assumption that the relation between the neural and the mental is via a mechanism of emergence is unreasonable, scientifically and conceptually. As Yehiel Mohar put in the quote at the beginning of the chapter: There is no silence within the storms, and no mental qualities hiding within the physical-neural. (II) Even if this assumption was reasonable and if we did adopt the emergent mechanisms and use them to explain mental events (which apparently cannot be done), that would still not be enough to rescue the materialist from the determinist conclusion that his position requires. Choice, in any event, cannot be introduced into the physicalist picture via emergence; and so the materialist is forced to be a determinist, and emergence does not change this. At this point in the discussion, materialism does not appear to be a reasonable option. There is no basis for the thesis that mental processes can be based on neural descriptions. Regarding determinism — we saw that if our conclusion is dualist, the question remains open. We can speak of determinist dualism, really epiphenomenalism, which accepts the separate existence of the mental but sees it as an epiphenomenon of the neural. Even in dualism, the relation between the material (neural) and the spiritual (mental) can be seen as a one-directional causal relationship, with the neural engendering the mental but with the spiritual unable to affect the material. While free will, as we saw, must assume also the influence of the spiritual on the material-neural. But there are two significant points here, which we mustn’t forget: First, if we have arrived at a dualist conclusion, then there is no longer any reason to assume determinism. The attraction of the simplicity in determinism is not on our horizon, anyway. In addition, we need to add the a priori arguments and intuitions we have accumulated thus far, and which also lead to a libertarian conclusion. Second, the primary motivation for determinism is a rejection of the possibility of interaction between the spiritual and the physical, usually called the psychophysical problem. But those who do not accept such interaction should also not accept the opposite direction of the interaction, from the physical to the spiritual. If we are dealing with two substances which cannot interact with one another, then the influence from the physical to the spiritual is also unreasonable. On the other hand, if we do accept the existence of such influence (say, some situations which create states of depression, which in turn create various states of matter), and in addition we are already holding a dualist position, the natural conclusion is that between these two levels can occur two-way interaction. Interactionist dualism is therefore certainly the more reasonable conclusion from the picture described thus far. Obviously if spirit or mental processes can move electrons, then we naturally arrive at libertarianism.

Still, it’s important to understand that this conclusion is not necessary. We are speaking of relatively high likelihood and not complete certainty, and certainly not solid scientific certainty. Therefore, if there is clear empirical evidence in favor of determinism, or for epiphenomenalism, then we must recall that this position has not yet been rejected and is only less likely at this stage. Now we reach the empirical-scientific question which breaks out in full force: What does the empirical evidences tell us? Is there proof for the determination of the mental by the physical or not? Does neuroscience provide us with concrete and direct empirical information relevant to this question? The next two chapters will be devoted to these issues. 167 John Searle, Mind, Brains, and Science, Harvard University Press, 1984. 168 This is very similar to the Mary’s Room thought experiment presented in chapter eleven. 169 Alan Turing, the British mathematician who founded computer science, defined a test named after him which would determine if a computer can be considered conscious. It is similar to Searle’s Chinese Room, and will be discussed in chapter sixteen. Searle’s example is of course influenced by the Turing test but appears aimed at disputing it. 170 Those who answer in the affirmative would probably not pass the Turing test itself… 171 In this sense, we can say that the neurosciences refute the Turing test. They provide a criterion for distinguishing between two states which Turing would not be able to. As far as Turing is concerned, the man in the Chinese room and the Chinese speaker are acting the same way, and they are therefore identical. But a brain scan of both would immediately show that these are two entirely different phenomena. It’s interesting to compare this to another finding of neuroscience: When a person imagines Grandma, he activates the same regions in his neural network which operate when she is physically in front of him and he identifies her in the real world. These are two states which even neuroscience has difficulty distinguishing between, although there is certainly a difference in terms of the flow of information from the sensory receptors to the brain’s processing centers. However, the Turing test is not like that: There it is clear that neuroscientific tools would be able to easily tell the difference between scenarios. 172 R and n are constants whose value is not important for our purposes. 173 Biologists almost unanimously agree today that vitalism is neither needed nor beneficial at the scientific level. Here I will only note that to the best of my judgment, this conclusion is too hasty in light of the present state of knowledge on the relationship between physics and chemistry on the one hand and biology on the other. 174 Ibid., p. 137. 175 Above, I distinguished between something which has no chance and something lacking a foundation or basis. It’s hard to speak of the odds that this argument is correct, as it can’t even be tested. I therefore used the expression “lacking foundation.” 176 See their article: A. A. Prinz, D. Bucher, & E. Marder, “Similar network activity from disparate circuit parameters,” Nature Neuroscience, 7 (12), 2004, pp. 1345-1352. 177 Although it isn’t clear to me how one can reject the argument that the other configurations are responsible for mental events and not motoric ones. 178 Gazzaniga raises this possibility in his book, but they do not sit with the laws of physics and statistical mechanics as we presently understand them. 179 As we said in chapter ten, quantum theory allows for superposition states, meaning the presence of a number of states in a combined state, but as we saw this does not happen at the scale of the brain. Beyond this, we saw that deciding between the different states is subject to a distribution which is dictated by the wave function, not human choice. 180 What we are describing here is the process of the veto. In chapter two, we saw that the circumstances dictate the next situation of the person (he will end up in a lower location in the topographical layout), unless he chooses to veto them and choose otherwise (ascend the mountain). This is precisely the description we saw here, as well. In the next chapter, we will see this from a different vantage point when discussing Libet’s experiment. 181 An interesting question is which of the ci micro states would be chosen in such a process. The mental choice is not to react, meaning a choice of macro state C. But this state has very many micro states which fit it, and the brain is supposed to choose one of them to realize macro state C. It may be that among all the ci micro states, the closest micro state in terms of the laws of physics to the initial state will be chosen, meaning the one which requires the least energy to move to. But this is of course speculation and there can be other mechanisms underlying this process.

Chapter Fourteen THE LIBET EXPERIMENTS AND MORAL DECISIONS Rabbi Hanina said: All is in the hands of Heaven, except for the fear of Heaven, as is says: ‘And now Israel What does the Lord your God ask of you but to fear.’ – Babylonian Talmud, Megilah 25a We don’t do what we want but want what we do. – Common saying in neuroscience This chapter, as well as the next, deals with the consequences of neuroscientific research for the discussion of determinism (and materialism). The first part of the chapter will deal with the important experiments of Libet, which tried to empirically test this question, as well as the follow-up experiments which continue to this day. In the second part we will deal with the manner of moral decision-making: Emotion vs intellect, and brain vs qualia.

A. The Libet Experiments In this part of the chapter, we will examine a fascinating series of experiments which ostensibly deals directly with the question of determinism. This is a clear and almost unique case of an attempt to take the question of determinism from the philosophical level to the scientific one. In the cases mentioned so far, we dealt at most with the assumptions of researchers, and effectively with nothing more than declarations. We have not yet presented any truly empirical findings which support the materialist-determinist thesis. The series of experiments started by Libet in a fairly early period of neuroscience (the beginning of the 1980s) is irregular, as these experiments directly test the question of determinism with scientific tools.

The Libel Experiments: The Rationale At the beginning of the 1980s, the results of experiments conducted by neurologist Benjamin Libet regarding voluntary actions were published.182 To understand the experiment we need to talk a bit about the essence of voluntary actions. Already in 1964, Hans Helmut Kornhuber and his student discovered that every motoric action we do (those which are not reflexes) is accompanied by a neural-electrical activity within the motor cortex. A form of electric potential is formed there which is called readiness potential (below: RP), which accompanies the decision to carry out the action. This sort of potential does not immediately appear but rather gradually awakens until reaching its peak after about half a second. Its general average form, as measured by an EEG machine, appears in the following graph:

Image 1: Readiness Potential as a Function of Time. Dotted line represents the moment the action is taken in practice — it occurs about a quarter-second after the RP reaches its peak.

This electric potential can be measured with an EEG attached to the skull of the test subject, registering the brain activity or waves. EEG is an old and fairly primitive machine in terms of today’s neuroscience, having been invented already in 1885, and serving neuroscientists and doctors to this day, alongside many other measuring and imaging tools, but when we are only interested in the general question of whether there is electrical activity in the brain and the region where it is occurring, the EEG is definitely sufficient. Libet’s idea was surprisingly simple. He understood that the RP may help us empirically examine the question of determinism. In chapter five, we saw that one of the characteristics of the causal relationship is the cause temporally preceding the effect. Meaning that, in principle, if the RP appears even before the person made the decision, then it turns out that his action was not done due to his decision but due to the influence of spontaneous neural activity in the brain, which determined that the action would be carried out. By contrast, if the libertarian is right, meaning that if the person’s decision created the action then it is clear that the decision should be adopted before the readiness potential appears. In essence, the question was whether the mental decision determines the neural state which leads to the action, or whether the neural state leads to the “decision” (which is just an epiphenomenal feeling of a decision). Do we do what we want or do we want what we do? The graph in image 1 represents the readiness potential as opposed to the effective time of carrying out the action (the dotted line). Libet understood that we need to add another line on the time line, a line which would note the moment in which the person’s conscious decision was made to carry out the action. If this point is to the left of the entire graph, then the libertarian is correct. But if this point appears after the

appearance of a significant part of the RP (relatively close to the dotted line), we will see this as clear evidence for determinism. But how will we measure the moment of the conscious decision-making? We already saw that mental processes are qualia, and that unlike physical phenomena they are not accessible to measuring tools. Libet proposed a solution which may sound very primitive, but which ultimately proved to be incredibly effective: We will simply ask the test subject. The Libet experiments were thus conducted in the following manner: The test subject sat in front of a clock, with his head connected to an EEG. The researcher instructed the subject to do some action in his own free time, meaning whenever the subject himself decided to do so. The action in question was something like lifting a finger or pressing a button. The subject looked at all times at the hand of the clock which ran in front of him, and was supposed to report to the investigator at the precise moment when he made the decision to press (a time obviously preceding the actual action). The researcher would register this time, and compare it to the results regarding the RP from the EEG, which measured the time when the RP appeared. The human reporting of the time of decision is the Achilles Heel of such an experiment, as it is a very inexact method of measurement. To improve precision, the clock the subject sat in front of ran quickly, as the faster the clock, the more accurate his report (the resolution went up, as every change in the clock represented a period of time much faster than a second). Still, such human reporting is obviously a very primitive mechanism.

The Libet Experiments: The Results To the researcher’s surprise, it turned out that a significant and clear readiness potential appeared already a third of a second before the time the subject made the decision.183 Image 2 shows the results of the experiment. This is a precise copy of image 1, only now we have placed the beginning of the graph at the point in which the decision was made:

Image 2: Libet Experiment Results. The origin of the axes is located at the point in time when the decision was made (according to the subject’s report), and the dotted line represents the time the action was taken (which can be seen with our eyes). The graph represents the results of the EEG measurement.

Here we can see again that the RP reached its peak almost a quarter second before the action was carried out, but it began to awaken more than half a second before the decision was made. Thus, the decisions we make begin in a neural process which precedes our conscious decision by a significant margin of time. This means that decisions are not actually what produce the actions which follow them. These actions are already set in advance by the neural activity, regardless of our decision-making process. To the contrary, the “decision” is nothing more than an epiphenomenon which accompanies the neural process. We therefore ostensibly conclude that the neural clearly determines the mental, and not vice versa. Subsequent experiments revealed even more amazing findings. In 2008, Soon and his colleagues checked a decision to lift the right or left hand at a given moment. It turned out that the decision to lift one of the hands was made a second before the action, per the subject’s report. This decision was adopted in two different regions of the brain: One responsible for setting the time of the action, and the second responsible for the decision of which hand to lift, the right or the left. It turned out that the RP in the region which determines when the action will happen appeared six seconds before the action itself, or five seconds before the decision was made. Meanwhile, the RP in the region which determines which hand would be raised started eight seconds (!!!) before the actual action. These time scales are already far more convincing; they can even be measured on an ordinary watch

measured in seconds. These results ostensibly show us that free will is nothing more than an illusion. Our actions are predetermined by the neural network in our brains, and conscious decision-making plays no part. The experience of the conscious choice which the subject reports seems more like the result of the neural activity than the factor which awakened the neural activity. The natural conclusion is that the micro determines the macro, and the feeling of freedom in our consciousness is nothing but an illusion. To paraphrase, we can say that the Libet experiments taught us that we do not do what we want, but want what we do. This amazing result sent the world of neuroscience (then still in its infancy), and even more so philosophers dealing with the question of free will, and especially the libertarians among then, into a tailspin. This ancient and undecided philosophical question was now ostensibly decided in an unequivocal, empirical manner. Determinism won in a knock out. It’s interesting to note that Libet himself, who espouses a libertarian approach, was no less surprised and confused, and as we will see he also tried to find a way out of the trap he had himself created, or really discovered. It is no wonder, then, that many philosophers and neuroscientists have offered various critiques of the surprising results of this experiment. Later and more complex versions of this experiment, which are conducted to this day, refuted some of these critiques.184 Gazzaniga185 and many others therefore see the results of these experiments as clear proof that our brain makes our decisions, and not the conscious level.

The Libet Experiments: Reactions and Critiques One can get an impression from the problematic nature of the Libet experiments, and no less from the hysterical reactions which its results aroused, by looking at the enormous number and variety of critical responses which these experiments received, as well as the number and variety of follow-up experiments which continue to be conducted to this day in an ever more sophisticated manner. My goal in this section is merely to show some of this enormous corpus of criticism, if only to reflect the importance which many scientists and philosophers see in these experiments for the determinism debate, importance which I will entirely reject later on. Libet himself, who as we said saw himself as a libertarian, addressed the large majority of the criticisms and proposed a broad and systematic picture of libertarianism. To the best of my knowledge it is very similar to what I presented here, and I will try and show this later on. In the second chapter of Ariel Ronen’s Master’s thesis,186 which is entirely devoted to surveying the Libet critiques, he divides the criticisms into four kinds: Technical-methodological, critiques of the modalization of cognitive psychology, conceptual critiques, and broad philosophical critiques. Every one of these categories contains dozens of different critiques of various kinds. Some have questioned the statistical reliability of the findings, arguing that the average times are not a sufficiently successful and precise metric. Libet rejected this and showed that his findings are of high statistical credibility (Ronen, p. 28). Another criticism had to do with the reliability of the subject. As we said, in these experiments, the time when the decision was made — a subjective mental event or qualia — was determined by the report of the subject, and not any objective measurement. It may be that the subject was unable to divide his attention between observing the clock, examining his consciousness, and the motoric action he was supposed to do. Some argued that it took the subject time until he saw, or absorbed, the clock. Later experiments neutralized the clock problem, showing that the effect still exists, and sometimes even strengthens.187 Others argued against him that he did not distinguish between various preceding stages, all of which appeared before the action: Will, planning, and decision. It may be that will came first, and only the planning or the decision appeared after the RP. We should note that even if the RP only expresses will or planning, and not a decision, in the end the fact is that it allows us to know what will happen even before a decision is made, and that is what is important in the discussion on determinism. Other arguments made the point that the decision of the subject forms gradually, and it is therefore unclear when the subject saw the situation as decided. It may be that the decision was effectively made before, and the subject’s report is at a later stage, when his decision was already complete. By the way, we can already see this as a phenomenon parallel to the gradual formation of the RP itself (as we see in the above graph, reaching its peak after the moment the decision is made). This argument is hard to accept regarding Soon’s experiment, which speaks of a gap of eight seconds. Others made an even more far-reaching claim. They argued that the subject’s decision forms in an unconscious manner. He report the time of the decision only after he is already aware that he decided, and this indeed only happens after the appearance of the RP. But it is not true to say that the RP creates the decision, but it may be that the (unconscious) decision is what engenders the RP, and that consciousness

only comes later. As Libet himself noted, this is a very problematic argument at the philosophical and ethical levels (Ronen, p. 42 passim; he deals there with the much broader question of whether one can even speak of unconscious psychological processes, or whether anything which is unconscious Is nothing more than mere neurophysiology). The accepted approach of libertarianism espouses conscious decision. If we are also willing to accept unconscious decisions as libertarianism, then even if there is a free, non-deterministic mechanism involved, such libertarianism has no philosophical or ethical significance, as Libet himself noted (Ibid., p. 47 passim). Judgment by definition is a conscious mental event, and without consciousness one cannot impose moral responsibility on an individual. This is imposed only on those deficient moral judgments and not unconscious neural processes, regardless of the question of whether they are free (nondeterministic) or not. This argument is thus problematic at two levels: (1) There is no reason to assume that there is a mechanism of free but unconscious decision-making. Libertarian intuition is based on our awareness of freedom of will and not something unconscious. (2) Even if there is a possibility to speak in principle of a free, unconscious decision, this is not libertarianism. Libertarianism is supposed to offer a basis for moral responsibility and our libertarian intuitions, and the picture being proposed here does not save it from the consequences of the Libet experiments. Put differently: With this sort of libertarianism, who needs determinism?!? We can of course speak of different levels of consciousness (which are along some sort of spectrum, which might be continuous) instead of a binary picture where there either is consciousness or there isn’t. We can also raise the possibility that there was consciousness during the decision, but that we forgot it. We cannot retrieve the information on this consciousness of ours from the past. Others have raised the possibility that our consciousness of will does not appear at the same time as the conscious will itself (Ibid; p. 49 passim; see also p. 52 passim regarding the objective measurement of subjective processes such as consciousness in general). Some questioned the very difference between conscious and unconscious (Ibid., p. 54 passim). Other spoke of various delaying methods which lead to the feeling of a decision being reported later than it was really adopted. Libet rejected these arguments as well (Ibid., p. 31). Some argued that one can affect the time of the appearance of will in various ways.188 Some hypothesized that events after the pressing of the button affect the time of the appearance of will.189 There are additional critiques of Libet’s arguments regarding the question of precisely when and what needs to be free for it to be considered an act of free will (Ronen, p. 75). Other critiques deal with the concept of time assumed by Libet (Ibid., p. 77 passim). Libet almost doesn’t address these critiques, but they all assume, as Libet himself insists, that his findings do indeed refer to an action of free choice. We will end this critique survey by mentioning a collection of conceptual critiques, most if not all referring to Libet assuming a picture of interactionist dualism (Ronen, p. 57 passim).190 As we pointed out in previous chapters, this is indeed the only possible foundation for libertarianism. The proposals for a libertarianism which is physicalist-materialist do not hold water. I will therefore not delve further into these critiques. In any event, we must not forget that the determinist option is always there. Bottom line, we can get the impression that if the libertarians have need of such wild speculations, then the question of free will seems a little beyond hopeless. I should note that Ronen himself (in the third and fourth chapter of his thesis) proposes his own libertarian solution (which I don’t agree with). But even if all the arguments surveyed here were summarily rejected, we could still show that the Libet experiments are not at all relevant to the determinism debate, as they do not measure an action of free will. As we will see immediately, there are three arguments which ultimately almost entirely neutralize the sting of the results of Libet’s experiments.

First Argument: Libet’s Veto Ironically enough, Libet himself is a libertarian, and he was therefore the first to be surprised by the results he got. It is therefore no surprise that he proposed the first explanation which left libertarianism in the picture. Libet argued that it is true that the RP precedes the decision, but the decision is not always carried out. The subjects reported states where they vetoed the RP. This means that whenever the action was carried out there was an RP about 0.35 seconds before the decision, but there were states when there was an RP and no subsequent action — the subject did not lift his finger or press the button. This means, according to Libet himself, that the subject had the ability to veto the preceding neural order.191 Libet ties the imposition of the veto to a part of the brain called the Caudate Nucleus. He also argues that people who suffered from pre-fontal brain damage, and who cannot delay unwanted conduct (people with Tourette’s Syndrome or OCD), could not in some cases impose such a veto (Ronen, pp. 21-22).

This of course leads to a serious question: Is the veto itself not driven by another RP? Perhaps it is also the consequence of a neural state and not our free decision (Ronen, p. 36, passim, section 2.2.3). Libet himself noted this, and said that there is obviously a neural state which is parallel to the imposition of the veto (nothing in human activity happens without a neural event). But in the context of the veto, the neural state does not precede the decision (there is no preceding RP). According to him, the reason for this is that the decision to impose the veto contains no value content in and of itself, only negative content. In any event, he adds, we do not indeed have clear empirical proof that the imposition of the veto is not also a result of a prior neural state, and it is clear that his experiments do not prove this. The question of determinism therefore remains open at the scientific level (Cf. Ronen, pp. 25-26, and p. 37). What does this mean? Why does there seem to be such a one-way connection between the RP and the decision? Whence the asymmetry of the connection (i.e., there cannot be an action without an RP, but there can be an RP but no action)? Here we return to the picture of the topographical layout which we presented in detail in chapter two. A person who makes some decision freely is not free of urges and inclinations. He acts within a complex environment, and this affects him in a variety of ways. He is affected by genetics, his environment, his emotions, various conditionings, neural influences of the brain, and so on. We described there how all these create a sort of topographical landscape which operates on the person in various directions (pushing or pulling) towards various modes of action or inaction. There are directions which are very easy to follow (valleys of varying depth), and other directions which challenge him (mountains of varying steepness). But libertarianism believes that, in the end, the person is supposed to and can decide what he does with all these pressures. The buck stops with him, not the topographical landscape. It’s easy to see that this is precisely the significance of the veto. The neural activity pressures the person to push the button. Now he has a valley in the direction of pushing and a mountain in the direction of not pushing, and this is precisely what Libet called the veto. In other words, Libet’s experiment did not measure the person’s decision so much as the topographical landscape within which the person operates. The appearance of the RP reflects the development of a neural urge in the direction of pushing (which is a result of all the environmental influences: education, genetics, environment, &c), and the veto describes the person’s ability to reject this push (meaning to climb the mountain) and not push. I will note here that Libet himself stated that his veto theory very much fits our feelings, as we all sometimes feel that we have a desire to do something and we decide not to in the end. By contrast, a state of action without a preceding RP may not be possible (although this isn’t clear), as the motoric action cannot begin without an order from the brain. This means that for the Libet experiment to be a good indication for determinism, it must succeed in 100% of cases. If there are a few cases where an RP appears but does not dictate the decision (i.e., a veto is imposed), this is sufficient to argue that the subjects sometimes act according to their neural impulses and sometimes not, entirely befitting the libertarian picture. This means that the results of the Libet experiments and their follow-up efforts are almost meaningless. No libertarian claims that there aren’t various urges operating on individuals before they make a decision. The libertarian’s argument is only that the individual can resist those urges, vetoing them, and the determinist will not agree to this. In chapter two, we described this dispute as follows: According to the determinist, this topographical layout is the person himself and it therefore determines what will happen in practice. According to the libertarian, the topographical layout tries to influence the person, but he can decide whether or not to assent to it or not. If we understand the model we presented in chapter two, then it is obvious why the results of the Libet experiments are irrelevant. Only a naïve conception of libertarianism, according to which liberty means freedom and that free will means that the individual acts without urges or inclinations but rather in an entirely free manner (in a vacuum), could lead people to such an incorrect interpretation of the results of the experiment. Libet’s veto is a predictable and almost trivial argument, at least if we are aware of the full significance of libertarianism. The debate with the determinist is only on the question of whether there even will be cases involving vetoes, and Libet’s experiments have little to say on the subject. We can also look at it from a different angle: In chapter two we saw that even the libertarian approach grants statistical significance to the topographical layout. It is clear that in the very same state, a group of people with some tendency will react differently than a group without this tendency. But the argument of the libertarian is that this is only an average at the statistical level. Every individual is free to react as his free will decides. Therefore, when a particular genetic structure leads statistically to aggression, sensitivity, or cowardice, for instance, none of this amounts to evidence for determinism. Statistical phenomena are

also acceptable to the libertarian. If so, in contrast to what is accepted in the scientific method in other contexts, on this issue only 100% success can decide the question of determinism. The conclusion we have arrived at so far is that the results of the Libet experiment are not relevant for our discussion, as there can be and are expected to be veto situations. But it still appears that such an experiment can in and of itself be relevant for the determinism debate. To understand this, let us think of a hypothetical situation in which results emerged in the experiment where there was no veto in all states (a 100% success). These results would ostensibly prove determinism, although even then there would be room to say that all the subjects chose to acquiesce to the RP by coincidence, though this seems a fairly far-fetched excuse. Thus, it seems that the experiment could be relevant, but all we can say for now is that the specific results of the experiment already conducted have left the question of determinism just as open as before.192

Second Argument: Picking VS Choosing Let’s now go one step forward and see that even complete success, meaning no cases of vetos at all, would not be a clear indication of determinism, and that the experiment itself is therefore fundamentally irrelevant for the philosophical discussion, not just any results therefrom. Let’s think what would happen if the experiment would produce no cases of vetoes, meaning that whenever an RP appeared, the subject would decide to press the button, and then the action of pressing would occur. This is one hundred percent success, the dream of every determinist. But in light of what I explained above, we can understand that even in this situation the results would have no significance; it would at most mean that the person decided in all cases not to apply the veto, but this does not mean that he does not have the option (or the ability) of imposing one. We saw a similar distinction in chapter seven, where I argued that the fact that we can predict the continuation of the numerical series does not mean that it is deterministic (calculation-wise). It is certainly possible that we are dealing with the result of the individual’s free choice, except that he chose at all times to act according to these fixed principles, or according to his topographical layout or urges. However, I already noted there that such an explanation seems a little far-fetched. Science is always built on generalizations based on our observations. It is more reasonable to accept the argument that a hundred percent success is needed, but the argument that even such a degree of success is insufficient here seems a little problematic, even though not entirely absurd. To understand why even that is not sufficient here, we will delve deeper into the Libet experiments. The Libet experiments speak of a situation where the subject must decide on a banal action like pressing a button or lifting his hand (and not murder, or rebellion against a tyrant, or another decision with such major consequences). In such situations it’s pretty clear that he will not veto the RP. Why would he? Does he have any important consideration which would lead him to go against his urges? In such a situation lacking any value content, it is very reasonable that the person will act according to his urges and pressures and not apply his autonomy. When the urge to press the button awakens is when he will actually do so. A man may fight his tendencies where there is a moral dilemma or complex issues of judgment involving two sides, but there is no such logic in banal cases such as these. We can put it this way: Why did the subject choose to press at that particular moment and not another? There’s not good reason to prefer one moment to another. So what led to the decision? Obviously because it is precisely at that moment that his RP awakened, and he therefore pressed at that moment. Even the libertarian will accept this. Here we should repeat the libertarian picture described in chapter two, reminding the reader of the three exceptions in the libertarian picture, which were mentioned at the end of the chapter (in the “Mechanical Actions” section, as well as chapter six). We saw a case there where a person is driven by an unconquerable urge, or the urge itself (without a willful decision) which determines the action. In such a case it is clear that the RP will always lead to the action and there will be no cases of vetoes. The second situation I presented there is of an action done for no reason, because the person does so without thinking and on instinct. The third case involves an action which has no reason, meaning an action which does not involve value considerations, and which therefore requires no judgment by definition. The decision to press the button in the Libet experiments is precisely this third case. Such actions are carried out due to a neural RP without any conscious decision. This is precisely what Libet found, and this “discovery” now seems fairly trivial. The libertarian would also agree that such actions are done without choice or judgment. We can now see the significance of libertarianism as described in chapter two. Anyone who understands the picture described there should understand that Libet’s experiments are not relevant for the debate, regardless of their results. This is one of many demonstrations of the importance of a discussion of the a priori principles and conceptions before we conduct experiments and analyze results. As some of his critics pointed out in terms of terminology, he looked into situations of “picking” and not

“choosing.”193 What they mean is that the situation Libet tested does not really involve judgment or dilemmas, and therefore we at most have a decision but certainly not a choice. In a case without judgment, we should not expect a veto; it turns out that what will determine the timing of the button pressing in such situations is a random urge which awakens at some point at the neural level (the person’s topographical layout). Libet himself tried to defend the relevance of his experiment on the grounds that what exists in a basic action must also take place in a more complex one (Ronen, pp. 35-36), but this is but speculation without much basis, and certainly not a scientific fact.

The Sophisticated Trolley Experiment As we said, this second argument leads us to the conclusion that even if there were no vetoes at all, there would be no real proof here of determinism, as the experiment examined fairly trivial situations. Even a hundred percent success would not help. We will now go one step forward and see if it is nevertheless possible to conduct a variation of the Libet experiments which would be relevant for the question of determinism. In light of what we’ve seen so far, it is clear that an experiment like that of Libet’s can deal with the question of determinism only if we conduct it in relation to situations which contain real dilemmas, such as the trolley scenario.194 In such situations, value judgments exist in both directions, and a person should be in a real dilemma. If we can predict — always and without exceptions, with a hundred percent success rate — based on the pre-decision RP whether the decision will be to divert the trolley or push the man off the bridge, this would already seem to cast a heavy shadow over libertarianism. In chapter twelve we saw that two kinds of experiments were made with the trolley scenario. The first was a statistical test of how people reacted to the two situations. This is an experiment unrelated to neuroscience, but psychology — the manner in which people react to different situations.195 The second type is a neuroscientific experiment, meaning a repeat of the trolley experiments with an EEG, to see which regions of the subject’s brain are activated when making these decisions. To conduct a trolley experiment which will be relevant for our discussion, we have to hone the second type of experiment: In addition to the EEG, we will have to place a clock before the subjects, so that they can report the time when they made the decision. If the time appears clearly after the appropriate RP for this sort of decision, then this will be a very clear step in the deterministic direction. If we accepted that we can in all situations predict the conduct of the subject even before he makes his decision — meaning, if we found this kind of distribution of results in which all those who decided to divert the trolley had an appropriate RP, and those who decided not to had a different neural signal appropriate to that decision,196 and all this before the conscious decision — this would ostensibly be a significant and clear indication in favor of determinism. But even in such a case there would be room for arguing that he does in principle have the ability to impose a veto and simply doesn’t use it, as we noted above. If so, even in such a situation we are still not dealing with an experiment which would serve as unequivocal proof of determinism. As we saw at the end of chapter two, people sometimes make decisions instinctively, without activating their higher thinking.197 Therefore, even if we find that all the people tested in these trolley experiments always follow the RP which preceded their decision, this may point to them not using their cognitive controls and acting only based on neural pressures; they always had the possibility of imposing a veto but simply didn’t use it. In chapter two, we dealt with this problem. We saw there two kinds of human actions done without judgment. The first kind is of a person who decides to push a button. No considerations are involved here at all, and this is therefore an action which has no fundamental connection to judgment. There is another kind, actions which do require judgment but where people nevertheless decide to do them based on instinct (such as the example we brought comparing people to bees, which showed that people like bees prefer the option which is always successful, even if the profit margin is lower). I explained there that the difference between people and bees according to the libertarian picture is that people can be taught to act rationally, meaning they can resist their instinct and use their judgment. The original Libet experiment involved the first kind of example, involving an action which by definition does not require judgment. The trolley experiment proposed here is equivalent to the second type, an action requiring judgment, but it is clear that people do not always use it. The way to handle this, as Daniel Kahneman stresses in almost all of his works, is to teach them not to surrender to instinct and use judgment. If so, we must train the subjects in the trolley experiment not to decide instinctively and to use moral and logical judgment, and only then conduct the experiment which we described above. After we taught the subjects to consciously reconsider the question, and not act automatically based on the neural urge, we can once again examine the temporal relationship between the RP and the decision-making in the trolley experiment. This is a more sophisticated trolley experiment, and if even now we always get an RP

prior to the decision, then we will ostensibly finally perhaps have a real indication in favor of determinism. But we still haven’t made it to the Promised Land. Now it may be that the training of the subjects simply changes their neural system — the structure of the network, as we saw in chapter twelve — and that they are once again making decisions in an entirely mechanical manner. If our training changed the network structure, then the RP will appear in a different manner or in another region of the brain. Instead of the decision being made in the region responsible for emotion or moral intuition, for instance, the region responsible for making complex decisions will be activated instead. But the RP in the new region may still appear before the conscious choice, which once again returns us to a deterministic state. This means that even in this situation, we will not have clear proof of determinism — people might still have the option of imposing a veto, but they simply do not use this ability. To further improve the experiment, we must train and teach the subjects to make rational decisions and avoid instinctive decisions entirely, and not just in the context of experiments. If we can train them to overcome their instincts and operate in a conscious, controlled fashion, then when they encounter a situation or dilemma requiring a decision, they will not act based on their gut. But this decision will not be the result of mere mechanical thinking, as their network was not trained to make decisions in these new situations. If such training succeeds, the libertarian prediction is that the mental will activate the neural, and that the RP is therefore expected to not always appear before the conscious decision but sometimes afterwards, or that the conscious decision will be different than what the RP predicts. Only when such an experiment is conducted will the question of determinism start to approach becoming a scientific one. Still, the truth needs to be said that even if we accept the results that the RP entirely predicts decisions in this perfect experiment, we still cannot confirm determinism for certain. The question will still remain whether we trained the subjects properly and thus whether the actions of the subjects were necessarily instinctive and determined by the RP, or that they had the option to impose a veto but didn’t. Moreover, it would be much harder in such a situation to point to the precise moment in which the decision was made as well as the region in the brain where the RP is supposed to appear, and therefore many of the critiques of the original Libet experiment will become ever more strident in this instance. This convoluted and exhausting description can give us the impression that even in such a complex experiment, the results will apparently not be unequivocal. Even the most sophisticated experiments will require results of a hundred percent success without any vetoes, and even such results will not amount to an unequivocal indication of determinism. It therefore appears that we are very far away from a test which would be both empirical and provide unequivocal answers for the question of determinism. We certainly cannot say that the scientific findings today point to such unequivocal indications. The declarations of philosophers and neuroscientists in this direction seem to me to be somewhat exaggerated, to say the least.

The Asymmetry Between the RP and the Action We will immediately go another step forward and entirely reject the relevance of these experiments. But before this, we need to add a brief clarification regarding the aforementioned asymmetry. We saw that Libet assumes that there were veto cases in his experiments, meaning situations where an RP appeared yet the person still did not act. The accepted assumption is that the reverse cannot take place, and indeed did not happen there. This means that it’s pretty clear that there cannot be a situation where there is no neural instruction yet the person still decides to press the button. There is thus an asymmetry between the RP and the action of pressing the button. Libertarianism, by contrast, espouses the position that a person’s decision can move electrons, meaning the neural system, and through it the motoric system (meaning, carry out actions). Therefore, even if we succeeded in “saving” the libertarian from Libet’s experiments’ results, we would still have to explain why we have not observed this phenomenon. The explanation for this may very well be similar to what we proposed in the previous section. The action of pressing a button, like any act of “picking,” is not an action requiring judgment. In such a state, will does not make decisions, and it therefore does not move electrons or give orders to the neural system. Here, behavior is determined only by the neural system, and will can at most impose a veto, and as we saw even this is usually unlikely to happen. In any event it is important to stress that even if we conducted the sophisticated trolley experiment described above, the normal libertarian picture would predict that the two possible situations can indeed occur: (1) An RP prior to the decision and action; (2) a decision prior to the RP and which effectively engenders it, and through it also the motoric action. Obviously the third situation, in which there is a decision and an action without any RP (neither before nor after the decision), is not possible at all. The

instruction to the motoric system must emanate from the brain, as even the libertarian agrees that will has no direct connection to the muscles, and that this connection is mediated by the brain. Therefore, even when will freely engenders the action, without any neural influence, will still needs to awaken a neural RP, and this in turn will activate the motoric system. Third Argument: The Timing of Ethical Decisions198 So far, we saw two aspects which neutralize the extent to which the Libet experiments serve as indications of determinism: (A) The argument that veto decisions are possible; (B) The argument that these are situations involving “picking,” which by definition do not require judgment. The third argument which I will represent here refers to the timing in which a person builds his value scale, and this would seem to neutralize the relevance of these experiments almost entirely. When a person faces a value dilemma, he is supposed to weigh the various options and decide how to act. His decisions are based on some value system, which states which value will override which. Sartre, in his book Existentialism is a Humanism,199 describes a case in which a student of his came to him during WWII and asked for his advice. He wanted to go over to Britain to fight the Germans and take revenge for the killing of his brother who died in the war. On the other hand, he lived with his mother, who was depressed at the death of her son and the conduct of her husband, who collaborated with the Germans, and who was also sick and needed treatment and care. The student was debating whether to stay and help his mother or leave in order to join the fight against the Germans. We have here a clash between two values; there is no practical way to fulfill them both. Obviously we don’t have a direct clash — there is no conflict in principle between the desire to care for one’s mother and to take revenge on the Germans. The clash was created by the practical situation, in which working towards one value does not allow working towards another, and vice versa. The values on both sides of this dilemma are the fight against evil on the one hand and care for an old and sick mother on the other. The decision will be made based on the relative status of these two values in the student’s value system, meaning which of the values is higher on his scale. Now we ask the question: Is the building of a set of values being done at precisely the moment of the dilemma? Maybe the student is already equipped with a set of values he already decided on, and now the “dilemma” is nothing more than a calculation of how to implement it in the present situation? If the student’s set of values already places the fight against evil higher than aid to his mother, and this has been built up within him for many years, then his dilemma is only apparent. There is a calculation which could be entirely mechanical, which calculates these two values and the consequences of the situation for these values, after which he arrives at the natural conclusion. This may not be necessary, and the scale may be under construction right now, or there may still be room for a dilemma within the scale. But the very fact that such a situation is possible means that there may not be an act of free choice here, only a calculation based on a choice made long ago. Let us now assume that we are measuring the time of the appearance of the RP in the student’s mind as opposed to the time when he makes his decision, and we discover that the RP everywhere and always precedes his decision, and that we can therefore predict it. This may indeed be a case of “choosing,” but the result still tells us almost nothing about the question of whether the student freely chose his values or whether the neural state determined and dictated his decision. It may well be that this specific decision is the consequence of a previous decision regarding his value scale, which the subject decided on long ago and in a free manner. What is happening now is simply a technical calculation to reach the correct implementation in the specific situation he is in now. It is no wonder, then, that we will get findings that the RP unequivocally determines the decision, and appears before the conscious choice. This does not of course tell us anything about the question of determinism itself. This consideration only neutralizes the possibility of examining the question in an empirical manner. As we saw, we can always say that the choice of basic values and the hierarchy between them was done in the past, and that we are therefore only observing their technical implementation, and this is indeed an entirely deterministic and mechanical calculation. It is no wonder that in such a situation, the RP determines the action to be taken. Even a devout libertarian could accept such a result.

Back to Weakness of the Will: Value-Value Dilemmas and Value-Urge Dilemmas The case described by Sartre mentioned above, deals with ethical decision-making between two options. The decision on the agenda is which of the two is more correct or more moral, and we saw that this depends on the value heirarchy. We can call this a value-value dilemma. But there are other kinds of decisions, which I presented in chapter two as dilemmas of weakness of the will. These are dilemmas

between the proper and right and the convenient and beneficial, which can be called value-urge dilemmas. The decision of Sartre’s student oscillates between two poles of proper, and he is torn as to which of them is more so. Here he builds a value scale or uses an existing value scale. But a decision whether to steal money or earn money, for instance, is not such a decision; here we have the value of not harming the property of others on the one side, and on the other side is self-interest. In chapter two, we presented these dilemmas as a choice whether to choose — the person wonders whether to act as a choosing person and act on his values, or be dragged along by his topographical mental layout expressed by his urges and do what’s easy. In such a dilemma, it is likely that if we ask that person at any other time what is the right thing to do, he would say that it is right not to steal. Yet still, in that moment, he is “tempted” to steal. If this distinction is correct, and there really are situations of weakness of the will, then it is very likely that choice in the form of choosing to choose is done precisely at that moment, not before. The description we gave above is relevant for value-value dilemmas — a situation in which a person builds a value system, and later on encounters a situation where he must decide how to apply it. My argument is that in such situations, it may well be that this person will do a mechanical calculation and not a choice, as he has already made the value choice in the past. But in the case of the thief, he has long since positioned the value of not stealing in his system — otherwise, we would have no dilemma. What has awakened within him is the conflict between this value and the momentary self-interest, and this dilemma therefore arises only now, when the contrary urge emerges. The libertarian is supposed to think that in cases of value-urge dilemmas we are not dealing with a mechanical calculation but a real decision, made at that moment. Thus, there might be an experiment which could also overcome the third difficulty, and this is an even more sophisticated Libet experiment of a fourth type: Repeating the aforementioned sophisticated Libet experiment, but conducting it in the manner of a value-urge dilemma (whether or not to choose) and not a value-value dilemma (which to choose).200 But another look at this proposal shows that even such a dilemma will not provide us with a clear indication. It may be that the ability to overcome the urge — the power of choice, which as we explained determines the degree of success or failure in dilemmas of weakness of the will — was created beforehand. For instance, the totality of experiences and insight which I accumulated that day create strength within me, large or small, to deal with urges. It may well be that this force was built up even before the moment of truth, and it has even perhaps changed within me at different times, sometimes being stronger and sometimes being weaker. But this does not mean that the force is being shaped at the moment of the dilemma itself. But if this is the situation, then such dilemmas are also decided by a kind of calculation based on prior information, and there is no decision at that moment. If so, even experiments testing valueurge dilemmas will not provide us with a decisive indication regarding the question of determinism. Even if all of the above is incorrect, we still need to understand that it’s very hard to conduct such a “perfect” experiment. We must locate the moment when he faces the dilemma201 and the precise moment when the self-interested urge awakens within him on the other hand, and precisely then to conduct the experiment. Second, we must ensure that this really is a value-urge dilemma for the subject and not a value-value one. How can we know if his dilemma on whether or not to steal is not value-based (between the desire to be rich and the desire not to steal), and then it is no wonder that the RP predicts the outcome (as at this moment he is only calculating)? This requires confirmation that one of the following two possibilities is the case: (A) He indeed has a value not to steal and does not have a value not to be rich, only an urge; (B) The value of wealth is lower on his ethical scale relative to the values of morality, and therefore this is a value-value dilemma. How can we find out? It would appear that at least at the scientific level, it’s hard to put faith in the subject’s own self-reporting. This would perhaps be possible in the future if we have a way of locating and diagnosing the region in the brain where his fixed set of values is, and see what his values are there and what is the relationship and hierarchy between them. Beyond all this, all the difficulties which accompanied the Libet experiments and their follow-ups which dealt with relatively simple situations will accompany us all the more so with such a complex experiment — difficulty in determining the moment of decision, vagueness in the conceptual system of the test subject, the gradual formation of the decision and all the other problems we saw above.

Interim Summary From everything we’ve noted in this chapter we see that not only do the findings in the specific experiments of Libet not attest to determinism, both due to various critiques of the experiment’s methodology and the subsequent conclusions, but that this sort of experiment is almost irrelevant for the discussion. We saw this primarily at three levels, one on top of the other:

The prediction by the RP must have a one hundred percent success rate. Otherwise there is a veto, which means there is free will. We need to examine situations involving “choosing,” which contain real dilemmas, because when they are simple matters of “picking,” even the libertarian would agree that the RP determines the moment that the subject presses the button. Finally, we need to refer to value-urge dilemmas — as in value-value dilemmas, it may be and is even likely that the decision was adopted long before the experiment was conducted, and that the experiment involves mere calculation on the part of the subject — but it is in any event clear to us that this is a kind of dilemma by the subject. But as I noted, it’s very difficult to conduct such experiments. Moreover, we saw that even an experiment testing such dilemmas does not provide us with an unequivocal indication of determinism. I should note that I did not include all kinds of secondary reservations in this discussion, such as the need to find out if we are truly dealing with choice and judgment and not an emotional action (which does not involve judgment, even if it should). The conclusion is that even in the future, an enormous scholarly effort will be required to move towards a serious empirical examination of the question of determinism, if it is even possible and indicative. Bottom line, it is not at all clear, even in principle, whether the question of determinism can be unequivocally decided empirically in that kind of experiments at some point in the future. We should remember that at least until now, the Libet experiments are the first and almost only ones to directly test the question of determinism in an empirical manner. This is a very problematic effort, and as we saw above it doesn’t even produce the desired certainty when it comes to determinism. The other arguments are purely philosophical and are not supported by any empirical scientific basis. The conclusion is that contrary to the declarations of many neuroscientists, it appears that the discussion of the question of determinism being conducted today is not fundamentally different than that which took place before the beginning of modern brain science. As of right now, the neurosciences contribute nothing essential to the discussion, aside from greater resolution when it comes to the brain mechanisms. The question of the relationship between these and our choices remains a philosophical question.

A Note on the Illusion of Choice Despite all these arguments and reservations, there is one aspect of the Libet experiments which is nevertheless relevant to our discussion. The subjects in these experiments felt that they are freely choosing the timing of pushing the button. They even reported that moment to the tester, meaning it was clear to them that they decided. But if their decision was indeed determined by the neural RP — and we saw that when it comes to picking, even the libertarian would agree with this argument — they still felt an experience of free choice, the libertarian would seem forced to agree that this is an illusion. After all, he also agrees that in such situations they did not actually choose, yet they still at that moment had a clear feeling of making a free choice. This ostensibly raises doubts regarding the entire libertarian intuition. Perhaps the experience of freedom of choice which libertarianism is based on is indeed no more than an illusion? If even a person who doesn’t decide experiences a feeling of freely deciding, why not assume that it might be that all our value decisions which are accompanied by a feeling of free choice are illusions? Indeed, experiments have been done in which a feeling of free choice of some action was artificially created within the subjects’ mind, even though they did not carry it out at all.202 Similar experiments, which created illusions of pain or of identification, produced similar results. In quite a number of situations a person experiences something which has no basis in the real world: We already noted that the identification of an actual object standing before us — Grandma — activates the same regions which operate whether we remember her or imagine her. To understand why this argument is far from convincing, we need to examine it on a number of levels. First, I believe that people may confuse the feeling that I decided and that I made up my mind. Before the pressing of the button in the Libet experiments, the person has a feeling that he decided, but not that he

was entirely set in his choice. A “decision” could also be a psychological expression of neural arousal, but making up one’s mind is something which is supposed to occur in our consciousness and derive from a process of judgment. The question is whether we are dealing with an illusion of choice (i.e. decision in a dilemma situation) or a non-illusory feeling of a simple decision. It’s important to distinguish between the two. Second, even if we are dealing here with the illusion of choice, we need to remember that there’s nothing new in our experiencing various forms of illusions. A fata morgana is a visual illusion, but no-one questions sight due to its existence. There is a phenomenon of phantom pains in the arms or leg, but that doesn’t mean anyone thinks that all their other pains are also illusions. The same is true of various simulations of the brain which create fictitious experiences and feelings. According to the same logic, experiments which show us that we can artificially create pain should make us abandon the belief that normal pain is caused by a wound. But is this the case? Do experiments in which the person hears voices without actually being addressed by someone in the objective world make him abandon the assumption that when he hears a voice it has an origin in the real world? So long as we are not inside a neuroscience lab, when we feel pain we assume that we’ve received a blow or have a wound. If so, why is it that when we feel we are choosing, even if it doesn’t happen inside a neuroscience lab, we have to state that it’s an illusion? I can artificially cause someone to believe in incorrect laws of physics. Does that mean that the laws of physics we use today are incorrect? This argument leads us into the sort of skepticism few take seriously. If so, the fact that we can create the illusion of some feeling does not mean that this feeling will always be an illusion when it arises. I believe that it was Henry Kissinger who said that even if someone is paranoid, that doesn’t mean they aren’t being followed. Third, another look at the issue shows that the illusion of free choice actually leads to the conclusion that we have free choice. Just as the very existence of a region of the brain whose job is to feel pain is an indication that real pain exists. At the same time, the very existence of the feeling of freedom of choice and will attests to there apparently being such freedom, as if not — how and why was such a feeling created? Where did it come from? The feeling of pain was created from the encounter with real pain. Even the ability to feel pain was created so that we can feel real pain. After this skill and type of freedom were created — illusions of pain can also form, meaning fictitious pain. The same is true when it comes to feelings and experiences of choice. These are also created from the encounter with real freedom. We feel this way because we actually encounter freedom. After the feeling of freedom exists, we can indeed create a fictitious form of it. But if we did not have free choice at all, it is unlikely that such a feeling would be created within us without some origin in reality. Therefore, the conclusion is that the phenomenon of freedom of choice being entirely fictitious seems very odd, and entirely inappropriate to how we usually treat these sorts of illusions. 182 See: B. Libet, , C. A., Gleason, E. W. Wright, & D. K. Pearl, “Time of conscious intention to act in relation to onset of cerebral activity (readiness-potential). The unconscious initiation of a freely voluntary act,” Brain, 106, 1983, pp. 623-642. B. Libet, “Unconscious cerebral initiative and the role of conscious will in voluntary action”, Behavioral and Brain Sciences, 8, 1985, pp. 529566. 183 Now it’s clear why a clock which moves at a regular speed is worthless in such an experiment, as the needed precision is much higher than the resolution is a single second. 184 Aside from the experiment of Soon and his team from 2008 which I already mentioned, that same year saw another experiment done by Matsuhashi & Hallet whose aim was to overcome the problems with the clock. Many other experiments have been conducted, of course. 185 Around p. 128 of his book. 186 In this section, I make much use of Ariel Ronen’s Master’s thesis, which deals at length with the Libet experiments and their significance. The work, supervised by Prof. Benny Shannon, is called The Freedom to Want Also When the Brain Wants: A Critical Analysis of the Findings of Libet (Hebrew), Psychology Department, Hebrew University, April 2008. Thanks to my friend Dr. Yehudit Ronen, Ariel’s mother, for the reference, which reached me after the first version of this book had already been written. Some of the relevant sources, primarily those which are the newest and most up to date, were brought to my attention by Dr. Ariel Furstenberg from the Computational Neuroscience Institute at Hebrew University, and I thank him for it. 187 See the experiment of Matsuhashi & Hallet, where the neutralization of the clock problems actually increased the effect, with the time difference being about 1.4 seconds as opposed to about 0.35 seconds by Libet. 188 See Lau et al, 2007, and an important criticism from Mele. 189 See Banks & Isham, 2007. 190 On page 61, Ronen shows that Libet himself claims that he is not a Decartes-style dualist, and is instead in favor of emergence (see p. 67 passim for a more detailed discussion). Thus, Libet espouses an approach of interactionism but without full dualism. We saw in the previous chapters that such a position is not possible in light of the physical knowledge we possess today. The main problem in this approach is that if the mental is nothing more than the collective trait of the neural, that it cannot possibly influence the neural, as interactionism would argue. After all, the claim is that there is no entity outside of the neural, and there is therefore nothing to influence it from that “outside.” 191 In the original Libet experiment, it was not an empirically based argument, as an RP cannot be identified without the accompanying action. But Libet checked this directly in later experiments, and claimed that he found that this phenomenon does actually exist.

192 This is an entirely hypothetical situation, as we do not really have a way to determine for certain that there was no veto in any of the cases where the RP awakened but was not followed through. 193 See here for the origin of this distinction: Sidney Morgenbesser & Edna Ullmann-Margalit, “Picking and Choosing,” Rationality, Choice, and Morality, 44(4), Winter 1977. 194 I already noted in chapter four in the context of the Newcomb experiment, that the group of Ariel Furstenberg is developing various choosing experiments. There is also a group at Caltech working on the same. 195 This is not related to ethics, as these experiments examined how people react, but did not refer at all to the question of how people should react or what the ethically correct reaction is. At least according to Kant, no such experiment is even possible. 196 It may be that in order to do so, we have to offer two alternatives involving action, not action and inaction, as in this case there are two types of RP, one for every option. In the Libet experiments, some did the experiment and asked the subject to decide whether to lift their right or left hand, while in the original experiment, the subject had to decide whether to lift their hand or not. 197 For a psychological discussion of these phenomena, see the book by Nobel Laureate Daniel Kahneman, Thinking, Fast and Slow, Penguin, 2011. Kahneman does not seem to deal with aspects of neuroscience and neural activity, but most if not all of his studies are definitely related to our subject (see the end of part two of this book). 198 See the fifth book in the Talmudic Logic series, which deals with conflict resolution in halacha and in general. 199 http://www.mrsmoser.com/uploads/8/5/0/1/8501319/english_11_ib_-_no_exit_-_existentialism_is_a_humanism_-_sartre.pdf 200 The validity of this proposal is conditional on there actually being a case of weakness of the will, meaning a value-urge dilemma, and that it is fundamentally different than value-value dilemmas. There is also room for the view that even the choice to choose is done after the value system is built. According to this approach, part of a person’s value system is the fact that he chooses and is not dragged along by the urge valleys in his mental topographical layout, or the determination of how much mental power he uses in these situations to realize his values, carry out his decisions, and overcome his urges. In this case, there does not seem to be a difference between the two situations, and the third argument against the Libet experiments related to timing stands in relations to these types of experiments as well. 201 This should be the real dilemma, as a dilemma in the framework of an experiment will not be seen by him as a true ethical dilemma. To create a real dilemma in an experiment is an ethical problem in and of itself, and this will not be accepted by the Helsinki Committee — and rightly so. 202 In the scientific context, see: D.M. Wegner, The Illusion of Conscious Will, MIT Press, Cambridge, MA, 2002. In the neural context, see: “Movement Intention After Parietal Cortex Stimulation in Humans,” Science, 324 (5928), May 8, 2009, pp. 811-813. The article also appeared online here: http://www.sciencemag.org/cgi/content/abstract/324/5928/811.

B. Moral Dilemmas In this part of the chapter, we will examine decisions which include judgment (choosing), unlike the Libet experiments which only involve mere “picking” without judgment. We will deal here with the distinction between emotion and intellect and brain and intellect, primarily in the context of moral decision-making. In chapters three and eleven, we dealt with moral decisions and determinism, and we will return to the subject next chapter as well. But here I wish to tie this to the picture arising from the neurosciences, to round out the discussion we had regarding Libet’s experiments.

On Emotion, Intellect, and Brain There are two different questions which researchers and philosophers deal with in this context, which can be easily confused: (A) Is a moral decision arrived at through emotion or in a rational-intellectual way? The question of emotion vs. intellect deals more with the issue of determinism, even if not precisely. (B) Is the moral decision made in the brain or in the intellect? This is a question of physical vs. mental, and it actually has to do with the question of materialism. As we showed, there is a close connection between the two issues of determinism and materialism, and it is therefore no wonder that there is a connection between these two questions. The first question deals entirely with the mental context, or qualia, as both intellect and emotion are mental functions. We can of course translate this question into physical terms, and ask whether these decisions are made in the brain region responsible for emotional functions or in the region responsible for the functions of rational decision-making (we did a brief survey of these regions in chapter twelve). In any event, the first question in both its variants does not deal with the line between the physical and the mental, but rather is deliberating between two options within the physical or two options within the mental. The second question, by contrast, deals with the relationship of the physical and the mental, meaning whether the physical determines the moral decision or there is a conscious, freely made choice. I dealt with the first question in the third book of my quartet Human as Hay. Here I only refer to it because quite a few of those who discuss the issue do not distinguish the two questions, and therefore arrive at mistaken answers. Why is this question of whether emotions have weight in our moral decisions and judgments not relevant to this discussion? Because if the answer to the second question is indeed that these decisions were made mechanically in the brain as a neural process, then its identification as either emotion or intellect is irrelevant. The determinist-materialist sees both emotion and thinking as appendages or epiphenomena to neural events; the neural network determines the decision, and the emotions or thoughts are merely mental events which accompany this process or emerge from it. So what difference does it make if the experience accompanying this mechanical calculation is the illusion of emotion or the illusion of thought? If so, only if the answer to the second question is that these decisions are made by the intellect, not the brain, can there be significance to the question of whether it is emotional or intellectual.

On Brain, Emotion, and Morality: Back to the Moral Responsibility Question Very intensive studies have been conducted in recent years on the significance of emotion in our moral decisions (Antonio Demasio wrote quite a few books and articles on the subject). American neurologist Michael Koenigs and his colleagues, like many others, are studying the effect of emotions on our moral decisions, both among healthy and sick subjects. The people Koenigs is testing are the people who due to physical damage to a particular region to the brain remained without emotions. In an article he published with his colleagues in Nature in March 2007, they report that brain damage in the emotional part of the brain leaves us with only a limited ability to conduct moral judgment. Test subjects lacking emotions morally judge in a cold and logical manner, and therefore express different positions than healthy subjects.203 To examine these questions, the researchers turned to the trolley and bridge examples. They found that while the healthy subjects tended to reject the active option of pushing the fat man off the bridge, it turns out that those whose emotional region was damaged preferred to push the fat man off the bridge and kill

him in order to save the five people tied on the track. Demasio and Koenigs explain this based on their decision-making being cold and emotionless.204 Jörg Blech and Rafaela von Bredow write in their article “The Sixth Sense,” 205 that the debate today touches on legal and meta-legal questions: One of the central questions in law is the extent to which a person is responsible for their actions. The basis for the entire modern system of law is punishment matching the guilt. But now researchers argue, among them neurologists like Wolf Singer from Frankfurt, Gerhard Rith from Bremen, and Hans Joachim Markowitsch from Bielefeld, that this principle is based on an error. According to them, free choice between good and evil is an illusion, and the evil deed is therefore solely the result of a biological process. ‘All our actions are determined via communications in our brain,’ claims Markowitsch. Sociologist Jan Philipp Reemtsma from Hamburg opposes this position: ‘The assumption that everyone is responsible for their actions is the condition for the existence of the justice system and the moral act. We cannot base the legal system on biology.’ The debate between these two positions occupies almost every respectable research institute in the field today. At the Open University in Berlin, philosopher Matthias Mahlmann and starting jurists discuss the issue ‘How does cognitive neurology put the “culture of law” to an extreme test.’ Seminars like ‘Paradigm changes in criminal law — advancing neurobiology,’ are being conducted today with the participation of neuroscientists and jurists, and lawyers in the United States are already using brain scans to try and claim their defendants’ limited responsibility. There is an implicit assumption here that neurology is putting our legal and moral assumptions to an extreme test. Is that so? First, it’s important to understand that this is not a philosophical-legal argument but a claim belonging to the natural sciences. If these are the scientific findings of the field of neuroscience, it’s not clear how a sociologist can oppose this and argue that the justice system should not be based on biology. If it has been scientifically proven that we “make up our minds” due to neural processes which are forced on us, then the very fact that we cannot base a legal system on this foundation is not an argument but a fact, distressing though it may be. Should we therefore change our scientific conclusions? Of course not. Moreover, we already saw in chapter three how the judge, the citizens, and even the legislators are compelled to act the way they do, and of course those who debate such questions are also compelled to do so, and these arguments therefore appear prima facie meaningless. Later in the article, the two authors discuss morality as a sort of universal internal grammar (like Chomsky’s inborn grammar), and examine its biological origin. But this of course does not touch on the question of the moral validity of these judgments and the question of condemnation and sanctioning of those who do not act properly. Whatever the origin of moral norms, so long as matters are mechanically handled by the brain (regardless of whether it’s the emotional or rational part thereof), we cannot impose moral or criminal responsibility on anyone. We should also note that the two questions we presented above are being confused here. Even if our emotion is involved in decision-making, this still does not mean that it is not a neural decision which is forced on us. Demasio and others identify the two questions as the same, and assume that if emotion is involved in decision-making then it was freely made, but even if it is a cold calculation then we are dealing with a deterministic process. However, as we saw above, this is a conceptual error. Whether we are dealing with emotion or intellect, we can ask whether these mental phenomena are the consequences of neural events and can therefore not be held morally responsible, or perhaps these are autonomous phenomena, and then we can impose moral responsibility, whether the decision is emotional or intellectual. Later in the article, the authors describe the dependence of our moral principles and judgments on our culture. As far as they are concerned, this connects to the argument that we don’t make these judgments autonomously, and we are therefore not truly responsible for them. If so, there are two main factors in our moral judgment: Inborn morality and social-cultural influences. At the end of chapter twelve we mentioned the work of Greene which shows that a healthy person deciding an ethical conflict experiences a tension between the emotional-social region and the cold logical one. This is another indication that our decisions and conflicts reflect neural events and nothing more. We must therefore now ask: Does neurology really leave us in this situation? If we really do see that our

judgment depends on brain structure, then it is clear that another brain will lead to another judgment, and our brain structure is of course not entirely in our hands. Moreover, someone whose brain was damaged is certainly not responsible for having a different brain. But even a healthy person who makes moral conclusions and judgments of one sort or another does so because of the structure of his brain. If so, what place is there for moral responsibility? We dealt with the a priori perspective on this question in chapter three. Now we return to it after discussing the neurosciences. Has our discussion changed? Do the conclusions described above undermine the libertarian approach? My answer to these questions is firmly in the negative. In chapter two, we pointed to every human decision being based on two levels: (A) The natural urge or tendency which his topographical environment applies to him (to slide into the valley or not climb the mountain); (B) His autonomous decision (whether to surrender to the urge or act differently — and nevertheless climb the mountain). We saw there that statistical studies on this question have no significance, as it was clear to all long before the neurosciences that the topographical layout around us slants the results statistically. The big question is not whether the topography slants or influences the results, but whether it dictates and determines them. The question is whether complete knowledge of the topographical layout amounts to complete knowledge of future moral decisions by a specific person, not just statistical knowledge of a group of people. We cannot find an answer to any of these based on the experiments described here. If the conclusion that the brain and the environment statistically influence decisions, then this is a fairly trivial result. But if the conclusion is that the brain and the environment dictate results (determinism), then this will truly be a revolutionary and very important conclusion. It’s just that the findings do not support such a conclusion. There is room to conclude here that we should sometimes reduce criminal responsibility and impose a lesser sentence on the criminal, but this has been clear and obvious since the beginning. A person’s educational and cultural environment influences his behavior, and this tends to be considered when he’s being judged. For instance, a legal system knows to be lenient with a criminal who grew up in a problematic environment or a person who operated under serious pressure. By contrast, the deterministic conclusion regarding the complete lack of responsibility for actions, meaning seeing all behavior as deterministically set by the circumstances is certainly an innovation, but this innovation does not at all emerge from the findings. Once again we see that those who make various deterministic arguments have a partial and incomplete picture of libertarianism, practically a caricature. The more complex and complete picture presented in chapter two is meant to prevent such errors, and here we see once again that it does the job. Another common error which we dealt with in chapters three and eleven can be seen later in Blech and von Bredow’s article: The German Kant (1724-1804) presented a different approach. His categorical imperative (‘conduct your actions only based on that practical rule which, upon accepting it, you would wish to also see it become a general law‘) was to be based on reason. But Mahlmann believes that in the end Kant’s philosophy and the new findings of neurology can find a common denominator. ‘Does inborn morality not explain Kant’s argument regarding our tendency to act rationally?’ he wonders. The error of the writers — that there can be a congruence between inborn morality and Kant’s argument — shows that they did not understand the latter. The entire essence of Kant’s moral approach is that a person must choose to do what he wishes to be a general rule. Kant is not a biologist, and he does not of course intend to teach us how a person effectively chooses (according to the determinist), and not even how he tends to choose (according to the libertarian). Kant intends to explain how a person should choose. He is speaking of an ethical norm and not a biological and/or psychological fact. Moreover, there is importance in terms of Kantian ethics for the motives and autonomy of the person and not just the content of his actions. Therefore, there cannot be a congruence between a deterministic approach and Kantian ethics. Later in the article, the authors describe the work of de Waal, which shows the existence of moral sense in monkeys and other animals. Various evolutionary explanations are made for this. The libertarian will say that this comparison between the behavior of monkeys and the behavior of human beings contains a double error: (A) Even if it appears that the behavior of animals is similar to human beings, this will not serve as proof for the question of determinism, as the similarity is solely statistical.206 (B) We are not dealing here with morality but good behavior (mimicking morality). This mistake appears throughout the article, mixing between the ought (how one should behave) and the is (how creatures behave). This is known in philosophy as the naturalistic fallacy, which was defined and described in chapter eleven, and we

will now return to it from the perspective of neuroscience.

Back to the Naturalistic Fallacy Chapter three dealt with aspects which demonstrate the problems in the deterministic approach in terms of moral judgment. Chapter eleven dealt with this issue from the materialist angle, and we already saw that the two questions are profoundly connected to one another. In that chapter, we presented the “naturalistic fallacy”: Hume and Moore argued that an argument which derives an ethical, normative, or judgmental conclusion from facts is an invalid argument. For instance, if I conclude that a wall is beautiful because it is cream colored, I have fallen for the naturalistic fallacy, as the assumption is a fact and the conclusion is judgment. There is an additional hidden assumption that cream is pretty, but this is a judgment, not a fact. This is when it comes to aesthetic judgment. We also saw that when it comes to ethical judgments, we cannot base a judgmental argument on mere facts. If a person killed someone that does not mean he is evil. The claim that he killed someone is a fact, while the conclusion that he is evil is a judgment. Here, too we must add the hidden assumption that those who kill (or murder, which is already a loaded expression) are evil. As I already noted there, we can already see an initial problem in the determinist-materialist thesis based on this fallacy, and especially when it comes to explain morality: It pegs judgments to facts. In fact, it does not even recognize the category of judgmental arguments. As far as it is concerned, there is no fundamental difference between a factual claim and a judgmental claim, as both are done neurally. We will now return to this point, but from the angle of the neurosciences as they have been presented so far. In the materialist-determinist picture, there is some physical state which causes our brain to conduct a calculation resulting in some argument, judgment, or fact. This entire process is deterministic. Moreover, the more detailed mechanism of judgment is composed of some external (stimulating) trigger which reaches my sensing organs, which transfer the information to the brain, which then does a calculation based on electric currents and physical-chemical processes and makes a conclusion which is the state of the neural network in the brain. This is the judgment. This mechanism shows us that judgment is the mechanical result of facts, and is in effect the facts themselves. We therefore see that norms are nothing more than epiphenomenal mental appendages to brain states or structures. In essence, it is more correct to say that there are no norms in the world of the determinist, and that this is but another name for a trait of certain brain structures. Thus, judgment in the determinist picture is also nothing more than a factual claim. We can put it differently as follows: Factual statements are the result of observation of factual circumstances. The brain creates these factual arguments based on observation in the form of mechanical calculation. But as we learned from Hume, judgments are categorically not the result of facts. What creates them, then? Is there a physical state which is not the result of another physical state? This contradicts the principle of causality. It is therefore natural that what creates judgments is the physical state of the brain (our system of norms or values), but such a state is also a fact. Judgment in the world of the determinist is therefore a logical or physical result of facts, and the naturalistic fallacy is thus wiped off the philosophical map. Indeed, there is no such map, as philosophy is a subject involving intellects and not brains, and the intellects of course function through their brains. Here we of course have to ask what nevertheless creates the insight at the basis of the naturalistic fallacy. If there is indeed no distinction between the two types of arguments, why do we nevertheless feel that there is such a distinction? The determinist will of course say that this is another illusion on an ever lengthening list of our insights which the determinist wishes to convince us are nothing more than illusions. But if the fact that the naturalistic fallacy is an argument which fails our test of feeling, and if our feeling is not an illusion but a correct philosophical statement, this means that there is a categorical difference between facts and judgments and norms. And this means necessarily that not everything is matter and not everything is neural processes. Only in the dualist picture can we get such a categorical difference. For precisely the same reason, a monkey which helps another monkey is not moral. It does not do so out of commitment to the imperative of morality but due to its nature, and actions not based on decision are not moral actions, by definition. Also a person who helps his friend under hypnosis is also not moral, as he did not decide to do so.207

203 I cannot but mention a wonderful book called The Rosie Project by Australian author Graeme Simsion. The book deals with a man with symptoms of Aspergers (never openly stated on the book, but hinted at). This is a man lacking emotional abilities and social skills, and the story is told from his own perspective. The book provides a picture in which these people act in an extremely rational manner while the normal people act in an illogical manner and for very strange and foreign reasons. The obvious conclusion is that the people with Aspergers are the only healthy ones in our world, and what makes us special is the degree to which we suffer from anti-Aspergers. 204 There’s no need to note that this determination of limited moral judgment has nothing to do with science. This is a value position (and an incorrect one, at that) of the experimenters and nothing more. It points to the experiments assuming that moral judgment must have an emotional tendency, and that those who make moral judgments coldly are “limited.” If so, beyond the factual statement that healthy people have emotions which play a part in moral judgments and decisions, there is a problematic ethical assumption here. If there is a person who is unable to show empathy (due to a brain flaw or some other reason), then he will certainly behave differently in the moral sense towards others. But this is a deficient consideration, as it may be that he does not understand how much others are suffering. It’s clear that those who do not understand how others feel will make decisions differently, but this does not point to different, or flawed values, only partial factual information which may affect the moral decision. This is not what people mean when they say emotion plays a part in moral decision-making, and we will go further into this issue at the end of chapter fifteen. 205 Jörg Blech and Rafaela von Bredow, “The Sixth Sense,” Der Spiegel, 2007 206 See more on this in our discussion at the end of chapter two on the comparison between bees and human beings regarding decisionmaking. 207 In this context, it’s worth noting the work of Frans de Waal, Good Natured: The Origins of Right and Wrong in Humans and Other Animals, Harvard University Press, Cambridge, 1996. De Waal deals with the morality of animals in this book, and as its name suggests — fails by committing the naturalistic fallacy when he comes to understanding the term “morality.”

Chapter Fifteen THE WONDERS OF THE DAMAGED BRAIN The pleasures and the feelings of the right half of my brain are too great for the left half of my brain to find the words to report to you on them. Roger Sperry, 1981 Nobel Prize winner In the first part of the previous chapter, we dealt with the Libet experiments and their significance. The second part of the previous chapter briefly went back to the issue of moral decision-making. The two parts combined pointed to the Libet experiments and other similar results from the neurosciences not producing any significant scientific indication for the materialist-determinist thesis. But our journey in the scientific context is not yet over. We have one additional context to look at where empirical claims are made on the issue of determinism, really more related to the question of materialism than anything else. These are questions which raise findings from studies on damaged or split brains. These issues are the subject of the present chapter. Let’s go back in time, to the phenomena and studies which began already in the nineteenth century, continued in the fifties and eighties of the twentieth century, and which reach our own day. Our aim here is to examine the relationship between the brain and the soul — or the physical and the mental — by observing damaged brains. These reflect a variety of interesting phenomena which teach us many things about the operation of a functioning and healthy brain. For instance, emotional deficiencies or changes in worldviews and outlooks, in the wake of a split brain or other forms of damage, demonstrate phenomena which quite a few researchers claim have direct consequences for the question of materialism — and therefore also, at least indirectly, for the question of determinism. 208

The Curious Case of Phineas Gage In 1848, while Phineas Gage, a young, 25-year old American manager, went out to work on his project to set up a rail line, he certainly could not imagine that he was about to become world famous (It’s likely that if he knew, he would have forgone the dubious honor and not shown up for work). While Gage was trying to blow up some stubborn rock with explosives, a metal bar about a meter and a half (about five feet) in length flew into his face underneath his left check. The bar penetrated directly into and through the frontal lobes of the brain until it came out again. Amazingly, Gage did not die and didn’t even lose consciousness. He was taken on his feet (!) to a carriage which took him to town. There two doctors treated him, Dr. Williams and Dr. Harlow, and it seemed that he was recovering and healing, despite the mortal wound in his head. Yet at the same time as he was showing himself to be a medical miracle, Gage started to demonstrate sharp changes in his behavior and character. He couldn’t plan ahead, and he started to be impulsive and unsociable. He turned from being a measured and responsible person to an intolerable one who couldn’t function in society. He started to curse loudly, something he would never allow himself to do previously. It’s no wonder that his friends said of him that “This is no longer Phineas Gage.” In his state, he could no longer go back to work for the train company and began making a living as a wandering carriage driver. Twelve years later he died, in 1861. His story has become one of the most famous in the history of medicine. Doctors at the time understood that the damage done to his brain is what was apparently responsible for the changes in his behavior and character, but they did not make much progress in systematically studying such damage, and certainly not in treating them. The technology and understanding of how the brain works, needed for such a study, simply didn’t exist yet. Gage’s case is part of the medical myth of the neurosciences even though he did not make a significant contribution to the progress of research in this field beyond the very awareness he raised in his time. In the 1930s, Dr. Walter Freeman began to conduct operations in the form of lobotomies, meaning removing part of the frontal lobe, with the entry to the brain being done through the eyes. This process began as a method of treatment of mental patients with violent tendencies, with the declared goal being to lead to changes in behavior and character of these patients. Since then, tens of thousands of such operations were carried out, which did indeed demonstrate profound changes in behavior and function: In some cases it did lead to substantial improvement, but there were quite a few cases where the reverse happened; some led to the deterioration of patients to the point of suicide. Today, it is accepted in the medical community that such operations are usually a serious mistake, and they were therefore almost entirely stopped. However, in cases of severe epilepsy, doctors even today sometimes carry out an operation known as a corpus colostomy — a removal of the corpus callosum. Still, today they do not remove the entire corpus even in the most extreme cases, but only part of it. All such cases demonstrate the connection between parts of the frontal lobe and forms of human behavior and character. Similar phenomena were tied to other lobes. As noted, all this was observed before there was a model or detailed explanation as to how this actually occurs, and the explanations have been coming in and forming slowly only in the last couple of decades. Gage’s brain was removed from its grave five years after the death of Dr. Harlow, one of the two doctors who treated him, and it has since been preserved at Harvard Medical School. A hundred and fifty years after his death, neurologists Antonio and Hanna Damasio (who deal a lot in questions related to morality and changes in conduct and the brain) photographed Gage’s skull from all angles. They examined it via x-rays and formed a three-dimensional model of his damaged brain. With the aid of a computer simulation, they succeeded in reconstructing the precise path of the metal bar through Gage’s brain tissue in the frontal lobe until it exited his skull. This reconstruction is significant, as the tools making it possible to examine the connection between Gage’s reported changes in behavior and the parts of his brain which were damaged are being formed.

Consequences for the Relationship Between the Physical and the Mental It is clear already from the general description that there is a close connection between behavior, and effectively character and personality, of a person and his brain. Prima facie, it would appear that the structure of the brain dictates our personality, behavior, and other mental characteristics. The path is short from here to identifying personality with the brain itself. Many claim that these findings point to mental events, and perhaps our personality in general, being the products of physical traits of the brain and nothing more. After all, if a different brain means a different personality, then why should we assume that personality and the brain are two different things? The conclusion many researchers derive from these phenomena is that we contain no additional component; if matter dictates and unequivocally determines the spirit, then the spirit and mental phenomena are at most emergent phenomena, meaning phenomena which emerge from the material which forms the brain. But we have already seen that such findings cannot be unequivocal. There is certainly the possibility that some change in the brain can engender changes in the “topographical map” surrounding a person and in which he operates. We already saw that such changes influence behavior, just as any physical event affects it. When we have a wound, for instance, it hurts. The wound is a physical event, but the pain is a feeling, meaning a mental event. Why would the influence of these brain changes on behavior and feelings surprise us more than the trivial influences of a wound on pain? Because pain resulting from a wound is not the result of any conscious decision. It is an entirely mechanical consequence of the wound, and we are therefore entirely unsurprised when the physical clearly determines the mental. By contrast, we would very much like to think that outlook and decision-making are voluntary, autonomous actions done by the individual, and it is therefore difficult for us to accept the argument that they also reflect nothing more than physical structures. It’s important to understand that these influences are not the same as regular decision-making. In chapter two, we spoke of the topographical layout in which a person operates, noting that some physical event may create a tendency in that person to, say, react with violence. He can certainly realize it, but he can also restrain himself — “vetoing” this urge, as Libet would put it. Bottom line, a person’s action is the consequence of his decision. But in cases of dramatic changes in the structure of the brain, it is not the person who decides to react to an event which happened to him one way or another. Here we are speaking of an automatic result of something which happened. The person who was injured did not decide to become a different person, he was simply turned into a different person, as the friends of Phineas Gage described We are not just dealing with changes in character; we are speaking of reports of changes in values and worldviews, which ostensibly point to both worldview and value decisions of a person being mechanically determined by the brain. This means that there is no stage of information processing and decision-making after the physical change in the brain. The person does not, for instance, decide to become a believer rather than an atheist, or a capitalist rather than a socialist. It just happened to him. Does this not point to an identity between brain and intellect, or between the physical and the mental? Can we still speak in such cases of judgment and decision-making?

Split Brain: Background Before we answer these questions, let’s discuss the fascinating phenomenon of the split brain for a bit. This issue also revolves more around materialism, but the determinist question also arises from it. Two researchers from Caltech, our man Gazzaniga and his partner in receiving the 1981 Nobel Prize in Physiology and Medicine, Roger Sperry, began to study people whose corpus callosum had been damaged, this being the part of the brain which connects the two hemispheres. This phenomenon allowed the researchers to examine the functioning of each hemisphere separately as well as the significance of the connection between the two hemispheres. We should note as an aside that the study of the split brain is problematic, as an initiated detachment of the corpus callosum is an immoral act, and it is of course forbidden to do so solely for research purposes. But it is still possible to study cases in which the callosum was damaged due to an accident or disease, or when medical circumstances require that it partially or fully be removed (as noted, this is indeed done sometimes in serious cases of epilepsy). This importance of the corpus callosum derives from the fact that both hemispheres have different functions, and therefore our ability to act harmoniously and in a coordinated fashion depends on the connection between the hemispheres, this being the corpus callosum. In chapter twelve, we saw that the left part of the brain is responsible among other things for verbal functioning (speech and writing), meaning it is responsible for our expressive skills. Naturally, this is the part which is also stronger when it comes to logical functioning and linear thinking. By contrast, the right side of the brain deals more with spatial functioning, and it is also the stronger part when it

comes to holistic, creative, emotional, and intuitive functioning. We should also note the fact that the connection between the brain and the body is crossed: The right part of the brain operates the left part of the body and is activated by it, while the left part of the brain is connected to the right part of the body. In practice, this means that the left hemisphere sees via the right eye, hears via the right ear and feels via the right hand, and it also moves the right hand and the right leg. The right hemisphere is related in the same manner to the left side of the body. It turns out that both hemispheres have similar functions in many cases; they just handle the same receptors in different ways. Quite a few of our functions require control which determines whether there will be a combination of the two hemispheres or whether one will be given control of the other. The simplest example is spatial vision, which is formed by both eyes together. Even the pinpointing of the origin of a sound via hearing combines the two ears. Our thinking also combines logic and rigid, recursive thinking from the left side of our brain with intuition and creativity from the right side. 209The conclusion is therefore that this is apparently one of the main roles of the corpus callosum, which connects the two hemispheres (although there are also weaker indirect connections, such as via the anterior commissure).

Split Brain Experiments Some of the experiments which examine the relations between hemispheres are done with a healthy brain. There are techniques in which the tester transfers information or gives instructions to just one part of the subject’s brain. Other experiments are done on people whose brain was damaged or split. Quite a few experiments are done on such subjects, and the results are wondrously diverse and interesting. 210 In an experiment done by a group of researchers from Dartmouth, headed by Michael Gazzaniga, pictures were presented to the left and right hemisphere of a patient with a split brain, after which each hemisphere was asked to fit a picture which would accompany the one showed it in the beginning. For instance, the right hemisphere of the subject — meaning the left part of the vision field, with the right eye being blocked — was shown a snow covered house. As a matching picture, the person chose a picture of a spoon for loading snow, whose relationship to the first picture is entirely clear. Now the left hemisphere of the same person — via the right side of the vision field — was shown the foot of a chicken, and he chose the head of a chicken — again, entirely logical. Afterwards, the researchers asked the patient to verbally explain his choices. The left hemisphere was the only one which could express itself, as it was responsible for speech functions, but it could not know why its colleague on the right chose the snow spoon, since the information on the snow covered house did not reach it due to the damaged corpus callosum. The explanation of the split brain patient was wonderfully creative: The head of the chicken was chosen because it fits the foot of the chicken, while the snow spoon was chosen because a loading spoon is needed to clean out the chicken coop. The left hemisphere acted as a translator for the person’s worldview, and simply invented a story which would fit his existing information. Scholars argue that the left hemisphere usually controls the right one; moreover, it fights to control it and preserve that control. The phenomenon we have described reflects the attempt of the left hemisphere to preserve its control. To do so, it completes the information gaps with creative explanations, whose connection to reality is very slight. Does this not remind you of Descartes’ demon we spoke of in the second part of chapter four? Now you have it before you: Our left hemisphere. In another experiment, the split brain subject was put before a screen which was also split in two. In the first stage, the right side of the screen showed a spoon. In this situation, the subject reported to us that he saw a spoon. When we move the picture of the spoon to the left side, the report is that he doesn’t see a thing. Still, when he is told to lift an object similar to what appears on the screen with his left hand (the spoon, which he ostensibly cannot see), he will take a spoon. When we ask him what he’s holding in his left hand, he will not be able to say. What is the meaning of these strange phenomena? As noted, the left eye is related to the right hemisphere, and vice versa. When the spoon appears on the right side, it is the right eye which notices it, and the information therefore goes to the left hemisphere. This hemisphere, being responsible for speech, reports to us that the subject saw a spoon. But when the spoon appears on the left side of the screen, the left eye sees it, and the information is transferred to the right hemisphere. Since there is a disconnect between the hemispheres, the information on seeing the spoon is not transferred to the left hemisphere, and it cannot describe to us that it saw a spoon. By contrast, the left hand is also connected to the right hemisphere, and therefore when we tell it to lift an object with its left hand which is similar to what it saw, it will indeed take a spoon. But the information that this is a spoon lies in the right hemisphere, and this cannot transfer the information to the left side, which as we noted is responsible for speech, and can therefore not report to us on this when we ask it what it’s holding. In another experiment, a split brain subject was presented with a split picture, with a man on the right half and a woman on the left half, like the following picture:

If we ask him what he saw, he will of course say a man. Why? Because the speaking part reporting to us is the left hemisphere. The information sent to it comes from the right eye, which sees a man. By contrast, if we ask it to point to a similar picture to what he sees, he will point to a woman, because pointing, in contrast to speech, is done by the right hemisphere. 211 Similar phenomena occur regarding hearing. In the experiment Gazzaniga reports on in his book, a teenage boy with a split brain is examined. He is asked out loud, so that both his ears hear: “Who is your favorite…,” and afterwards adding in a soft voice, so that only his left ear can hear “…girlfriend?” It turns out that the youth giggled in embarrassment, as though he heard the whole question, but his answer was: “I didn’t hear the last word.” Finally, strangely enough, he used the left hand to spell the name of his girlfriend. Again, the explanation is that the left ear and the left hand are connected to the right hemisphere, and vice versa. Some of those who have split brains suffer from troubles created by their own hands: Sometimes their right hand tries to button up their shirt while the left hand is trying to open it. The right hand places objects in the shopping cart, and the left hand returns them to the shelves. Such a person can read a book in his left hand, when suddenly this hand itself will remove the book in the middle of reading; reading, like speech, is the kind of activity done by the left hemisphere, and meantime the right hemisphere gets bored and orders the left hand its connected to to put the book down. One of the subjects, who had speech capabilities on his right side as well, was asked what profession he wishes to have. The right hemisphere answered race car driver; the left preferred to be a draftsman. Even his opinion on the American President at that time (Nixon) was literally split: The left hemisphere supported him and liked him, while the right hemisphere had reservations about him. We should note that we are talking here about political and value judgments, not just character and behavior. So, it appears that we have two different creatures within the same skull. They see things differently, think differently, report differently. And sometimes even have different political attitudes. We saw that when they are detached from each other due to damage to the corpus callosum, they act separately, and when there is a connection between them they create a united front in some way. The functioning of a healthy person is some sort of average or connection between the two, and sometimes even one side controlling the other (usually left over right).

Is the Physical the Same as the Mental? If our mental world is a phenomenon which occurs in the soul and is not determined by our physical form (i.e., the brain), then an obvious conclusion from the split brain experiments is that they have two souls. If the corpus callosum is cut, does a person get another soul put into him? If he was a Democrat before, then the split put a Republican soul in him, and vice versa? It sounds like fiction. It’s not reasonable to assume that souls enter the body on demand, every time we wish to explain a result which does not fit the worldview of the libertarian dualist. These are ad hoc additions which appear very problematic. The explanation that there is no such thing as a soul, and that the mental is but a trait of the material, is therefore simpler and more logical. 212 All this ostensibly teaches us that there is an absolute identity between the intellect, or the soul, and the brain. The natural conclusion is therefore a materialist one, which means that people have no spiritual component, or that it at the very least is nothing more than an epiphenomenon determined in a one-sided manner by its material dimension (or perhaps only emerges from our physical brain). Therefore, when our brain undergoes a change — in this case splitting in two — it’s no wonder that a mental duality also forms. Our traits are determined by the brain, or to be more precise the relevant part of the brain, and therefore the number of our “souls” is as the number of our brains. In other words, the person is a Republican or a Democrat according to his brain, and not as a result of choice or values he decides to believe in. The article of Eran Zeidel (we should note that this is from the 1980s),213 ends with these words: The findings describe fit the opinion that the intellect is the brain, with its traits. Since each side of the brain has a more or less complete conscious system, each halfbrain can be seen as a separate intellect. And since we believe that the two halves of the brain are different from each other at every level of the system, they thus constitute two separate intellects. Are the two hemispheres conscious? And if so, are these separate from each other, capable of acting at the same time? Can it be that each half-brain will operate separately when it comes to free choice or moral responsibility? And what will happen if a conflict forms regarding value judgment? When one hemisphere takes over

does the other bear responsibility, as well? […] A nice definition of consciousness refers to the complexity of perception, cognition: A cognitive system, or an animal, has consciousness if it has enough mechanisms to be able to represent its own image, represent its surroundings, have concepts of the past and the future, choose and plan. According to this definition, the meaning of the theory of the independence of the hemispheres is that the two hemispheres can maintain separate consciousness at the same time. In fact, this should not shock us so much, as we are witness to a bevy of beliefs, opinions, and behaviors which contradict each other and exist alongside each other in the same person, without this preventing him from arriving at decisions in practice and accepting moral responsibility, as a unified personality. Therefore, so long as the channels of communication are functioning properly and there is hemispheric superiority, there is no fear of new challenges to ethics, logic, or the theory of consciousness. In situations of deficiency, when there are inter-hemispheric conflicts, accepted approaches of knowledge and morality are reexamined. Perhaps particularly embarrassing is the recognition that in a functioning situation, where there is hemispheric independence, a kind of split brain symptom reveals itself within us. Zeidel is wrong here in the philosophical context. If the brain does indeed determine the mental, then there is a problem in terms of ethics and consciousness, and moral responsibility entirely loses its meaning (see above in chapters three and eleven). He claims that so long as there is a proper connection between the hemispheres, this does not prevent a person from accepting moral responsibility. But the question is not whether he accepts moral responsibility — a functional, practical question — but whether it can be imposed on him, which is a matter of the philosophy of morality. Put simply: Does he deserve it? The roots of this error lie in the fact that the problem regarding moral responsibility is not the consequence of the duality. For instance, we can speak of both halves sharing moral responsibility, just like the responsibility imposed on a business with two partners or some legal corporation. The problem runs deeper: Is someone whose behavior is entirely mechanically determined by physical brain-related events and structures even subject to moral responsibility? I explained in chapter three that even in the framework of the deterministic picture, we can speak of moral responsibility by analogy, meaning the handling/correction of behavior via moral condemnation, with condemnation and sanctions in this picture being meant to fix the person (via the constraints which mechanically lead to a change in his brain). In this functional sense, in contrast to the essential meaning of moral responsibility, it is a sort of garage or lab for fixing brains, just as we would fix a car or drill which is broken or damaged. The moral responsibility Zeidel is speaking of is nothing more than the mechanism of imposing sanctions to mechanically change human brains, meaning their training. I already noted that not long ago, this would have been called “re-education” (and is indeed still called this in Communist China). To sharpen these issues, let’s take a look at the subject of subliminal advertising, which arouses quite a few debates and a lot of public discussion. Subliminal advertising is based on messages sent to us without our noticing them. A person who hosts a TV show and wears something or drinks something, and this sends us a message about it. Sometimes we are broadcast pictures at a very high frequency, which we cannot be aware of, but it affects our brain without conscious scrutiny. But in light of what we’ve described here, this debate seems absurd. After all, regular advertising operates in precisely the same way as subliminal advertising. Even when it comes to normal, aboveboard ads, there isn’t really rational scrutiny on our part when it comes to our decisions and desires. What is happening around us shapes our brains, and “us” through it. We therefore see that Zeidel’s argument, according to which there are not ethical problems from the emerging picture from split brain research, is based on a philosophical misunderstanding. But beyond this misunderstanding, the essential argument of Zeidel and his colleagues stands: If the split brain does indeed automatically split the intellect, character, worldview, values, and consciousness, then it’s very likely that all of these are nothing more than traits of the brain. It would appear that libertarian dualism faces a serious challenge here.

The Unity of Consciousness Gazzaniga, in the third chapter of his book, 214wonders how the collection of all the mental skills and activities done in the brain combine into one consciousness. What causes every one of us to feel like a complete person? What unites all these functions into a single collective entity with a united consciousness? In effect, he is asking why we are not filled with noise and do not feel another sound and another function every moment, but instead feel as though everything functions harmoniously as though it were one wonderful unity? We can really ask how our “I” is created if we are nothing more than a collection of molecules neurons which create a collection of mental functions which are created by various activities of different parts of the brain. His response is that the left hemisphere is what creates a feeling of unified consciousness within us by uniting all the thousands and millions of voices gushing within us and fusing them into a single unity. We saw at the beginning of this chapter that our left hemisphere creates illusions and invents explanations to synchronize the picture we are viewing, even if it is done in the right hemisphere, thus leading us to coherence and internal peace. According to Gazzaniga, this is also the cause of the unity of our consciousness — nothing more than another fabrication of the left hemisphere, our interpreter. But another problem arises here: The very presentation of the question contains the assumption that we should be feeling an enormous polyphony of many voices and tendencies and stirrings, while what actually happens is entirely harmonious. But who is the “it” who should be feeling all this? What consciousness should experience this polyphony? Is it some “I” which all these are a part of? We can put it this way: What does it mean when we say “our hemisphere”? Whose? This interpreter is my function. But who is the “I” which is using the interpreter as an aid? It’s important to understand that in all such statements, meaning those which deal with the consciousness which absorbs its own unity, consciousness effectively serves here in two different meanings: It is the absorbing subject while at the same time serving as the absorbed object.215 The explanation which Gazzaniga proposes for the left hemisphere may provide us with an explanation for the absorbed object, as it tries to explain to us how our consciousness is absorbed by us as united (with this simply being a fabrication created by our clever interpreter, the left hemisphere). But how is there a “we” and “by us”? Who is it who receives the information from the interpreter? After all, the word “I” appears in these sentences in both meanings at the same time: “I” am both the subject of the statement and the one who expresses it. My argument is that even if the first “I” is a fiction, the work of our left hemisphere, the second cannot be a fiction. This is no more than a version of Descartes’ cogito argument presented in the prologue: I absorb (even if what I absorb is polyphonic), ergo I am one; there is one united consciousness which experiences (or doesn’t experience) the polyphony. Moreover, what we are encountering here is nothing more than another reflection of the fool’s cycle we encountered already in the fourth chapter. If our left hemisphere is indeed fooling us (who is “us”?), meaning that it is Descartes’ deceptive demon, which we finally managed to locate within us, then how do we know that this theory itself is not the result of this deception? As we saw in chapter four, the determinist assumes the existence of such a demon, and even knows that it deceives us many times (and at the most basic levels), and we now even have a clear identification of the demon in the form of our left hemisphere. So what is even left of all our insights and consciousness? In effect, even the identifier uses that same naughty left hemisphere, so how can we trust this identification? It’s not clear to me how we can speak of deceptions which create our most basic insights and observations, while at the same time believing this talk itself. Indeed, beyond the question of whether the deception itself is also a deception, it is not clear who it is deceiving. Is there some “I” remaining complete and credible which the demon is deceiving? Is the “I” itself free from this whole devil’s dance? Perhaps the determined physicalist has the answers. The simple explanation, of course, is that we may have illusions and sometimes even form fictions in our mind, due to our unrestrained left hemisphere, but this does not mean that we must necessarily give up on our most basic insights wholesale. In chapter four, we pointed to the fact that we know that a fata morgana is sometimes possible, but that doesn’t mean we disbelieve our own eyes at all times. There are also phantom pains, for those whose foot or hand were removed, yet a person who feels pain elsewhere will not ignore it or treat it as a fiction. Thus, we have rejected Gazzaniga’s explanation for the unity of consciousness, and we indeed see the unity of consciousness itself as strong indication for libertarian dualism. Our interpreter cannot be the left hemisphere itself. It is instead our “I” which perhaps uses the left hemisphere (which sometimes lets us down by deceiving us). We are still left with two types of findings which we must take into consideration: (A) Brain damaged people who change their behavior, outlook, and sometimes even their values and moral decisions. (B) When the brain is split, it appears as though there are two different “souls” with different characters. We already noted that many claim that these two types of findings ostensibly point to a materialistic identity between the neural and the mental. Is that so, though? Do these findings necessarily lead to the materialist-determinist conclusion regarding the identity between intellect and brain, with the neural determining the mental? My answer to this question, which I will expand on later in this chapter, is negative: None of this amounts to scientific proof for our discussion, which remains at the philosophical level it occupied before modern neuroscience ever began.

Reminder: Two Models Let’s go back to the two competing models: Materialism, which identifies the mental as an emergent phenomenon of the neural array, and dualism, which espouses freedom of will and its non-subjection to the laws of physics and matter (and of course the brain in general). The determinist model refers to our brain as part of the material world and the physical-causal processes which happen therein. The picture which emerges within regarding human decisions, appears as follows: Image 1: A Schematic of Decision-making in the Deterministic Picture

In chapter one, we saw that “circumstances” includes both internal and external factors. Genetics, the present structure of the brain, and all external circumstances (education, environment, home, school, &c) — all these become input for the brain, it processes it and issues a “decision,” which is then translated into electric currents which activate the relevant organs to carry it out. In other words, three different brain functions are represented in the central circle: Input and processing of data, decisionmaking, and the transfer of instructions to organs (in case of a motoric action). On the other hand, many err in thinking that libertarianism means that decision-making is entirely free, and that the circumstances around us (the topographical layout) has no influence on our choices. According to this incorrect interpretation, if there are influences then this means that the decision is not freely made, and is instead a process set deterministically by other factors, whether external or internal. Such an approach views libertarianism thusly:

Image 2: Erroneous Picture of Libertarianism — Freedom of Will

In this picture of libertarianism, decision replaces the circumstances we saw in the determinist picture. There is a decision made entirely freely, detached from circumstances (we see that the decision is preceded by nothing at all), and it activates the brain to create the needed signals to activate the organs to carry out the decision and do the act. According to this interpretation, the brain has no part in the decision; it only sees to its implementation. The decision itself is made entirely autonomously in the soul or spirit. I already noted that this is but a caricature of libertarianism which may be easier to attack but which does not faithfully reflect the approach. The error in this simplistic understanding of libertarianism is the same we dealt with in chapter two: Freedom of will instead of liberty of will. As we saw there, freedom is a state without limits. Decisions without any constraints or influences are indeed free, but libertarianism does not mean that there is no factor affecting the person’s decision or contributing to it. The libertarian would also agree that there are factual data which the brain absorbs before making a decision, as it’s impossible to ignore the fact that I need data that someone wants to hit me, before I make a decision as to how to react to that fact. But in the picture in Image 2, this information is not part of my decision. The decision whether to react and how is made entirely freely, and therefore there are no circumstances in that schematic. The libertarian does not deny that frustration leads to aggression, or that education for tolerance may moderate it, or that the RP pushes the person to press the button at a particular moment in this or that manner. The libertarian only argues that after all these influences there is another stage of decision, which is not determined unequivocally by the influencing circumstances. The debate is not the very existence of influences but the connection between them and the decision itself. In other words, for the libertarian, there is not complete fixing of the decision by the circumstances, but there is certainly influence. I made an identical argument in the previous chapter as well. The libertarian argues that after the brain has made its point, there is also the possibility of imposing a veto. This is human autonomy, the way in which a person is free to decide whether to yield to the pressures and limitations applied to him (i.e., his surrounding mental topography). As we saw in the previous chapter, the Libet experiments demonstrate this quite well, as there is no 100% match between the RP and the decision. Thus, in the libertarian-dualist picture, which recognizes the influence of circumstances on decision-making, as we said, it is required that we add the following prior stages to the decision:

Image 3: Schematic of Decision-making in the Libertarian-Dualist- Picture — Liberty of Will

The difference between the determinist picture (Image 1) and the picture here (Image 3) is the separation between the various stages in the process: Here, in the center of the process there is a stage of decision-making (stage 3), in which the person — the will and the intellect, not the brain — decides. Before this stage there is the processing of data by the brain (stage 2), and after the decision stage there is again a move through the brain (stage 4), which translates the decision made into neural signals which activate the organs. These two functions of the brain exist in the deterministic picture, as well, but there it is all represented in the central circle, because there it is all attributed to the mechanical processing of information, and it is all seen as neural-physical processes which are entirely deterministic, entirely done in and by the brain, and therefore there is no need to conduct such a functional breakdown. Thus, in principle, we could describe the deterministic picture precisely like in Image 3, but there the decision stage would also be nothing more than a mechanical consequence of the preceding stages. But in the libertarian picture, the second arrow from the left (from circle 2 to circle 3) is different, because as opposed to all the other arrows in this schematic — it is not deterministic. Libertarianism argues that circumstances do not determine the decision in a mechanical and necessary manner; they only serve as influencing mental input. The party which freely makes the decision, the intellect or will, takes them into consideration but is not determined by them. Therefore, the decision it makes and transfers onward is not a deterministic product of the circumstances. The two external arrows in this schematic describe a material-physical connection, but the special arrow, like the one following it, express a psychophysical connection — the first one from the material input to the intellect/will and the second from the intellect/will to the material output — and not a merely physical relationship. This is another example of the importance of the full picture of libertarianism (the liberty of will), as presented in chapter two, and the errors which may arise from addressing its deficient versions (freedom of the will). We can now return to the findings of neuroscience described in the present chapter and see if they contain proof against the libertarian-dualist picture.

Conclusion: Two Sorts of Ways Brain States Influence Decisions and Our Character How can the brain influence human decisions and behavior in the libertarian picture? As we saw in Image 3, the brain comes into play in two places during decision-making: Stage 2, where the absorption and processing of the data occurs, and State 4, where the orders for implementing the decision are carried out. Based on this schematic, we can conclude that a change in the state of the brain, such as physical damage, can generally be expressed in two ways. When there is brain damage in Stage 2, the results of the processing, or RP, which comes out from that absorption and processing of data is different. The signal which

enters the decision circle (Stage 3) is different and it is therefore expected that the decision itself will be different at this stage compared to the decision of someone whose brain is functional and healthy. Does this mean that it’s a different person? Certainly not. This is the same person receiving a different input or processing the data in a different manner, and his decision might therefore be different, though it could of course be similar. A blind man doesn’t see an approaching threat and therefore does not respond to it. It is not therefore reasonable to treat a person who lost his eyesight as a different person (i.e., say that he lacks fear). At the same time, a person whose access to other data is barred to him because of brain damage is not a different person. By contrast, when there is harm to Stage 4, the decision of the deciding factor (Stage 3), even if it is identical, will not be realized in the same manner as with a healthy brain. For instance, if that part of the brain responsible for emotions or speech is damaged, the person will not express emotions or will not speak, even if he wishes to. Does that mean that he is another person? Of course not. It’s the same person, but he is not expressed in the same manner at the physical-practical level. The practical result is after all dependent on the messaging system (the brain, Stage 4) and not just the origin (decision, Stage 3). It is very much like a prism, which when we shine a white light on it, the light splits into different hues. 216

Does that mean that the hues are the creation of the prism and that they were not contained within the original beam of light? Obviously not. We do after all know that the combination of all these hues is what creates the light which appears white to us, meaning that all the hues were contained within it. The prism only gives them full expression. What will happen if we create a prism which blocks some of the hues, and which is not transparent to some of them? Obviously, in such a case some of the hues will not pass, and the white light will suddenly come out green or perhaps red. Does that mean that the light which hit the prism is different? Or that there isn’t even a light but only material created in the prism and which only characterizes it itself? Again, no. But in such a prism we will only notice some of the components of the light. Every object bearing color around us is but a kind of prism. A white light hits it (from the sun), and it only reflects back one or more hues. The red color of a table, for instance, is created from the fact that the material with which it is covered only reflects the color red back to us, and all the other components of the white light are swallowed up by it). The same is true of any other colored object. Thus, it is not the prism which creates the light which we see. There is a white light which comes from the sun, and the prism screens some of its hues, and thus allows only part of the light’s contents which reach it to be expressed. It does not allow the other part of the light to have practical expression. In any event, the light itself is not different; it’s just that sometimes not all of it is expressed before our eyes. Another change in the prism will express light in another color, of course. The same prism can produce a number of different kinds of light, except that in all these cases it is transferring the same light. The libertarian argues that at least in some cases, this is the situation when it comes to damaged brains. Even if the person is the same person (i.e., the soul is the same soul), and even if his decisions are similar, there may be a situation in which these decisions will not be expressed, because the part of the brain responsible for their expression (Stage 4) has been damaged. The physical prism screens some of the decisions he made. For instance, we saw that even when the split brain receives some experience, if it is done in the right hemisphere then it will not be verbally expressed, as the part of the brain responsible for speech is on the left side of the brain, and this part did not receive that experience. This means that the experience will not have verbal expression, but it is incorrect to say that the “I” did not experience it. If the soul is the light and the brain is the prism, we can understand that different expressions do not necessarily mean a different message or soul; a change in the brain will change the practical expression of the soul, not the soul itself. In the situation we just described, the problem is that the person cannot express what he wants, meaning that the root of the problem lies in Stage 4, the state of expression and realization. We can already argue that the split brain is in many cases a situation symptomatic of problems of the second kind (Stage 4). By contrast, if the left side of the brain of that person was mortally damaged (and not just the connection between the hemispheres), then that person would not have absorbed part of the information transferred to him at all — for instance, information from the right eye, or emotional and intuitive information — and then we would be dealing with problems at Stage 2. We can now go back and discuss the findings from the study of damaged or split brains, and see why they cannot be seen as sufficient basis for deterministic materialism. In other words, we will now see in further detail how the libertarian views these findings.

Damaged Brain: The Significance of the Findings in the Libertarian Picture The first finding we described here involved damage to a person’s brain, leading to his character and behavior changing. Does that necessarily mean that a person’s character is nothing more than an expression of his brain, its emergent property? Certainly not. This can derive from one of two phenomena we described in the previous section. There might be a problem here of the first type, according to which the owner of the damaged brain did not receive the information in the same manner as a healthy person. If so, it is no wonder that his responses and decisions might be different, as the environment in which he is operating also reacts differently to them (his RP is different). It is clear that when a person understands the circumstances in which he operates differently, he also reacts differently. In addition, problems of the second kind may form here, in which the person wants to express something or behave in a certain way, but he cannot do so because of the damaged part of his brain. For example, think of a situation in which the center of emotions in someone’s brain was damaged. He therefore does not feel what he himself would if his brain was healthy and fully functional. His decisive power — his intellect and will — receives different data than that which would be transferred to him via a healthy brain. When such a person sees another person receiving a blow, he does not interpret it as an incident which causes the other pain, and therefore may decide to give the person who was hit yet another blow. At the same time, a person who does not understand the consequences of his actions will make different decisions. Does that mean that this is a different decider or a person with a different personality? Are we now observing a different, less moral person? Not necessarily. That same person underwent a change, and in our terms it would be more correct to say that the environment in which he operates underwent a change, and that he now makes decisions based on different data, which means his decisions may also be different. When a person is depressed, he makes different decisions than he himself would make if he was in good spirits. When a person is in pain due to a wound or an illness, he also makes different decisions. A person under pressure — in a state of war, say — makes different decisions than someone resting comfortably in his armchair at home. Does that mean we’re talking about a different person in each and every one of these situations? Definitely not. This is the same person operating in a different environment or mental topography, and his decisions may therefore be different than the sort he himself would make if he were in a healthy situation. These phenomena were known long before the latest studies in neuroscience, and no one concluded from any of them that a person cannot decide freely, or that every one of us has a split personality. Thus, even when the brain is damaged, that is not fundamentally different than any other change in a person’s mental or physical state. The arguments taken from these phenomena in favor of deterministic materialism err in presenting an erroneous picture of libertarianism as the counterpoint to their determinism. They see Image 2, in which freely made decisions must be affected by nothing external or internal, as being representative of the libertarian approach, while real libertarianism is actually represented by Image 3. Image 3 leads us directly to the two forms of influence we described above, and thus naturally to the conclusion that the findings from damaged brains have no consequence for our discussion, and certainly cannot be seen as proof against libertarian dualism.

More on the Brain, Emotion, and Morality In the previous chapter, we presented the example of a person whose emotional center of the brain was damaged, and the argument was that he makes decisions coolly and without sentiment. Smadar Reisfeld’s article “On the Connection Between Fear and Violence,” noted there in a footnote, sharpens my argument here. 217 Reisfeld begins her article by describing a woman, whom she names S.M., who is entirely fearless. S.M. does not fear walking in dangerous areas, and attacks on her including sticking a knife to her throat did not scare her at all. She continues to go to these places without any qualms. All this with a sense of vitality and joy of life, which no event, no matter how traumatic, can harm. She doesn’t fear snakes, and she even stretched out her hand to pet a hairy and scary tarantula, even though she was repeatedly warned that the creature is poisonous and dangerous. Horror movies are particularly amusing to her, and even make her very excited. She simply lacks fear. We’re not talking about an abnormal person with incredible psychological strength who can overcome all these scary events, but simply someone who has a different brain than the usual human fare. Because of a rare childhood illness, she does not possess a fear center, which is located in the part of the brain called the amygdala. She therefore simply does not know how to fear. Studies done on soldiers who served in Vietnam revealed that some of them had a damaged amygdala in the opposite direction (i.e., they were too frightened), and they therefore suffered from the reverse: They were post-traumatic, meaning they could not break free of trauma they had experienced in the past, even years after it happened. Other soldiers, with a damaged amygdala similar to that of S.M., did not develop post-trauma in the wake of the war. This damage protected them from post-trauma.

Other studies show that lack of fear can lead to anti-social behavior. The same people, or children, do not know how to identify fear among others. They don’t identify expressions of fear in body language, such as facial expressions of fear. They are not deterred by conduct which makes others feel afraid, as they simply do not recognize this phenomenon. It turns out that these people also do not identify pain or suffering by others, and they therefore cannot feel empathetic towards them. Children with such brains ignore the strong cries of babies, and act with indifference when their mother is injured. They also feel no regret or guilt for bad things which they did. Many researchers have pointed to a person who is unable to feel empathy acting in a manner which is less moral or considerate. Nothing will prevent such a person to harm others, and he will be indifferent to their suffering. When this derives from damage to the brain region responsible for empathy, does that mean that the person is a different one after the damage? Certainly not. This is the same person with the same values exposed to different data, and he therefore reacts to it differently. Like the example of the blind man presented above, we can say that this person is blind to the feelings of others, and it is therefore clear that he cannot take them into consideration. James Blair, a well-known neuroscientist at the United States National Institute of Mental Health, claims that the lack of fear often leads to psychopathic and criminal behavior. In many situations, such people act with extreme violence to achieve their goals. The future punishments such people can expect for their actions do not deter them, because they do not fear them. In 1998, Israel was shocked by the story of Arbel Aloni and Moshe Ben Ivgi, two 14-year old boys, who coolly murdered a taxi driver with repeated gun fire, for no reason at all. This is behavior typical of someone with a damaged amygdala. Blair and his colleagues conducted brain scans of such youths, and indeed found them to have damaged amygdalae. This trait is also attributed to murderers famous for their cruelty (like Charles Manson). Blair sums up and says that there is a direct connection between the low activity of the amygdala and morally deficient conduct, due to the decisive role the amygdala has in the emotional system controlling fear. By contrast, it’s interesting that Dr. Feinstein, who has been studying and following S.M. for years, reports that she shows no signs of anti-social tendencies. She is a happy and good-hearted woman. One day she gave her last ten dollars to a homeless man so he could buy himself a coat for winter. This behavior could be interpreted based on the assumption that her brain damage is different, or based on the assumption that she made different choices with the same neural-brain situation. What does this mean in terms of the moral responsibility which we can impose on such people? In the picture described here, the significance is that such a person lacks part of the data he requires to make the moral choice. A person whose vision is deficient makes decisions differently than someone who can see, as he has either a different, partial picture or none at all. The same is true of someone who lacks the ability to feel empathy for someone else; he suffers from a kind of (mental) blindness, and he therefore makes different decisions regarding others, including moral ones. 218 It’s very much like children who often cruelly abuse a friend, because their ability to feel empathy and arrive at the necessary behavioral conclusions is still developing within them. Someone who doesn’t understand the pain caused by an insult will not hesitate to insult. Is this necessarily a less moral person? Certainly not. It may well be that the values of a person with a damaged amygdala potentially include the prohibition on causing suffering to others; it’s just that he doesn’t know that his action is causing them suffering, or that he doesn’t know what suffering means for the person who is experiencing it. He is simply blind to these aspects. But if we believe that this person does indeed not include the prohibition on causing suffering in his value system, it still might be only because he does not understand what suffering is (his suffering center, or his empathy for others, is damaged). If he understood it, he would likely take care not to cause suffering to anyone. The conclusion is that this is not a different person, and not a less moral person, but rather someone who makes decisions in light of partial, or different data, and he therefore makes different choices. What, then, can we conclude when it comes to the question of whether such a person bears moral responsibility? Such circumstances can certainly serve as grounds for leniency when it comes to the punishment imposed on the criminal acts of such a person. These findings have important legal consequences, but we cannot use them to arrive at the deterministic conclusion that this person is compelled to act the way he acts, or that his brain “calculates” this behavior rather than him choosing it.

A Note on Kant’s Categorical Imperative In chapter three, we described Kant’s categorical imperative and its meaning. We saw there that in Kant’s eyes, morality is not the result of mercy, pangs of conscience, or any other emotion. Moral conduct is not even the result of a teleological examination of the act, in which we look at the expected results of the action. For Kant, moral conduct is conduct done based on respect for the law, or imperative, meaning a cold decision to be committed to the values of morality. Kant derived the categorical imperative from an a priori consideration, which is a purely theoretical one, without reference to facts or emotions. The results of his inquiry where that every thinking person should arrive at the conclusion that he must act in a manner which would be worthy of being made a general law. We can now ask ourselves: What of someone who has no moral sentiment? Can he arrive at Kantian conclusions? The Kantian way of thinking shows us that the answer, in principle, is yes. We are speaking of judgment and not emotion. Therefore, even when a person lacks the ability to use his emotion, there is nothing preventing that person in principle from conducting an a priori inquiry and arriving at such a conclusion himself. It would therefore appear that he should be subject to the moral responsibility deriving from this understanding. At most, we could reduce his sentence if he lacked the emotional assistance which the rest of us have. If he understands the problems with his behavior — say, people explain it to him, even though he himself does not feel it — he must avoid doing it. Again, there is room to be lenient when it comes to the punishment, but not for a principled argument absolving him of responsibility. It is true that a person lacking emotion sometimes will not understand why his action cannot become a general rule. If he does not understand that others suffer from his actions, he may think that there is nothing preventing the whole world from doing such actions. But this is true only when it comes to very severe brain damage and very specific actions, whose damage they cause cannot be understood intellectually. In any event, even in such extreme cases we are speaking of a lack of information and not a moral failure. If such a person acts immorally, this does not mean he is a different person, but only that he lacks data.

Split Brain: The Significance of the Findings in the Libertarian Picture Above, we presented two sorts of problems which could arise among those with damaged brains: Problems with absorbing information (input) and problems in implementation (output). So far, we dealt with problems related to the first sort, these being deficiencies in either the reception or processing of data. But the primary findings which come up in discussions about determinism and libertarianism are the various phenomena of the split brain which include, perhaps primarily, problems of the second sort, these being a gap between the decision and the implementation. In a split brain, the two hemispheres are complete and functional, and input and processing are therefore more or less normal. 219 What is damaged is the connection and the transfer between hemispheres. We thus end up with situations in which a person receives certain input but reports a different input. As we saw, the reason for this is that the input was indeed received, but the person has no way of expressing this, since the transfer between the receiving hemisphere on the right and the hemisphere responsible for expressing this input through speech on the left is damaged. Zeidel himself pointed out in his article that we know very well that a number of opinions, desires, and directions of thought exist within us at the same time. In the end, every one of us makes a decisions which amounts to some calculation of all these arguments and considerations. In chapter two, when we dealt with the weakness of the will, I brought the example of the chocolate dilemma: On the one hand, I want to eat it because it’s delicious. On the other hand, I don’t want to eat it because I will gain weight if I do so??. Is such a situation an internal contradiction? Do we contain two souls with two different values (health and pleasure)? Absolutely not. These are two desires, both of which exist within me at the same time. Even if I decide to eat the chocolate, that does not mean I do not want to lose weight, and even if I decide not to eat the chocolate, it does not mean I don’t like its taste. Bottom line, I must calculate my desires against each other and make one ultimate decision. In any dilemma, valuerelated or otherwise, a number of desires clash with each other at the same time. Each one of them appears reasonable to us, and the dilemma is what to do, and not which one of them is correct. We do not decide whether or not chocolate is fattening or not. We also are not on the fence as to whether chocolate is tasty or whether it’s worthwhile to east something tasty. It’s clear to us that factually, chocolate is both fattening and tasty, and it is also clear that we both wish to lose weight and wish to enjoy a sweet tasting snack. Our dilemma is only which of these two “values” — pleasure or health — is more important on our value scale. Is it more worthwhile to get fat and eat something tasty, or avoid tasty food to stay thin? Let’s assume for the sake of discussion that the left hemisphere represents the cold calculation, the desire to be healthy, while the right hemisphere represents the more emotional desire to eat something tasty. If my brain was split, then the tester who asked me if I want to eat chocolate would receive two different answers from the two hemispheres. Does that mean I have two souls? Not at all. When one part of the brain represents one side and the second part of the brain represents another, this presents no principled problem. Moreover, in chapter twelve, and also here we noted the studies of Josh Greene, who pointed to a number of regions in our brain being active when we are trying to decide a value conflict — including the region responsible for emotion and sociability and the region responsible for decision-making. Does that mean that I am split? Absolutely not. All it means is that every function of my soul is expressed in another part of my brain. What will happen if one of the parts is damaged? In such a situation, the effective decision will only take one of the considerations into account and not both of them, and a different decision will therefore be made. There is no different person here, and not even the same person with a different value system. There are certainly not two people in the same body. What we have here is different data being taken into account for the final decision, or values being weighed differently when making a decision. Even when it comes to the spit brain, there is nothing preventing us from seeing each hemisphere as expressing one aspect of the soul. When the right side opposes Nixon and the left side identifies with him and his policy, this means that Nixon has a number of aspects to him, some of which the person likes and some of which the person likes less. Moreover, the positive sides of Nixon (in the eyes of the subject) are those which belong to the left side of the brain, and the less positive sides of Nixon are those which involve the right side. Therefore when the left side of the brain is asked, it actually supports Nixon, while the right side opposes him. When there is no connection between the hemispheres, this may even be expressed at the polls. A healthy person will calculate all the aspects, and will form one bottom line general position regarding President Nixon. The same is true for any ideological split. Almost any ideology contains positive and negative aspects, and even those who oppose it can find positive sides to it. It is therefore even to be expected that an ideological conflict will arise between the two hemispheres. Does that mean that there are two different souls here? Or that there is an identity between soul and brain? Of course not. We are not dealing with a split personality but rather a split functionality. Each part of the brain deals with a different aspect of the problem, and therefore arrives at different conclusions. The final conclusion of the person is the weighted calculation of both sides together. Thus, it is precisely the person who is healthy in soul and mind who should understand that every issue has sides to it, and he must therefore calculate all of them when coming to form an opinion on it. It is precisely the unified view, which Zeidel seems to consider an expression of a complete and functioning person, which now seems

problematic. Can that same person not understand that Nixon has both positive and negative aspects to him? Those who do not understand this — may be the one who has damaged thinking and maybe a damaged brain. It’s obvious that when there is no connection between the hemispheres, the final decision cannot be done based on consideration of both sides together. If the result is expressed in speech — the left hemisphere will determine matters. If the result is expressed in action — the matter will be decided another way. But none of this is to say that we are dealing with two souls. It simply means that a certain part of my one soul cannot be expressed in the world. Does that mean that the source itself only contains those hues? As we saw in the example of the prism, certainly not. The white light includes it all. It’s just the prism which will not express it all.

A Note on Dissociative Personality Disorder To end the chapter, I would like to briefly comment on damage which is not physical but mental. There are quite a few people — experts estimate it at 1% of the population — who suffer from what is known as multiple personality disorder, or in professional terms “dissociative personality disorder.” This disturbance usually forms due to ongoing assault and abuse during childhood, when the victims adopt various characters they invent to deal with the suffering, helplessness, and other such feelings. To the best of my understanding, those who point to the split brain being an ostensible indication of a materialist identity between brain and soul should see precisely this phenomenon as actually proving the dualist distinction between the two. To be more concrete, I will discuss one famous case in Israel from one of the episodes of Amnon Levi’s TV series, True Faces, which dealt with a young girl named Yuval Shitrit, who contains thirty different such characters. 220 Yuval is a young woman — 23 years old when the episode was being recorded — whose personality would disappear for a few hours or even a whole day, and she would take on an entirely different character. Afterwards, she would return to being Yuval, not remembering what had happened in the interim at all. She would find herself in various situations and not understand how she ended up in them. In those same breaks in time, different characters would form: Sometimes she was a 4-year old boy named Danny. Sometimes she was a 3-year old girl, and sometimes she was a girl with no age. Sometimes she was a blind 13-year old girl named May. This disturbance presents a serious problem for the materialist, and this is to an extent the opposite of the split brain case: A split brain has two brains, while here the brain is one and the personality (or soul) is split. Prima facie, it is difficult to explain these phenomena in a materialist framework. Experts claim that these are usually influenced by abuse, meaning psychological influences, which create psychological transformations. How will the materialist explain it? He will seemingly explain that the brain responds in this or that way to the impression the abuse left on her, functioning in an entirely different manner, as though there were someone else here. The mental events, the experiences, are merely an epiphenomenon of neural events. But now we must move forward two additional stages in our discussion. In the first stage, we must understand that every such character has her own memories, her own way of speaking, her own customs, and more. Danny, for instance, plays with dolls, sucks on a pacifier, and speaks just like a little boy. How will the materialist explain these phenomena? We must take note that all these are subjective phenomena, as these are behaviors which cannot be tested to see if they are “correct.” There is no real boy named Danny, and we can therefore not compare the memories of the character which Yuval has taken on to his memories or her way of speech with his. A materialist explanation for such a collection of phenomena can therefore be found. The materialist will explain that as part of taking on the character, Yuval invents memories and a form of speech. The materialist will explain that these are all the antics of our human brain. We may attribute it to coping with abuse, but this is in effect a physical reaction to physical events. The brain moves from state A to state B, and the experiences which accompany it at the mental level are but epiphenomena. But there is another stage in the discussion, which places our determined materialist in some pretty dire straits. In the case of the character of the blind girl May, it turns out that she has objective skills which Yuval herself never acquired. May plays the piano very well, a capability which Yuval never developed in the real world. How is it possible? Another character which Yuval embodies is fluent in sign language, even though Yuval herself never learnt it. Here, too, we must ask how she knows a language she never studied. These two examples are cases of objective capabilities. In contrast to Danny’s way of speaking, memories, or playing with dolls, which Yuval may very well have invented and are therefore no more than neural events, when it comes to playing the piano, we can compare what she does to proper playing of the piano, like someone who truly learned to play would do, and the comparison shows that she plays it correctly. This is no coincidental state of the brain. This is a skill which is very hard to believe was formed randomly. The same is true when it comes to sign language, which we can also test against deaf people who use it after learning it. Here, too, it is difficult to accept a materialist explanation which would see it as the physical move of the brain from one state to another. Does the brain move by chance to a neural state which contains perfect knowledge of how to play the piano or the use of sign language? The odds that a random change in the brain, without any studying, would lead someone to a state where they would contain the knowledge and skill of piano playing or deciphering language is almost nil. As Fred Hoyle once asked on a similar issue: What are the odds that a random hurricane passing over a junkyard would form a functioning Boeing 747 from the parts? We dealt above with the phenomena of the split brain. We saw that when the brain is split, the person acts as though he contains two souls, and this is unlikely. The split therefore appears to be evidence for materialism and the identity between soul and brain. But here we are talking about an entirely different situation. First, we are speaking here only of a split of the soul, not the brain. Second, the split is not simultaneous, meaning only one soul occupies the body at any given time. Third, in the case of the split brain, there are no different objective skills which were not learned beforehand, only different character (the different parts of the personality which appears on each side). Here, by contrast, we are dealing with objective capabilities which were acquired without prior study. If so, what can explain the fact that in these cases, Yuval can play without learning? The natural explanation is dualist, meaning that there is a component here beyond the brain. We cannot deny that it sounds very strange even in a dualist picture. Does she now have thirty souls? Or perhaps these souls take turns occupying and using her body, so that each time only one soul occupies it? 221It may be that her soul undergoes a change, not that a different soul enters her body each time. I do not know, but it is in any event apparent that it is very difficult to explain or accept these phenomena at the materialist level. It’s very likely that something else is operating there beyond the brain and neural events. In chapter thirteen above, we mentioned the argument of the “god of the gaps.” I explained that lack of information can prove nothing. Here it seems that we are speaking of a “soul of the gaps,” meaning proof for the existence of the soul due to lack of scientific knowledge. But it’s not precisely the same thing. I am not offering an explanation instead of the scientific one. I am only making an initial conclusion based on the fact: The neural dimension does not provide a reasonable explanation for these phenomena, and therefore at least as of now it’s hard to accept the arguments of the materialists that there is no need to assume the existence of additional components. What are these components and what is their nature? I’ve no idea. That is, it is difficult and dangerous to arrive at conclusions from phenomena we don’t know how to explain, but it seems to me that these phenomena nevertheless undermine the materialist thesis which sees all our behavior and feelings as being nothing more than the result of an entirely material neural system. A change in the brain to a state where it knows how to play piano does not appear to be the result of blind laws of physics which coincidentally ended up there. It is more likely that we are dealing with influence at the macro mental-psychological level, which causes the brain (at the micro level) to undergo these changes. This would seem a strong indication of dualism.

Summary From what we saw thus far, the findings presented relating to damaged and split brains are not unequivocal proof in any direction for our discussion. The root of the problem lies in arguments which tie them to the question of materialism and determinism which are based on an incorrect understanding of the dualist-libertarian model. In effect, we are dealing with a failure to distinguish between the (correct) argument that the environment influences decisions to the (incorrect) argument that the environment determines the decision. Alternatively, we are dealing with a confusion between a conception of freedom of the will (decisions without any influence at all) and a conception of liberty of the will (autonomous decisions subject to various influences). The incorrect interpretations of the findings presented in the last two chapters, according to which these findings ostensibly prove determinism, are all based on the same error: The freedom of will argument is but a caricature of libertarianism, which speaks not of complete lack of limitations and influences (freedom) but rather an autonomous action within limitations (liberty).

208 We will not go deep into detail here about the findings and experimental methods, and only discuss them to the extent that they are important for the discussion here. Our goal here is primarily to discuss the phenomena in general and discuss their significance. 209 See the opening footnote of chapter twelve on the case of Dr. Jill Bolte Taylor, which demonstrates the relationship between the function of the two hemispheres and what happens when one of them is damaged. 210 Here we will only point to those which are relevant to our discussion. A succinct description and many sources can be found in the Wikipedia entry “Split Brain.” See also Gazzaniga’s book, primarily chapter three. 211 This distinction has consequences for the issue of indication in analytical philosophy. Bertrand Russell, in his classic article from 1905, distinguished between two types of reference to a concept or object: Description and indication. For instance, we can speak of the first Prime Minister of Israel, and this is reference by describing him through his position. We can also speak of David Ben Gurion, and this is indication, done linguistically through names and not titles. Here we see that each one of these mechanisms of reference takes place within a different hemisphere. 212 We are now back to Ockham’s Razor, which as we saw in chapter eleven is an argument in favor of materialism. 213 Eran Zeidel’s (fairly old; Hebrew) article, “Between His Right and His Left — Two Halves of the Brain,” Mada, 31:6, pp. 1987-1988 214 See more on this in his earlier book: M. S. Gazzaniga & J. E. LeDoux, The Integrated Mind, Plenum press, NY, 1978. 215 Ron Aharoni pointed to this distinction in his (Hebrew) book The Cat That Isn’t There. His main argument is that all philosophy is nothing more than a series of errors, all of which stem from this failure to distinguish, meaning our failure to distinguish between an object and a subject, which is so necessary in statements referring to our subject. The title of the book is based on William James’ immortal saying: “The philosopher is like a blind man searching a dark room for a black cat that isn’t there.” I do not at all agree with this argument, and am even insulted in the name of philosophers and philosophy in general. But Gazzaniga is indeed a “philosopher” in Aharoni’s terminology. He is confusing between these roles of consciousness, and does not at all distinguish between the two roles which consciousness uses. 216 Picture from Wikipedia entry “prism.” 217 Available here (Hebrew): http://www.haaretz.co.il/magazine/the-edge/1.1658338. 218 For the empathy mechanism in the brain, and the mimicking neurons and their role in this context, see Gazzaniga’s book at the end of chapter five. 219 This is a schematic and imprecise picture, as it is clear that there may also be harm to the processing of information in such a situation, data whose reception requires a combination of both hemispheres. But our discussion is solely on the matters of principle, and I am not trying to analyze the findings themselves but only define how to address them. 220 The show was aired on Israeli Channel 10 on 9.1.2012 and can be viewed here (Hebrew): http://panim.nana10.co.il/Article/? ArticleID=857681. 221 From my impression from watching the program, she does not recognize any more than one soul at any given time. The souls only know Yuval and do not know each other, and Yuval doesn’t know any of them.

Part Four Ending: Freedom of Will and Dualism From a Bird’s Eye View After presenting the relevant updated information which contemporary science has on the question of freedom of will, it’s now time to sum up, and that’s what this part of the book will be devoted to. The sixteenth chapter presents a series of thought experiments which are offered to the reader in coming to form their own ultimate opinion on the question of determinism. The seventeenth chapter presents a summary of what we learned from all that we’ve covered in the book. This summary will be done in the form of a legal discussion whose aim is to figure out the question of determinism in light of what we saw — the chapter draws a general outline of the picture, presenting the various arguments in favor and against determinism and libertarianism.

Chapter Sixteen TEST YOURSELF: A SERIES OF THOUGHT EXPERIMENTS AS AN ENCORE Take note of the soul, return there and heal, and its light is like the sun, sevenfold like the light of the morning. – Rabbi Shamyahu Koson, 16th century In previous chapters, we demonstrated that the findings in neuroscience don’t really change the equation regarding either the question of materialism or the question of determinism. We are still in a situation where the question must be resolved in the philosophical sphere. How can we consider things in this sphere? Not with scientific findings but solely via a priori considerations or thought experiments. To prevent obvious misunderstandings, I will start with an important distinction between two types of thought experiments. We saw thought experiments in chapter four whose aim was to examine the coherence of an existing theory. In such experiments, we try to see if the principles of the theory do not lead to a contradiction (i.e., problematic results). By contrast, there are thought experiments which are conducted in a situation in which we cannot actually carry out the experiment itself. In such a situation, I try to think what results I would expect if I conducted it. The goal of such experiments is not to discover the truth but to help me form a position on an issue which is not empirically accessible to me. Such experiments do not amount to proof or even confirmation of any one position, but they do aid people to look at themselves and see where their opinion leans. The person asks himself: If I conducted such an experiment — What result would I expect? This expectation is an indication for some approach or theoretical principle. My expectation for result X clarifies to me that I myself seem to hold worldview or belief A, and by contrast, if I expect result Y — then I seemingly hold worldview B. Such an experiment aids me in examining and sharpening my own beliefs and thus make a decision. In this chapter, I wish to propose a series of thought experiments which the reader can conduct while forming his opinion on these two philosophical questions. The discussion in the book was conducted in three stages: (A) The philosophical-a priori stage. It seems to me that its conclusions tend very much in the libertarian direction. (B) The empirical stage. The conclusions from here do not tend in either direction. (C) The self-diagnosis stage. Since we arrived at the conclusion that the experiments needed to form an opinion on the question of determinism are not empirically accessible (at least for now), what we are left with is examining where we ourselves stand on these issues. This chapter deals with stage C, and I hope it will provide every one of us with tools to try and answer the latter question.

Elitzur’s Series of Experiments Avshalom Elitzur, in his book Time and Consciousness222, demonstrates the distinction between the material dimension and qualia in the following thought experiment in chapter 11 of his book. Let’s assume that you, the reader, receive the following proposal from me: Let me kill you in a painless manner, with the promise that I will build a person which is entirely identical to you biologically — down to the atom and molecule and neuron, and precisely in the same physical state. Afterwards, I will give this biological twin of yours a million dollars. Would you agree? The consistent materialist should agree to this, as his double is identical to him. It has the same thoughts, the same personality, the same memories and experiences, beliefs and values, desires and character; there is nothing in this twin which is different than him, and it is in effect — him. Per his system as a materialist, people are nothing more than a collection of these — memories, beliefs, experiences, &c — and therefore that clone is really him. He has, in effect, not died at all. Meaning, the materialist would see this offer as nothing more than a proposal of a million dollars to him. So what’s wrong? Still, I assume that none of my faithful readers (or my less faithful readers) would agree to conduct this experiment on themselves. Why? Elitzur argues that the decision not to agree assumes a dualist conception, meaning the approach that sees human beings as containing a dimension beyond their biology. This clone may be entirely identical to me biologically, yet it is still clear that it is not me.

If a materialist who insists on consistency nevertheless appears before us and agrees to undergo this procedure, Elizur then proposes the following subsequent experiment: While talking in the lab, and immediately after I made the proposal you agreed to (you are, after all, consistent), I suddenly tell you that the technician in the other room has already created this clone, and he has already gone home, to his house (meaning, your house). Now I merely propose to kill you in a very comfortable way, without pain, and thus seal the deal. At this stage, the consistent materialist is shocked and reacts hysterically, imagining this clone going to his house, living in it and with his family in his stead, sleeping in his bed, wearing his clothes, having his experiences, yet still he himself (who is “he himself”?) is not partner to it. I now return to what I wrote in the beginning of the chapter: These experiments, in and of themselves, prove nothing. This is not a debate meant to force someone to change a position or outlook. A person can say that he would refuse or agree, and still stand firm on his opinion. The goal of these experiments is not to prove anything, but rather to give each person tools to test themselves. If he refuses these proposals, he will learn that he himself is likely not a materialist. Even he agrees that the phenomena of qualia cannot be considered identical with the neural process which engender them. If he agrees — then he is apparently a materialist. Is it possible to make an even more far-reaching conclusion from this: That the person who refuses believed in dualism or vitalism? Prima facie, it would appear that this conclusion is not a necessary result of these experiments, as it may be that these qualia do exist, and as noted they are also not identical to neural-physical phenomena, but it can still be that they emerge from biology itself. They do not occur within any spiritual substance; they are instead traits of matter itself. We must now move one step forward. If neural phenomena truly emerge from biology, then my biological twin has the same qualia as I do (meaning, the same psychological structure and experiences). After all, if our physics and physiology are identical, then whatever emerges from it should be identical (in chapter thirteen we saw that there cannot be two different mental states emerging from the same system based on the same physical state). The conclusion was that this clone is simply you yourself, including in terms of the mental (spiritual?!) dimension. If you refuse such an attractive proposal, then this is a sign that you’re a dualist. If so, it would seem that such a refusal also points to dualism and not just the (trivial) distinction between qualia and neural processes.

The Principle of the Indivisibles: On the Relationship Between Self and Traits Despite it all, there is still some room for arguing that this refusal does not necessarily express support for dualism. To understand this, let’s start with a description of the Leibniz’s principle of the identity of indivisibles. 223Philosopher Gottfried Wilhelm Leibniz, who lived and worked in the eighteenth century, claimed that two entities with precisely the same traits are really the same entity. His argument is that there cannot be two separate objects which possess precisely the same traits. Per his system, it would turn out that if there were two drops of rain which look exactly the same — and for our purposes, we will assume that they are also located in the same location in space — then it is the same drop. Leibniz even proves this argument himself, through process of elimination. Let us assume that there are two entities which are not identical — meaning, they are two objects and not one — but they are not different in any one of their traits. If this is indeed the situation, then object A has a trait which is being not object B, which is not true of object B itself, of course. If so, there is a trait in object B which does not exist in A, and this contradicts the assumption that they are identical in all their traits. We must therefore forgo the assumption that there may be a situation in which there are two objects with the same traits. The problem with such a proof is that it begs the question. After all, the position which opposes this principle argues that there may be two different objects with the same traits. If so, in what sense are they two? What is the difference between them? There would appear to be something beyond its collection of traits. That something determines its Otherness or individuality. According to this system, which can be called essentialism, every object contains essence, or individuality, meaning that it has these traits, or that these traits belong to it. This entity carries these traits, and can therefore not be identified with the collection of traits it possesses. According to the essentialist approach, Otherness is no trait. When I say that A is Other than B (not different in traits, but other), I am not referring to any given trait it has. This is an approach according to which indivisibles (in traits) can be non-identical — two different entities. Now that we’ve defined the opposing approach, we can examine the attack which Leibniz critiques. It’s easy to see that if we assume this position, Leibniz’s attack collapses on its own, as it begs the question. His argument clearly assumes that Otherness is a trait, and this serves as the basis of his critique. But this is precisely the point of dispute. Leibniz in attacking the position is assuming that individuality is not

included in the collection of traits (and therefore Otherness is not a trait), and it assumes in its argument that the totality of traits is him himself. Thus, despite Leibniz’s attack, the position according to which indivisibles are not necessarily identical remains consistent, and to the best of my judgment is even quite reasonable. It is entirely clear to me that the object is not just the collection of its traits. The object is what (or who) bears these traits. By my approach, two drops of water which appear exactly the same and are in the same location are still two drops and not one. It is clear to me that if we weigh both drops, we will end up with twice the weight of each drop separately, for instance. We thus have two different drops here and not two which are one. We can support the essentialist philosophical position with an interesting finding in quantum physics. Modern physics speaks of two kinds of particles. The first are the fermions (named after Italian physicist Enrico Fermi), whose primary characteristic is that two such particles cannot be in the same state — i.e., have the same set of traits — which means that these are particles where every one of them are necessarily in a different state. By contrast, there are also other particles called bosons (named after Indian physicist Bose) — which can all be in the same physical state or possess the same set of traits. The conclusion is that while two fermions really can’t be with the same traits, the situation when it comes to the bosons is different. Thus, this characteristic of the fermions — where two different particles always have different traits — is a physical and not a logical characteristic (Leibniz was not speaking of physics but logic, of course), and is therefore not necessary (as we can see, it does not occur in the case of bosons). Our conclusion from all this is that a particle is not just a collection of traits. There is something, its individuality, which possesses these traits. At the same time, it’s reasonable to assume that people are also not just a collection of their traits. A person possesses these traits or bears them (or is gifted with them), but they are not him himself.

Back to the Thought Experiment and Its Conclusions If we do agree that Leibniz’s principle is not correct, then it is obvious that even if we build a biological structure identical to you, the reader, even the materialist will have to agree that this is not you yourself but at most another creature which is identical to you in terms of traits (his personality, memories, beliefs, &c). It’s true that without consciousness and qualia he will seemingly have nothing different from you in terms of traits and characteristics, but the person who bears these memories and values will still be someone else. This is possible, of course, as this is another entity with precisely the same traits. But it’s still not clear to me why it bothers you that you will die and this clone will live in your stead. Who is this “you” that so disturbs you? Meaning, it’s possible to propose a third subsequent continuation to Elitzur’s thought experiment. Now I offer to kill you gently without creating someone else in your stead. What does it matter to you if there is another copy of you or not? Or maybe I’ll lie to you and tell you that I’ve already created such a double, without really doing so. What’s the problem here? Now we arrive at the question of the value of life and human beings in the materialist picture. Why is there a problem with such a switch? Moreover, why should I even ask you? Maybe I’ll kill you (without pain, of course) without even asking you, and that’s it. Should a collection of molecules agree to be dismantled? Does a chemist in a lab ask matter he’s experimenting on whether it wants to be taken apart? Thus, even if the philosophical questioning is correct, meaning that this double is indeed not you, it’s still not clear to us why a materialist like you would oppose winning a million dollars in such an experiment. Maybe there is no “you” winning a million dollars, and it’s not at all clear why you’re even resisting, and it’s even less clear to me why I should care about your opposition. Why give such value to the life of a collection of molecules? For us, maybe we can say that this is the way we are conditioned, it’s important to us, and that’s enough. It’s a psychological fact. But here we must once again ask ourselves the moral-normative question: What about others? Why should we avoid doing so to them without asking them? To the contrary, to place another in such a dilemma is in itself an improper act, as we place him in a troublesome and problematic situation which includes fears about his fate (unjustified, of course). So let us spare him this torment, and just do it to him without asking. What’s wrong with that? And if the materialist tells me that there is indeed no reason not to do so, but that we are conditioned not to do so (evolution is the deus ex machina of all these problems), what could he tell me if I revealed the secret that I myself am not so conditioned? I myself have no problem doing so, to him or anyone else. Could he condemn me for it? We of course dealt with this question in chapters three, eleven, and fourteen.

Another Experiment, Maybe Not So Thought-Related

We could think of a more troubling experiment: We will take each and every molecule from you and reform them and thus reconstruct you. Will such a person be another human being, or is this duplicate you yourself? Could we treat such an action as murder? If we give your copy a million dollars, would you oppose this experiment? If so, why? What’s missing in the present setup that wasn’t there before? On the question of what he now has that he didn’t before, there’s a good answer: a million dollars. But what does he lack compared to before? There would not seem to be another candidate aside from a soul or spirit, no? Is this just a thought experiment? It may be that we could treat this as a real experiment, if not on people then at least other creatures. Dr. Aubrey De Grey is a gerontologist, a researcher of aging processes, and an autodidact who published a book in 1999 on aging processes and the ways to stop them, earning him a doctorate and a research position at Cambridge University. De Grey is conducting a very controversial experiment, both morally and scientifically, in which he is examining the possibility of defeating death and achieving eternal life. Itay Lahat, in his (Hebrew) article “Man of Eternity,” 224interviewed De Grey and discussed his work. The article begins with describing the life of a medusa the size of a pinky, whose scientific name is turritopsis nutricula, or the hydra. When this medusa suffers from hunger or distress, it can turn back into a polyp — the medusa’s childhood stage — and grow again. It reverses its biological clock, and then reverses it again. Daniel Martinez, a marine biologist from the University of Pomona in California, grows such hydras in his aquarium, which are over a decade old, meaning twenty times older than the average medusa’s life expectancy. In their first four years in the aquarium, they replaced all the cells in their body sixty times. Let us now ask: Is this hydra, which replaced all the cells in her body sixty times already, the same hydra which lived twenty times longer than a normal hydra, or is it in effect twenty different hydras? In what sense is the second hydra, or the seventeenth, the same hydra which was here before? No molecule in its body was present in the original hydra. Does this hydra have something which is beyond its collection of molecules?225 So far, medusas. And what about people? De Grey argues stridently that these processes are also applicable to humans. As noted, De Grey’s project is controversial, at least ethically. What will we do without the redeeming death which grants meaning to our lives? If we do get eternal life, what will make us get up in the morning and do something, and not put it off to tomorrow, or another million years, or a billion? But at the scientific level, it would so far seem that these arguments are not entirely absurd. Dr. Sherwin Nuland of Yale University, a surgeon more than seventy years old who has dealt with medical ethics for many years, wrote a sharp letter in 2005 to MIT’s prestigious journal Technology Review, in which he stated that he has found no weak points in De Grey’s thesis, but he claims that his ambitions are destructive. At the same time, the journal proposed a prize of $20,000 for someone who would write an article which would convince a team of outside judges that De Grey’s argument is absurd and that aging cannot be stopped by hundreds of years with existing technology — i.e., not futurism, but tools we have today! Among the articles sent in response to this challenge, three articles were published by the journal, each of which contained a response from De Grey. The judges were not convinced, and they ultimately declared that no-one was found worthy of the prize. Now let us think what happens if De Grey succeeds and gives everyone eternal life. The cells in our body will all be replaced thousands and millions of times. Moreover, when asked how we could contain all the memories we will accumulate in an endless life, he responded that our nature sees to that. We forget everything which happens to us after forty years, and thus we are always left with room for new memories. Thus, even our memories will not remain as they were. So what will define this future creature as a continuation of us? Will this future creature still be us? What will be in him which will belong to us (present creatures)? The memories and the nature acquired throughout our lives? This can in principle be created artificially, as we saw above. If we are willing to accept that this creature is us ourselves, then why shouldn’t we agree to the tempting proposals of Avshalom Elitzur? Adi Tzemach, a philosopher and literature scholar from Hebrew University, once wrote an article226 in which he proposed an original basis for our ethical obligations. He argues that there is no reason to see myself tomorrow as a continuation of myself today. This is a person with different memories (by a few or by many), in a different mood, and even the cells in my body, which are being replaced all the time, are a little different. If so, why am I so worried about “my” future? And if I nevertheless do so, why should I not do so for someone else, entirely? What, really, is the difference? It may be that the last thought experiment is beginning to look like a practical experiment. In any event, for our purposes, it doesn’t really matter if De Grey is right and whether we could ever effectively live

forever. The principled question can also be asked solely within the framework of a thought experiment: If it happened, would there still be something there which would still be me myself? Is there even any point in making the effort to live forever, when I myself make the effort, and the fruits are enjoyed by someone who is effectively someone else? 227

Buridan’s Donkey There is another aspect I haven’t touched on so far, and which seems to create a serious problem for determinism. Many do not notice that just as determinism cannot explain free choice, it can also not explain a situation in which we make another non-causal decision. If a human being is nothing more than a material collection then all his actions are of a causal-deterministic nature and are controlled by the laws of physics. Such a picture rules out not just freedom of choice but also randomness, at least at a macroscopic scale. To demonstrate the surprising consequences of this aspect, we need to do another thought experiment, in which we will tell the story of one of the most famous donkeys in history, Buridan’s donkey. Jean Buridan was Rector of the University of Paris during the fourteenth century. Among other things, he proposed an explanation for a body continuing to move even after the force applied to it stopped its work. For this purpose, Buridan invented the concept of “impetus,” which expresses a sort of energy embedded in the body by the force being applied to it and which causes it to continue to move. Later, inertia was established as the first law of Newtonian mechanics, and the impetus which Buridan invented could now be translated into a term which had solid empirical and theoretical basis — kinetic energy. But Buridan’s donkey is an even more well-known and influential theoretical concept than his impetus. Buridan used this donkey to discuss, and effectively to undermine, rationalism. Here I will present the argument in a more modern version that he did, and I will use it a little differently. Think of a donkey which is equidistant from two identical troughs. His whole environment is entirely symmetrical in both directions, and the assumption is that his body is just as symmetrical (physicists always say that the problem should first be solved for a point donkey, for simplicity’s sake, and only then go over to a real donkey). Thus, our (point) donkey stands before the two troughs while enjoying chewing grass and with his face facing precisely forward. After the grass runs out, he rests, and upon feeling hungry again he wonders which of the two troughs in front of him he should go to. Because of the symmetry he does not see any possibility of preferring either of them, and he thus stands in place paralyzed by his thoughts and unable to make a decision. At the tragic end of the story, our donkey dies from hunger. His inability to make a decision where to head costs him his life.

A Paraphrase of Buridan: An American caricature from the beginning of the twentieth century, which expresses Congress’ dilemma of whether to dig the canal through Panama or Nicaragua.

The donkey parable was meant to criticize the concepts of rationalism, on the grounds that if the donkey insists on being rational, it will necessarily starve to death. Why is this so? Ostensibly, this is utterly irrational behavior. If he is really a rational point donkey, then he shouldn’t be such an ass! Let him choose one of the troughs as he wishes, and bon appetit. Buridan’s concept of rationalism assumes that every action has a justification or a causal explanation. His assumption is that a rational-causal creature chooses to do action A and not B only if it has good reason to actually do so. But here our donkey has no good reason to prefer either trough over the other, and a rational-causal donkey should indeed starve to death.

Implementation: Buridan’s Human

What would a person do in such a situation? Ostensibly, he is no less a rational creature than the donkey, and will therefore also die of hunger. On the other hand, it is intuitively clear to all of us that the fate of the human being will be different. The person will do some sort of lottery, meaning he will arbitrarily choose one of the troughs, anything to avoid death by starvation. This may be a random act, done without cause in the mechanical sense, but it certainly appears rational to us. To the contrary, rationality itself tells us to act randomly and not causally —and simply flip a coin. This is a different rationality than Buridan was talking about, who was speaking of causal rationalism. But this answer is not so simple. To understand why, let us set aside the question of rationalism Buridan was dealing with, and go back to our own discussion. In the mechanistic-physicalistic picture of Man, every action of his is carried out due to a prior physical cause. Let’s assume that a (point) person standing before two troughs feels hunger. This means that there is a region in our brain coming to life and shouting “I’m hungry!” But the determinist explains that this shout is nothing more than an epiphenomenon, an associated phenomenon. What’s actually happening is a neural awakening, which is sending an instruction to the organs to do some action (= walk to the trough and eat). At the same time and unrelatedly, a feeling of hunger is awakening within us in the mental dimension, but this does not in any way affect the neural mechanisms, and therefore also not our behavior. It only accompanies them in a passive manner. If so, the action we will do in such a situation is solely the result of physical processes. But if the initial situation — both the external one and of the brain — is symmetrical (and we will assume so for the purposes of our thought experiment), then it is not clear how the instruction will in the end be sent to the organs to walk in the direction of one of the troughs. Let’s pose the problem more precisely: For the physical system to start acting, some physical cause is needed. If physicalist determinism sees a person as a physical system, then every action of his requires a physical cause. Such a reason is describable via one of the physical laws, described mathematically via equations and initial and boundary conditions. Yet here, there is a basic principle in mathematics which is very useful and important — the principle of symmetry. Symmetry theorems in mathematics state that in an environment with some symmetry (usually described in an equation and with initial and boundary conditions), the result (= the equation’s solution) must be at least as symmetrical as the circumstances themselves. 228If the equation’s symmetry and the boundary and initial conditions are spherical then the solution must necessary involve spherical symmetry. If the symmetry is one of translational character, meaning that translation to some distance does not change the forces and the variables of the equation and the environment, then the solution will also have such symmetry. In our example, there is symmetry to both sides, in terms of the space (the two troughs) and the structure of the donkey itself. If so, the solution (= the movement of the donkey) must have such symmetry (at least). In the physics of the end of the twentieth century, the Buridan donkey is used to demonstrate what is known as “breaking symmetry.” For instance, let’s think again about the tiny ball from chapter nine at the top of a round mountain. The situation there is symmetrical as well, and we have no way of predicting where it will fall — just as we don’t know which trough the person, or donkey, will eat out of. In the end there will be a minor event of one sort or another, in the environment of the tiny ball (or the person/donkey), which will break the symmetry and thus create a tiny force in one of the directions. For instance, a slight breeze, or a change in the environment’s air density will occur, and this will move our tiny ball from its spot in one of the directions. A real experiment will thus always have something in the environment which will break the symmetry. We did see already in chapter nine that we cannot predict which trough the donkey or person will eat out of, but it’s clear to us that both will end up at one of them. They certainly will not die of hunger. This is precisely why I defined this as a thought experiment, as here the environment is entirely neutral and symmetrical. There is no disturbance of any kind, and therefore the situation remains in place. In such a hypothetical situation, our tiny ball will of course remain on the mountaintop forever, and our donkey will of course die from hunger. And what will we say of a person in such a thought experiment? In a materialistmechanistic picture, such a person will also die of hunger, as there is no physical cause which can make him approach one of the sides. Symmetry forces him not to prefer any of the sides. Anyone who understands human action as being the deterministic result of physical laws must conclude that a person’s actions will bear the symmetry of the consequences in which they operate. What does that mean to the poor guy equidistant between the two troughs? The symmetry of the problem makes both sides equal, and the symmetry of the solution must therefore be the same. There are only two kinds of results with this type of symmetry: (A) Go in both directions at once, which is of course not possible (due to our accursed and stubborn laws of physics). (B) Remain in place and die from hunger. This is mathematical

proof that in the deterministic picture, within the conditions of our thought experiment, a (point) man in a Buridanian situation will die of hunger.

On Donkeys and Men To the best of my estimation, a point donkey in such a hypothetical situation will apparently really die from hunger. The reason for this is that I assume (perhaps mistakenly), as Descartes does, that a donkey only acts causally, and therefore the practical result in his case must be symmetrical, as I proved above. But what about a person in such a situation? Will he also die from hunger? Intuition would seem to say no. Man is a rational creature, but not in the causal sense Buridan was talking about. It’s not reasonable to assume that people always operate based on causes in the simple deterministic sense Buridan was speaking of, but rather in the sense of judgment, meaning acting for goals. The real rationality, then, is not causal-rational but rather the rationality of judgment. That same person can tell himself: “Pal, if you continue to be causal-rational (i.e., determinist), you will die from hunger!” In this framework he could also make an arbitrary decision, against the laws of physics, to turn to one of the troughs to survive and not starve. The decision which direction to take could be carried out by some lottery, or just because. But the very decision to make the turn is a rational decision based on judgment. 229 The significance of such a decision is that the person’s action is not determined by the physical circumstances which existed before it. After all, the circumstances were symmetrical and the decision was asymmetrical. Meaning, there was something beyond the physical which participated in the determination of the physical state in the next moment. But no worries, if and when the Buridanian person is put on trial in determinism court for violating the laws of physics, he could of course defend himself on the grounds of protecting his life. Meaning, if such a person in this hypothetical situation does indeed carry out a lottery and survive, this does not fit with the physicalist assumption. In the physicalist picture, man is a material creature operating only according to the laws of physics. In chapters nine and ten we saw that there are no gaps in our physics, meaning that circumstances always dictate the results. There is no room for neither randomness or free choice in physics. We did see in chapter ten that quantum physics may contain non-deterministic gaps (at least according to presently accepted interpretations), but we also saw there that they are not relevant for the scales at which the brain operates. At these scales, the quantum character is “stretched,” and it is the macroscopic laws of physics which are in control, and these contain no random dimension. Thus, quantum theory will not save us from the expected cruel death which awaits us in the Buridanian situation if the materialist-determinist picture is indeed true. This chapter is about providing a number of diagnostic tools for people to use on themselves. Now every one of us can ask themselves if they consider it reasonable if such a person in such a situation will indeed die. If yes, then he has self-diagnosed as a materialist-determinist. If not — then at least deep inside he would appear to be a libertarian.

Conclusions: Determinism, Randomness, and Freedom — Again The surprising conclusion from all this is that determinism does not just rule out an action based on free will, but also a random action like a lottery. Even though it’s very easy not to notice this, such a mechanism is also not possible in the physicalist-determinist picture. In this picture, if the person made an asymmetrical action, in which he approached one of the two troughs to eat, then there must have necessarily been something in the previous situation which caused this meaning the circumstances were not truly entirely symmetrical, and this is contrary to our assumptions in our thought experiment.230 Rationalism in its deterministic sense will bring a person subject to it and which is in this situation, to starve to death. Simple intuition would seemingly rule this out as unreasonable. It’s important to understand that even if we assume that Buridanian Man would not die in such a situation, there is still no direct indication here of freedom of choice, as randomness could also save him from death. What Buridanian Man gives us is an indication against determinism (or physicalism) and not in favor of libertarianism. Except that if a person has something which is not material and is not subject to the laws of physics, then we have no real reason to reject our intuition and assume that he lacks freedom of will and merely acts randomly. If it’s not true that the whole world is physics, then there is no good cause (!) for sticking to determinism. Moreover, the very judgment of Buridanian Man who can decide to go to one side and eat is in effect a form of libertarianism. He exercises judgment and isn’t simply driven by circumstances, random or deterministic. For those readers who insist on remaining determinists, may I suggest that they not end up, even

coincidentally (is coincidence even possible?), in a symmetrical situation (to the extent that he can even make that choice). Bon appetite, and lots of health for all!

Do Sperm Cells Have Mathematical Capability? I wish to present another consideration, which should have actually appeared in chapter eleven due to its being an a priori argument against physicalism, but to present it here in full, it was important that we get through the chapters in part three regarding emergence and neuroscience, and I therefore arrive at it only here. 231 A headline from the Israeli newspaper Haaretz on 4.4.2012, informed us “Researchers: Sperm Cells Have Mathematical Capabilities.” 232The article describes the phenomenon according to which human sperm cells, like sea urchins, find their way to eggs via mathematical mechanisms. The chemical signals sent off by the eggs attract sperm cells, and they move towards it through a process which is based, among other things, on mathematical derivatives, and this is the conclusion presented in the article: This fact may seem simple to us, but it presents us with an amazing discovery: The sperm cells know how to calculate derivatives (what young people usually only learn in advanced classes in high school) and in very brief time periods, as the responses of sperm cells are done at the rate of a second, and sometimes even less than a second. Hence, the slope of the travel path of the sperm cells is the result of this calculation. For our energetic reporter (who apparently is French, because it is a translation from le Monde), the obvious and “incredible” conclusion, as he puts it, is that sperm cells are seemingly smarter than high school students. But according to that same principle, an electric capacitor or solenoid is at least as smart as the most intelligent sperm cells, as it obtains a derivative and an integral — in a capacitor, the current is a derivative of the voltage, and in a solenoid the result is the opposite — in a far shorter time than the pathetically slow second required for the sperm cells. The electrons which move within the capacitor derive the voltage activated on them, and the result determines the speed at which they move (the electric current. At the same time, light is in effect a mathematical genius, as according to Fermat’s principle, at any given moment it calculates the fastest way to reach its destination, 234 and directs itself towards it (it’s not just a derivative but a differential equation. It’s almost exactly the same activity that sperm cells “carry out”). By the way, it also does so much faster than sperm cells (as we know, light moves at the speed of light, and can therefore also win the Olympic medal for running; it turns out that it’s better than us both mentally and physically). The natural conclusion is that the light’s intelligence is not only greater than that of high school students but also of mathematicians and physicists, who also take far longer to do so. By the way, the light also knows all this from its conception, while they only learn this at the university. And what do they learn about in advanced degree studies in physics and mathematics? Among other things, the behavior of other physical objects. As we said, those physical objects are already at the level of intelligence of graduate students. Moreover, some of the physics is known only to a few scholars in the entire world, and another part of it is not known to this day (as if not, the physicists would be out of work). According to that train of thought, physical objects operating according to these rules make calculations no human being in the world known how to do, and are thus smarter than Einstein, Feynman, Bohr, and all the other physicists in the world (at least as of now). What do we get out of all this? After all, the wonder that all of inanimate nature “speaks math,” operating according to laws which are well defined in the language of mathematics (including the advanced stuff most of us have a lot of difficulty with), has been known for a very long time. What the above article should have said, if there is even room to write it, is that sperm cells are programmed to do an action we describe through mathematical derivatives. All of nature behaves in this way, and we knew this long before someone decided to investigate the amazing journey from sperm to egg. We need to understand that we are not just dealing here with a clickbait article meant to attract eyeballs to articles. We’re already used to that. This is an article lacking content, entirely focused on making a trivial point of no interest to the public. If a person writes a sensational article stating that physics uses mathematical tools, he isn’t getting a Pulitzer. Ostensibly, this is but further demonstration of the common tricks pulled by researchers who try to generate headlines, impressing laymen who read the article or funders and/or potential donors for their research. Those who pay attention to these things can find such headlines almost every day in the popular press, and I would expect science journalists in the press to stop

allowing headline-hungry researchers to continue to yank their chain, and put an end to these absurd and mockery-worthy phenomena. 233 But beyond this common academia pathology, in the context of interest to us here, this trick also raises a slightly more serious question: Is this really just a trick? What is the difference between what sperm cells “know,” or the light, and what a physicist or biologist knows when studying their behavior? What is the effective difference between them and us? After all, the same physicist who carries out the derivative, does so mechanically with his brain cells. He, like sperm cells, receives some function as input and produces its derivative as output and according to the physicalistic picture does it mechanically. What’s the difference between that and our smart sperm? What do we know that it doesn’t? Let’s sharpen the question further: What is the difference between what we know about derivatives and what a numeric computer which does precisely the same action “knows” about it? The process taking place inside the computer is very largely similar to what happens inside our brain, and is certainly its equivalent mathematically. Is it true to say that the computer “knows” how to carry out mathematical derivatives in the same manner that we know how to? And what of our intelligent sperm cells? Were they really gifted with a profound understanding of infinitesimal mathematics, this without the generous help of Newton and Leibniz (who discovered, or created, it simultaneously in the seventeenth century? The deeper question is: What is “understanding” or “knowing”? What do we have that animals and inanimate matter do not? Before we touch on this question, we must define the tools we will use to look for the answer. Obviously understanding is not just the ability to carry out actions and implement them, as these also exist among sperm cells and capacitors. They also carry out actions of derivation, at least de facto. The answer to the question of what understanding is can never be given in terms of the ability to do mechanical actions or in terms of input-output. The ability to do a derivative is not given to every system which receives a function as input and produces its derivative or integral as output. Such capabilities are at most indications that we have an understanding of something, but they are not the understanding itself. The computer does such actions, but we nevertheless have no doubt that it does not “understand” in the sense we commonly use the term. The conclusion is that understanding entirely belongs to the level of qualia, not the neurobiological realm, to the intellect and not the brain. A person who carries out a derivative understands what he does not because he knows how to do it but because this action is accompanied by a mental process which we refer to as understanding. The action is an indication that he understands. This understanding allows him to also implement what he understands in different contexts, but it’s clear that this implementation is but an indication that there is understanding here, not that this is the understanding itself. 234 This distinction is at the heart of the two “rooms” we dealt with in previous chapters: The Chinese Room and Mary’s Room. The example of the Chinese Room tries to sharpen the distinction between understanding and technical inputoutput action, while Mary’s Room distinguishes between intellectual understanding and the unmediated understanding of experience. Let’s frame this distinction a little differently: The understanding deals with infinitesimal mathematics and its ideas, and not any implementation thereof. Any implementation — like that carried out by sperm cells, light operating according to Fermat’s principle, or any other natural process, is a detailed process which stands on its own. In and of itself, it has no connection to other implementations of that mathematical idea of process. The fact that both cases involve the conducting of a mathematical derivation is our own abstraction (and understanding) of what’s going on, and not necessarily something inherent in what we are observing. Thus, understanding refers to the shared aspect of all the different implementations, or the idea which serves as the basis for all of them. By contrast, the implementation deals with a specific act, unrelated to other implementations. The sperm cell does not “see” any connection between what it does and the progress of the light. This connection does not exist at the practical level, only at the level of reason or conception. This connection was only formed in our own understanding and not in the context of the phenomena themselves. This is another aspect of the difference between man and machine, and perhaps any other object (unless we assume that animals have understanding in the same sense, even if this understanding is less sophisticated). But this is precisely the difference between intellect and brain. The brain carries out mechanical actions, just like sperm cells or a computer. The intellect, by contrast, carries out these actions based on understanding. At most, we can say that it uses the brain to do so, just as we use our legs to walk. The search for a scientific definition of understanding is a vain one. We will probably not be able to find such a definition, as it belongs to our subjective level, which always eludes an objective scientific definition.

This means that the understanding and definition of the concept of “understanding,” like the understanding and definition of all the concepts of qualia and our mental concepts, continue to evade us in an ongoing process, and it’s hard to see an end to this annoying vanishing act. It would appear to be an inherent part of our mental sphere. This is similar to our inability to define the color red, the feeling of pleasure, the phenomenon of desire or love. None of these have any objective definition, and can at most have various indicators which characterize them — the pains accompanying love, the calm accompanying understanding, or the stimulation we feel when seeing the color red.

The Turing Test Let’s sharpen matters a bit more, and assume that there’s a sperm cell with consciousness. Does that grant it human-like capabilities? What about birds who can navigate enormous distances and demonstrate skills requiring mathematical capabilities which pilots need very complex measuring tools to carry out? Why shouldn’t we say that these birds have mathematical capabilities? What do they lack for us to be able to say of them that they know and understand mathematics? Let’s go further, and think of a person who plays soccer. He sees a ball rolling and runs towards it to catch up to it. Afterwards, he kicks it to get it to move to another location on the field. All these actions involved carrying out fairly complex mathematical calculations; robots which play soccer are highly complex machines mathematically and mechanically. Should we say that this soccer player is conducting a derivation or other form of mathematical calculation when he kicks the ball or decides to run in its direction? Of course not. He does so instinctively, and there is therefore no need to say that he knows mathematics. Note that here we are already talking about a creature with consciousness and an action which is the result of thinking, decision-making, and implementation. Yet we still do not necessarily attribute mathematical skill to that soccer player. Why not? Because mathematical ability is not measured by performance, even when it is accompanied by consciousness. Mathematical ability exists only for those who are aware of the mathematical action they are conducting, and who conceptualize it before they use it. A sperm cell, even if it had such awareness, is at most like the soccer player. It is certainly not correct to say that he has mathematical capability or understanding. This is simply its nature. That’s how it operates. Mathematical knowledge exists only when the person conducting the action, or building a robot which conducts it, is aware of the mathematics at the basis of it all and uses it explicitly and consciously. This of course naturally leads us to the Turing Test, so popular among artificial intelligence people. Turing, one of the founding fathers of computer sciences, was aware of the problems these machines raise when it comes to the definition of human thinking. He proposed a test which would determine if and when a computer would be entitled to be called human. To test the computer before us, we must place it in closed room A, and at the same time seat a human being in room B. Now we will bring another person who will sit in front of these two locked rooms who will insert questions in some manner to the two rooms, without his knowing who sits where. He will receive printed answers from both rooms, and will continue to conduct a conversation with them. During the conversation, he must determine which of the two rooms contains the computer and which one contains the person. If the questioning person cannot do so, Turing states that the computer he “conversed” with is entitled to be called “human.” Artificial intelligence people treat this test with great seriousness. As far as they are concerned, such a computer is indeed human. 235 This test is very interesting, precisely because of the error at its heart, which identifies understanding with its implementation. For simplicity’s sake, let us assume that we are speaking mathematically with the occupants of the two rooms. Now we introduce a function as input into both rooms, and instruct them to do a derivative of it. The output which leaves the two rooms is the derivative of the function we sent in. So we continue to do with various and sundry functions, and each time we receive a correct answer. Is it correct to say that this computer “knows” mathematics like human beings? Of course not. The computer carried out a technical action, in a manner which is not accompanied by “understanding,” and the person decided based on his judgment what to do and then did it. This computer is like those sperm cells which “know” how to do a derivative. To the same extent, we could have placed an electric capacitor or spool in one of the rooms. Mathematical understanding or knowledge is not summed up in the relations between input and output. As every one of us knows, implementation capability does not necessarily attest to understanding. For the computer to be able to do a derivative, it must be aware of the mathematical action it is carrying out, and it must decide to carry it out. It’s not enough that it carries it out in practice. Here judgment once again enters the picture as a decisive criterion in the definition of understanding. As we have noted quite often, without judgment there is no understanding, a soccer player who moves towards the ball does not necessarily have mathematical knowledge, as he may be aware of his actions, but he is not aware that he is now conducting a

mathematical action. He acts intuitively according to his nature. Only we, from the side, describe what he does as a mathematical action. The view of the human being as a kind of machine, identifying the intellect with the brain, forgoes the most prominent aspect of humanity: our subjectivity, which by its very nature is inaccessible to objective scientific definitions. Again, we are not speaking of subjectivity solely in terms of awareness. The soccer player has awareness, yet he still lacks mathematical understanding and knowledge. The subjective dimension related to understanding is judgment and the conscious choice to carry out actions conceptualized as a mathematical action, not just an action which an outside observer will describe with mathematical tools. The materialist-physicalist should not see anything wrong with the headline and content of the article noted above. As far as he is concerned, sperm cells do indeed “know” how to do derivatives, precisely in the same way that people know how to. We are nothing more than more sophisticated sperm cells. Is it any wonder, given that we come from them?! The physicalist may be able to once again argue that the mental should indeed not be identified with the physical-neural. He is effectively raising the emergence argument: The mental emerges from the neural. According to this proposal, the calculation done in the brain is accompanied by the mental phenomenon of understanding, the definition of an epiphenomenon. This understanding is nothing more than the result emerging from the mechanical calculation. This phenomenon only exists among human beings, and does not seem to exist in computers or sperm cells, and which is what the physicalist picture would appear to call “understanding.” But this distinction appears technical and insignificant for our principled discussion. Understanding as it is expressed among people involves a mechanical action being done based on our decision and judgment, which are mental events. A person who is faced with some problem decides to derive a function and then carry out the derivative, while a sperm cell or capacitor is simply programmed to do so. It’s not just that the sperm cells are not accompanied by any mental event when conducting the derivation (and by the way, who says they don’t?!). The root of the matter here is the mechanical nature of the action by them. The order is important here. Does the brain carry out mathematical actions and understanding is attached to it later, or does understanding and judgment decide to carry out the action and order the brain to carry it out? For the materialist, teaching a person math, including how to derive functions, is exactly the same as programming a computer to do so, or building a sperm cell to operate based on a derivative. But understanding is not in the computer which carries out the action, only by the programmer who introduces this capability into the machine. To the same extent, understanding is not contained in the sperm cell or an electromagnetic wave, as they only carry out mechanical actions. The understanding is only possessed by the researcher of how they operate and who artificially reconstructs that activity. If there was some programmer of the light or the sperm cells to do what they do, then we might be able to attribute understanding and mathematical capability to him. But that takes us to different areas of discussion (meaning theology). This discussion is also in effect a series of thought experiments or diagnostic tools. Every person can ask themselves: Doe he really think that sperm cells have mathematical capabilities? Does the person in the Chinese Room really know Chinese? If his responses are negative, then he would seemingly reveal himself to be a dualist and even a libertarian.

What Does This All Mean? To finish this chapter, we must return for a moment to the hypothetical nature of the thought experiments described here and the significance of this aspect of the experiments which we briefly discussed in the introduction to the chapter. A thought experiment suffers from an inherent problem: A person could always defend himself on the grounds that the results we expect are not the real results. After all, we did not truly measure the results of an experiment we are unable to conduct in the real world. We can always say: The sperm cell does know mathematics. Who says they don’t? Alternatively, I could say that I will agree to a biological switcheroo since it can’t actually happen in practice, or: Yes, the person in the Buridanian experiment will indeed die from hunger. Who could carry out such an experiment, especially without the approval of the Helsinki Committee for Human Rights? So what to do? This is where our intuition comes in. As I explained, a thought experiment is meant to teach us something about ourselves, not necessarily about the world. We examine our intuitions by standing in front of a mirror. When a person claims he’s a determinist, let him test himself to see if that argument makes it through the Buridanian test. Does he really think that he himself would die in a

Buridanian situation? If so, then he really is a determinist, and the experiment taught him nothing new. But if he’s honest with himself and admits that he is convinced that the person will not die, then he must reexamine his deterministic worldview. A thought experiment will never prove that determinism is not true; it will at most prove that you, the person who does not believe that the Buridanian person will die, do not believe in it and are thus not a determinist. By contrast, those who say that our point person will indeed die of starvation, or those who really should accept the tempting offers made by Elitzur — and if he doesn’t, that’s just a question of instinct or habit — then he can indeed remain a determinist. This is the deficiency of the thought experiment compared to the empirical one, in which results are measured and are therefore usually accepted as real. The “results” of the thought experiment are unknown, and we can only speculate on them. Thus, this series of experiments helps every one of us to demonstrate to ourselves where we stand on these questions. There is no proof here that materialism is not true. These are at most proofs meant for you, the reader, and they are meant to help you form a position on these issues for lack of a way to do so via scientific-empirical means. Still, it seems to me that there is something here which is perhaps stronger than all the proofs and philosophical and scientific arguments. These deal with the question of what is correct, while this series of experiments deals with the question of what we think and what we really believe in. As we saw more than once, we do not in the end have unmediated access to the truth itself; at most we have access to what we think about it. 236These thought experiments teach us precisely that: What we think about it all.

222 Published by the Israeli Broadcast University and Defense Ministry, 1994. 223 On this principle and its meaning, see Two Carts and a Hot Air Balloon, second edition. 224 Available online here (Hebrew): http://www.calcalist.co.il/local/articles/0,7340,L-3417710,00.html. 225 Philosophy describes this problem via Theseus’ Ship; see the relevant Wikipedia entry. 226 Published in the book On the Just and the Unjust (Hebrew), Tel Aviv, 1977. 227 I assume that the skilled reader noticed that this series of experiments tests materialism and not necessarily determinism, at least not directly. But as we’ve seen more than once, these ideas are interdependent, even if not necessarily in the strictly logical sense. We could have more directly involved determinism here and asked what would be the meaning of the decision in relation to these proposals. Are you choosing it? Is your refusal, if you are refusing, based on instinct or principle decision? And if it is an instinct, why should I consider it? And my consideration, if I even have control over it, which may also be forced on me? 228 Mathematician Emmy Noether published the famous theorems named after her in 1918, which state that we derive from any such symmetry the existence of a preserved quantity, thus explaining all the laws of preservation in physics (like the preservation of momentum and energy). All these theorems are the result of the phenomenon noted here. For more on the subject of symmetry, see Marcus de Sautoy, Symmetry: A Journey Into the Patterns of Nature, Harper 2009. 229 Another philosophical note, which I can’t help but present here. This decision itself is hard to explain in a physicalist picture, as we are after all thinking here of an idea which does not sit with the laws of physics, We conceive of a creature in a symmetrical environment which acts in an asymmetrical manner. Where did we derive such a possibility? How did this asymmetrical thought even emerge in our physicalist brain, which is built on physicalist principles and which has no source for making such assumptions? This is itself an argument, even if not decisive, in favor of libertarian dualism, or at least against physicalist determinism. This argument is similar to Descartes’ anthropological argument, which states that we cannot think about concepts which do not exist by us in an essential manner (meaning that are not a trivial extrapolation of our experience) and which were also not drawn from anywhere in the external world. If both we and the world are built in a purely physicalist manner, then it is not clear how we can have a frame of reference which does not align with all this. This argument can only be rejected with the statement that human beings can think wrongly. For instance, until the modern era human beings thought in Aristotelian terms, meaning that they understood nature in an erroneous manner (which does not fit how nature actually operates). 230 As noted, this is a thought experiment. In the real world, such a perfectly symmetrical situation cannot exist. Already the Jewish Sages stated (in the Talmud, Bechorot 18) that it cannot be subject to tzimtzum or precision, not by man and not by heaven. 231 This discussion started while preparing for the Jewish holiday of Passover, 2012. My thanks to my son Nachman, the conversation with whom produced the section you are now reading. 232 See here online (Hebrew): http://www.haaretz.co.il/news/science/1.1679607. 233 I assume that any researcher in any field, certainly in the natural sciences but not only them, encounters dozens of examples like this, and may have even partaken in some of them. Shimon Marom, in the article we referenced at the beginning of the book, pointed to the sensational headlines on neuroscience being based on these motivations. See his words noted in the introduction to this book, but even more so the following honest admission: Just how justified is the optimism of neuroscience in light of the task it has explicitly taken upon itself, to understand the origin of spirit from matter? What is the true status of the neurosciences? Not the one meant for the ears of interest-driven donors, which we (yes, including the present author) direct to potentially thirsty science journalists, government officials who are in charge of research budgets, or colorful and attractive formulations? 234 This is the accepted definition for the purpose of exercises in university. The goal of the lecturer in the academic course is to pass on the body of knowledge, and thus to create knowledge and understanding, while the exerciser’s goal is to teach the student to implement this knowledge and understanding in various contexts. The exercises aid in understanding, but the ability to implement our insights is not in itself their understanding. 235 One of the most prominent and popular writers in this field is Ray Kurzweill. Kurzweill is a man who deals with artificial intelligence and is a very resourceful inventor, primarily in the field of aids for disabled people. He received the National Medal of Technology and Innovation in 1999, and three years later merited entering the Inventors Hall of Fame in the United States. His most well-known book is The Age of Spiritual Machines. In the book he deals primarily in futurism, predicting technology which will expand human intelligence and will in effect bypass it. His futuristic calculations have come in for quite a bit of criticism; I will only say here that his arguments are based on far too reductionist assumptions on the identity between artificial intelligence and people. Kurzweill entirely ignores the subjective dimension of “understanding.” 236 And once again we see Descartes’ cogito argument, which we saw already in the Prologue.

Chapter Seventeen SUMMARY A philosopher is like a blind man searching a dark room for a black cat that isn’t there. – William James We have no region in the brain which is responsible for moral responsibility. Moral responsibility is not in the brain. – Michael Gazzaniga We can sum up the physicalist-determinist argument in the following schematic manner: 237 1. 2. 3. 4. 5. 6.

The brain is a necessary and sufficient condition for the mental. The brain is an entirely physical object. The physical world is fixed, and contains no gaps. The brain is therefore also fixed, with no gaps [based on (2) and (3)] The mental is also entirely fixed [based on (1) and (4)] The experience of free will is an illusion. [based on (5)]

Gazzaniga points to arguments (1) and (2) being a matter of complete consensus among researchers. Argument (3) is under serious attack from the direction of quantum theory, chaos theory, and emergent systems. He therefore concludes that argument (4) is also not obligatory, and therefore (5) and (6) aren’t either. Bottom line, he concludes that there is free will, but to my opinion he does so for the wrong reasons and based on incorrect definitions of free will. What do we see from our book? As we showed, assumption (2) is of course correct: The brain is a physical object. We also saw that Gazzaniga’s argument regarding assumption (3) is not relevant for the brain: There are no gaps in chaos and it is therefore irrelevant for our discussion. Quantum theory is also not relevant for our discussion, both due to the issue of scale and because randomness is not the same thing as liberty. Emergence is not relevant because all the known examples of emergence only demonstrate weak emergence. To explain freedom of will based on the neural and material totality, we need strong emergence, and such an argument will forever remain a speculation without foundation, since its own definition prevents it from having an empirical basis. Thus, ostensibly arguments (4) and onward are necessarily true. Still, the libertarian of course argues that at least arguments (5) and (6) are incorrect. Why? As we saw, dualism argues that free will can create force fields which move electrons. This insight undermines argument (4), even without the theories of quantum physics, chaos, and emergence. The brain is not entirely fixed even though it is a physical entity, as will also affects it. There was room to raise doubt for the same reason regarding assumption (1), as if will can move electrons then perhaps it can also operate my hand without the brain. But in chapter two, we saw that it’s more reasonable to assume that will does not move electrons but creates force fields. Based on a similar argument, it would appear more reasonable that will creates fields which move electrons only in the brain, and that all the subsequent results — activation of organs, feelings, &c — occur through the brain. I therefore hypothesize, at least for the purpose of this discussion, that argument (1) is correct, even if not entirely necessary. So, where do we stand now? Arguments (1)-(3) are correct. In spite of this, as we said, argument (4) is incorrect. It is thus much the same when it comes to arguments (5) and (6). What is at the root of the debate? Physicalism assumes materialism, and it therefore considers it clear that if any object is physical then it is fixed. But libertarianism is based on interactionist dualism, meaning that it assumes that spirit can influence material and the mental can influence the neural, and it therefore opposes deriving (4) from (1)-(3).

What we have beyond this are the findings of neuroscience, which as we saw from chapters fourteen and fifteen do not prove anything beyond this. At most, they prove argument (1) that the brain fixes the mental. But what is lacking there is proof that the brain itself is fixed deterministically, something which the libertarian obviously doesn’t accept.

Court! So far, we saw that contrary to the consensus among neuroscientists, this field has not really changed the framework of discussion when it comes to the problem of free will. We saw that the libertarian position is possible, and that physics and neuroscience do not force us to abandon it. The field is thus open, as both positions are consistent and in accord with the scientific findings. What we now need to do is return to the philosophical discussion, the one conducted prior to the emergence of neuroscience, and see which of these possible positions is more reasonable. We did this primarily in parts one and two of the book. Let us remind that in the previous chapter we offered another way to look at the issue in the form of selfdiagnosis. Later in the chapter we will conduct a mini-trial between the libertarian and the determinist, in which I will present the philosophical arguments we encountered. 238

A Preliminary Discussion As we saw, the roots of the dilemma between libertarianism and determinism is the clash between the two very basic intuitions we have: The intuition of freedom of will and moral responsibility on the one hand, and the physicalist-causal intuition on the other. Do we have what to say on this aside from the very placing of these positions opposite each other? Which of them is in a better “legal” position? In chapter six, we saw that a reasonable doctrine when it comes to such a clash is preferring the more specific intuition, or Lex specialis. What conclusion do we derive from this doctrine? A wholesale adoption of the principle of causality will lead us to entirely forgo the intuition of freedom of will, while the adoption of free will only requires that we place a slight restriction on the principle of causality: In the case of the free activity of people, and only there, physical events may occur without a cause. Moreover, in chapter six we saw that the principle of causality can indeed continue to be accepted wholesale in the physical world, meaning that we can continue to hold that every physical event has a cause. The concession we must make in order to remain libertarians is twofold: (A) There are certain mental events (free wills) which occur without cause. (B) The cause for physical events always exists, but it is not necessarily a physical cause. Will can also be a purpose, and effectively a cause, for a physical event.

On Evidence and Rules of Decision Thus, Lex Specialis grants clear preference to the intuition of freedom of will. This is the starting point of the discussion, when we established the initial balance of forces. Yet still, if there is proof in favor of one of the sides, we would have to rule in its favor. It’s important to understand that rules and doctrines of decision have two purposes: (A) They determine which side bears the burden of proof. (B) Lacking proof, they also determine who wins the trial. In any event, the rules of decision are only relevant when decisive proof is lacking. This is like a case where Reuven possesses an object of some kind and Shimon comes and sues him in court. Shimon claims that Reuven took the object from him without permission. Shimon is effectively asking the court to take the object from Reuven and give it back to him. The initial situation in this legal dispute is that Reuven has the upper hand, as he is in possession of the object. This is the preliminary discussion, but certainly not the end of the legal deliberation. This analysis only shows that Shimon is the one who bears the burden of proof. 239If the initial analysis shows that the initial burden of proof lies with Shimon, then so far as he has not met that burden — Reuven will win at trial. But when there is proof, then this preliminary discussion loses its force; if Shimon brings proof for his claim, he will of course win the object, even though Reuven is in possession of it. Thus, even regarding our subject, after the preliminary discussion we must now move on to the evidence phase. But first I will conduct another discussion, at the level between proof and a priori intuitions.

The Status of the Series of “Illusions” Throughout the book, we saw that determinism requires that we declare quite a few of our assumptions and insights to be illusions. Beyond the intuition of freedom of will and choice as well as moral responsibility, we saw a fairly impressive list of additional “illusions”: The illusion of the value failure or “weakness of the will” (discussed at the end of chapters one and two), the illusion of conflict and judgment

which a person feels when making a choice (see the second part of chapter four and chapter fourteen), the illusion that we have two paths of actions before us, and that we can determine which one is taken, the illusion that when someone chooses, his actions are not caused by any external reason (see the end of chapter two), the illusion of moral condemnation and rage towards horrific acts (condemnation in the ethical sense, not the psychological one), the illusion of the guilt felt by the criminal, the illusion of the existence of moral norms, and of course the illusion of commitment thereto (see chapter three), the illusion that one mental state leads to another mental state, the illusion that there is a difference between norm and value statements of the “ought” variety and factual statements of the “is” variety, meaning the illusion that the naturalist fallacy is indeed a fallacy (see the end of chapter fourteen), the illusion that it is worth it to work to get results (the lazy man argument), the illusion (of whom?) that we have consciousness, and specifically a unified consciousness, and finally, we must add our argument in chapter four (the fool’s cycle) — which adds a whole collection of illusions which we call scientific, ethical, or philosophical conclusions. As we saw there, the determinist would consider all our conclusions and beliefs to seemingly be illusions (and as I noted there, it is not clear why this is not also applied to the determinist-materialist conclusion itself). Is all this in and of itself not a series of arguments against determinism? To establish this, we must discuss the status of these “illusions.” Ostensibly, all these insights are our own a priori assumptions, and we can therefore see the argument from illusions as only an addition to the preliminary discussion: In addition to the intuition of free will, we can add a long list of other intuitions we all have which support libertarianism. Even determinists feel what the libertarian feels, and they especially feel that they are free to act as they wish. Gazzaniga describes this “illusion” as follows: 240”The simple truth is that even the loudest determinists and fatalists cannot deny the fact that at the personal psychological level, they do not really believe that they are pawns on the chess board of their brains.” Determinism forces us to give up these intuitions, and this is ostensibly another argument against it. According to this approach, all these are not proofs but rather additional a priori motivations. On the other hand, almost all of our thinking is based on such a prior assumptions. Science is based on the principles of induction and causality, Ockham’s Razor which states that we should choose the simplest theory, and more. All these have no empirical support, and they are but assumptions made by our reason. The distinction between scientific proofs and a priori insights is therefore not so simple. In other words, we cannot abandon the weight of our a priori assumptions and only see them as starting points, as if not then even empirical and scientific evidence will not be acceptable. Thus, it seems that there is definitely evidentiary weight to this series of “illusions,” and this for a few reasons: First, since our intuitions prove themselves to be fairly reliable, including in the field of science. Second, because anyone who claims that some human insight is an illusion must also explain how it was created and why — for instance, what is its evolutionary value. To reverse the argument (see more in chapter four), this is in effect Descartes’ anthropological argument: If an insight is formed in our minds, there must be a reason that it formed. The formation of an illusion without a basis is an unreasonable scenario, unless there is a convincing explanation as to how it was formed. For instance, if an insight formed within me that I have free will — it’s hard to explain this without assuming that I am actually watching or experiencing this phenomenon. After all, if this concept did not exist in the world around me, then I have no source from which to take it. This is true sevenfold when the concept exists among all people; even those who deny it understand its significance — they simply deny that the claim is tangible. In a certain sense this is but the implementation of the principle of causality. Third, every argument is of course based on certain assumptions, and if we conduct a discussion without assumptions, since every assumption is suspected to be an illusion, then how will we present out arguments? Fourth, we saw the argument of the deceptive demon in chapter four which attacks determinism. We presented the determinist with the following challenge: If he is indeed claiming that a long and impressive list of human insights and intuitions are nothing more than an illusion, then this itself is proof that the demon not only exists but is indeed deceiving us in many areas. Finally, in chapter fifteen, we even identified this wicked demon with the help of Gazzaniga: The left hemisphere of our brain. So what is the basis for the trust we place in this demon (i.e., our system of thinking and consciousness) in other fields? Why do we trust those hemispheres which provide us with these causal and deterministic intuitions, or the laws of science, or determinism in general? They have, after all, been proven to be deceptive. We saw there that this deterministic argument is not the same thing as simple skepticism, as it shows that this approach lacks any foundation and is almost entirely unlikely. Thus, the simple assumption of the entire discussion is that our a priori assumptions are correct so long as they have not been proven to be otherwise, at least if it comes to insights shared by most if not all

human beings. And it is indeed just unreasonable in general to expect people to give up on their assumptions without a reason. Hence the starting point of the discussion: Bottom line, the libertarian approach is closer to our intuitions, and it is therefore preferable to determinism. The burden of proof therefore falls on the determinist. If he wants to convince us (and himself) that almost everything I feel is an illusion, he must bring weighty arguments to support his case.

The Evidence Stage We must examine two types of proof now: Philosophical and scientific. We saw in the last few chapters that the scientific evidence doesn’t really take us in either direction. There might be an exception, and it is actually in favor of free will: The indication of libertarianism which we brought from dissociative disturbance (the case of Yuval Shitrit). At the philosophical level, we pointed to some proofs in favor of libertarianism in chapter four: The determinist cannot morally condemn or judge. Even if he has no basis for doubt, he cannot believe his own scientific and philosophical conclusions. We saw that they are not hopeless in the deterministic picture, but they certainly lack foundation. The fool’s cycle argument summarily rejects the determinist conclusion, regardless of any future argument findings presented in its favor. Even when deterministic materialism is proven, we will have to ask ourselves what justification we have for believing the proof, since this is a conclusion which is forced on us. In chapter four, we also saw the arguments of the deceptive demon which we also noted above, as well as Newcomb’s Paradox, which also proves that if there is a preferred strategy in this experiment then there is necessarily free will, as the determinist will have to conclude that freedom of will contains a logical contradiction and not just that it does not exist in reality, a conclusion which is unreasonable and which contradicts what we showed in chapter seven, when we presented a mathematical argument which presents a consistent and contradiction-free definition of this concept. Most of these proofs, aside from the various forms of the fool’s cycle, can ostensibly be rejected on the grounds that they are a priori insights and are classifiable as illusions. If so, the philosophical proofs can join the previous section, as a person can say that these are intuitions and not empirical findings. In addition, it’s hard to imagine scientific findings in favor of freedom of will. However, as we already saw, we cannot and should not dismiss the value of such a priori proofs. In addition, scientific findings are also based on some form of interpretation and on principles assumed by science — indeed, we saw in chapters five and six that even the principle of causality is an a priori assumption and not an empirical finding — and all these are but a collection of a priori and philosophical considerations. No-one claims that he saw with his own eyes that people do or don’t have free will. At most, this is an interpretation of scientific findings, and as such belongs to the philosophical-a priori field, and not empiricism in its rigid and direct sense. Thus, even empirical evidence can be rejected in the same manner, and this discussion, like any other, is doomed to be found irrelevant. We once again must face the deceptive demon argument: If all these are indeed illusions, what remains acceptable in the eyes of the determinist? It’s hard to think of a clearer and more obvious intuition than the freedom we possess. If that’s an illusion, then it does not appear that there is anything which is immune to this attack. Another argument against physicalism is based on evolution. As Plantinga shows, it is hard to see a reason for evolution forming correct judgments within us. At most, it should create correct behaviors which benefit our survival. If so, what is the basis for the trust which we have in our own judgment system which was formed in an evolutionary process? Moreover, why does evolution even produce judgments, correct or not? After all, if our mental dimension is but an epiphenomenon, then our behavior is dictated only by our neural component; it’s just that this is accompanied by various mental events which do not in turn affect the neural. So why to even form such a mental dimension? It would have been sufficient to create a neural system, which ensures the best levels of survival. Thus, we have no apparent reason to deviate from what is clear to us all. First, because the determinist does not meet the burden of proof. Second, because there are very strong arguments in favor of libertarianism at least on the philosophical level. The questions of determinism and materialism were and remained philosophical questions, and the sort where the libertarian-dualist conclusions seem natural and logical. What will happen in the future? It’s hard to predict. We analyzed and presented the various possibilities of conducting a Libet experiment which would be more significant for our philosophical discussion, but as we saw at the moment, it’s hard to even draw the outlines of such an experiment. What could even be the results which would force us to forgo our insights and intuitions? And even if we arrive at opposite conclusions, would we even be able to believe them? The fool’s cycle would rise again.

Forming a Position in Light of the Thought Experiments Bottom line, after all the philosophical arguments and interpretation of scientific findings, we are also left with the series of thought experiments. With them, you the reader can decide if one of the sides convinces you. As we saw in the previous chapter, these are not arguments in favor of one side or the other, but a sort of laboratory for self-examination. If you wish to know where you yourself stand after all the arguments, conduct these thought experiments on yourself, and then you will know the answer.

Physicalism’s Munchausen Syndrome Gazzaniga, who presents an entirely physicalist position in his book, tries, like some of his colleagues, to nevertheless defend the idea of human moral and criminal responsibility. He writes (see the opening quote for this chapter) that there is no region within the brain for moral responsibility, concluding that the responsibility is not in this organ, but in the relationship of people with their environment and society. Already in the first chapter, we noted the phenomenon of the plasticity of the brain, and how the argument which sees it as a solution to the question of determinism and moral responsibility is absurd. If human behavior is entirely determined by other factors, be they external or internal, then one still has no responsibility for their actions. Gazzaniga commits the same fallacy Ciechanover did in chapter one. Relying on our having no moral responsibility region in the brain obviously begs the question, and this tells us almost the whole story. There is indeed no such region in the brain, as moral responsibility comes from the spirit which affects the brain. The fact that it is not the brain but the soul which makes our decisions, meaning that the person makes his decisions freely, is the only possible basis for imposing moral responsibility on a person’s choices.241 It’s amazing to discover how such prominent and talented neuroscientists fall for such crude fallacies in their attempt to protect their positions and examine their significance. This is very reminiscent of James’ statement brought at the beginning of the chapter, on the blind man searching a dark room for a black cat which isn’t there. The materialist neuroscientist is the blind man in this story, as he does not have the tools to discover what he’s looking for, using physical measuring instruments to search for a non-physical entity. The cat is free will. For the neuroscientist, it is black, as it cannot be seen when observing the brain and certainly not with physical measuring instruments. It is therefore no wonder that his conclusion is that the cat in question isn’t there. But can we conclude based on this that the cat really isn’t present? Moral responsibility cannot be in the dark room, this being our physical components and the physical world in general in our case. The attempts to find it there have all failed. Those who make such attempts fail to understand that the cure for what ails them is not in physics, which contains no gaps, nor in sociology, as there is no factor external to people. The source of our moral responsibility, if it exists, is present only in ourselves. The “I” thinks, ergo it exists. Gazzaniga’s arguments are just one of many examples of the “heroic struggle which the physicalist fights with himself. This struggle is always accompanied by a deep and profound denial of the serious consequences which naturally emerge from his position, as though the cure for the ails which derive from him can be found in physicalism. We can compare this struggle to two of the famous stories of the beloved Baron Munchausen. The Baron tells that when he fell in a deep hole, he was not paralyzed by despair but rather immediately took his fate and initiative in his own hands. He ran home, brought a ladder, and used it to climb out of the hole and come home happy and content. The same is the case with our obstinate determinist. He is in the physicalist hole, imprisoned by the laws of physics and causality, and is therefore subject to the poisonous influence of the deceptive demon — the left hemisphere of his own brain. But he, too, does not despair. He immediately brings a ladder from the outside — insights which can be believed, arrived at through conceptual tools and contents not subject to the demon’s influence — and thus comes out of the hole and recognizes reality for what it is, meaning he understands that he can believe nothing. That’s as far as physicalist materialism is concerned. In the second story, the legendary baron rides in on his knightly horse, and begins to sink in a swamp which threatens to drown him. Here, too, he maintains his calm, and immediately pulls himself out by the hair on his head and escapes certain drowning. Our physicalist also pulls the hair on his (bald) head, meaning he uses freedom (or judgment) which he does not have to save himself from the freedom which threatens him. It sounds amusing in the literary context. In the scientific and philosophical contexts, it’s more pitiable than entertaining.

Summary In the introduction, I defined the two goals of the book as follows: (1) Demonstrate that the present findings of neuroscience do not change the status of the two questions we dealt with — the question of materialism and the question of determinism. Then as now, these two questions lie in the philosophical sphere and not the scientific one. (2) The more reasonable philosophical conclusion is that we have freedom of will and choice. The reader who has made the entire trek is now called upon to examine his position regarding the two questions we dealt with. It seems to me that even those who do not agree with the intuitions which lead me to conclusion 2, still need to agree with me at least as regards conclusion 1. In the previous chapter we also saw that there is a third, diagnostic level: If we do not insist on denial, we will see easily that we ourselves usually hold to a libertarian position. The libertarian intuition accompanies every one of us, including determinists. It seems to me that we saw that the determinist doesn’t really succeed in pointing to any reason to abandon it. This experience is the most fundamental ways in which we see ourselves, and it is stronger than any other a priori assumption which can ostensibly be placed against it. As we saw, the basis for the strongest arguments against libertarianism is the need to base freedom of choice on randomness or causality — as we saw in the case of Van Inwagen’s argument, which repeats itself throughout almost every text dealing with this issue. But we also saw that these two mechanisms are far less apparent to us and understandable to us than our free will. The deterministic arguments which reject our most basic intuition just because they cannot be based on one of two mechanisms which are entirely unfamiliar to us from observation, seems prima facie utterly unreasonable. The common feeling is that the libertarian is ostensibly supposed to apologize for a position which is not rational or for adherence to making prior assumptions — is the opposite of reality. It is the determinist who (also) suffers from committing these two fallacies, in addition to many other difficulties. So if science has nothing to say in the matter, and if we were all gifted with libertarian intuitions, and if the philosophical considerations also lead to this conclusion, it would seem to me that the natural conclusion is clear as day. As they say: What’s green on the outside and red on the inside and has watermelon seeds, is a watermelon… 237 This structure is based on Gazzaniga’s summary on p. 129 of his book, but with a slightly different formulation. 238 I will not repeat these arguments here, as I have done so many times throughout this book. 239 The halachic term is “He who wishes to remove from his friend, the [burden of] proof is on him.” 240 P. 105. 241 In chapter three, we saw that in the deterministic picture, there can be justification for imposing punishments, but not for condemnation and moral judgment. In this picture, punishment is not an expression of moral responsibility but a mechanical “corrective means” for the person’s brain.

Epilogue COGITO ERGO ELIGO — “I THINK, THEREFORE I CHOOSE” The Cogito of the Libertarian In the prologue, we dealt with the question of materialism and dualism via Descartes’ cogito principle. We saw there that it is actually spirit whose existence is more certain than body and objective material reality. Qualia precedes the physical processes which engender it and is more certain than them. It is what recognizes material and the physical laws which control it. Does this have consequences for the question of determinism-libertarianism? It would certainly appear that this is the case. In the second part of chapter four, we saw that the conclusion of the determinist himself regarding the character of the world is based on his judgment and decisions. Therefore, if this is forced on him, meaning that it is not based on judgment and free choice, then there was no real basis for believing this conclusion and for seeing it as a real claim about the world. If so, the possibility of espousing a deterministic position is made possible only in a world in which there is judgment and freedom to decide. Therefore, the situation regarding the question of determinism appears similar to what I described above when it comes to materialism. Here, too, the foundation of the deterministic position is libertarianism. In other words, we can formulate a parallel principle to the cogito, one which applies to determinism: “I think, therefore I choose.” If I believe what I think (which means I think), I therefore make a choice. Without the ability to choose, or to use judgment, there would be no room for seeing me as a thinking creature. The Latin formulation would be: Cogito Ergo Eligo.

A Negative Formulation The conclusion from the book is that our ability to reach a deterministic conclusion or any other conclusion is also conditional on the faith we place in our judgment. Absurdly enough, it is libertarianism which allows this. Just as we saw in the previous, materialist-dualist context where qualia is what allows us to recognize the existence of matter. But while the materialist question deals with the question of what is more certain — spirit or matter — here we deal with the question of what is correct: Determinism or libertarianism? We must decide between the two, not just form a hierarchy among them. Our argument in chapter four was that if the determinist belief is only made possible by the libertarian foundation, then it refutes itself. As we saw there, this is the fool’s cycle, as we have a contradiction between assumption and conclusion. The libertarian cogito principle can be formulated thus: I believe (or espouse) determinism, therefore determinism is incorrect, since I can believe or disbelieve it, which means that this is subject to my judgment. Thus, the similarity to the cogito principle is only partial. In the original formulation, matter and spirit have existence, but the cogito shows that spirit is more solid and fundamental. After we proved the latter’s existence, we can then go on to confirm the existence of the former. Whereas in the second context, if we arrived at the libertarian conclusion, then determinism is untrue, and not simply less certain.

Back to Dualism What we can conclude similar to the original is a softer argument. I — meaning my consciousness, which is part of my spirit — do indeed observe inanimate nature and see — i.e., conclude — its deterministic character. I arrive at this conclusion, based on judgment and decision, regarding the fixed nature of the relations between cause and effect, and effectively recognize the principle of causality of inanimate nature, meaning the determinism of the laws of physics. We saw throughout the book that quantum theory and chaos theory are not irregular in this regard. Only spirit can be an exception to physical causality, because it does not belong to inanimate nature, and is therefore not subject to the deterministic picture which arises from the laws of physics. We can now see that the picture we described here does not contradict itself. We must moderate our cogito principle somewhat in its determinism-libertarianism context. Its formulation would be something like this: I understand that the physical world is deterministic, ergo I (my spirit) has free choice. Here we are no longer dealing with the fool’s cycle, as there is no contradiction between assumption and conclusion. The assumption is that I have freedom of decision and judgment, and by using these I conclude that the

material world around me — though not me myself — is deterministic. This is not a contradiction, as long as we adopt a dualist world view, in which spirit and matter are separated, and we recognize that material alone is subject to the laws of physics, but not the spirit which observes it and studies it. The free spirit recognizes that the world around it is deterministic. The error of the determinists is that they apply this conclusion for the consciousness which reached this conclusion in the first place — thus entering the fool’s cycle.

Summary The circle thus closes here. Libertarianism leads us back to dualism. The cogito of the libertarian means: I exist, ergo I have spirit (and not just material), which is necessarily not subject to the principle of causality (to which material is subject), and ergo I think (meaning my thought can be believed), and ergo I also choose. The two questions which accompanied us throughout the book — the question of materialism and the question of dualism — are now retied into a Gordian Knot. The attempt to cut this connection leads to contradictions. It’s true that there is no way to introduce freedom into physics, but it is certainly not true that physics forces us to arrive at determinism. The consideration of Lex specialis leads us to the conclusion that a coherent picture of the world must be dualist, with the understanding that there is spirit alongside the material. In effect, the cogito teaches us that it is not alongside it, but at its foundation…

Appendix A BRIEF DISCUSSION OF RELIGIOUS FAITH AND CAUSAL DETERMINISM, WITH AN EMPHASIS ON JUDAISM In chapter four, we dealt with the question of theological determinism, this being the conception that God’s omnipotence requires that he know what will happen in the future. We saw that there are also religious thinkers — albeit very few of them — who derive a deterministic approach from this, and we also pointed to some of the difficulties which it raises. But the main issue which this book deals with is physical-causal determinism, meaning a deterministic approach based on physicalism. We pointed to the differences between these deterministic approaches. Among other things, we saw that the solution of R. Hisdai Crescas, which explains that the freedom that we have is only in determining our attitude towards events and not our conduct itself, does not even pretend to answer the theological question-problem. In this appendix, we will expand and discuss the question of whether religious faith can fit with an approach of physical determinism. The religious faith I am dealing with here will be the Jewish one, but the following analysis can also hold for other religions. I already noted in chapter four that the Protestant Christian and Druze faiths are actually based on determinism — this determinism is very fatalistic for the Druze, while it is quite activist in Protestant Christianity; the approach of the Protestants is that the fate of man is sealed, and his actions only prove retroactively if he is one of God’s Elect or whether he belongs to the weak. It is therefore very important for them to try and succeed to be considered one of the elect. There is no need to point out that such an approach involves serious philosophical and logical problems, and has been subject to many philosophical critiques. 242 My goal in this appendix, then, is to briefly examine a position which combines Jewish religiosity with determinism. One of the prominent figures espousing such an approach in contemporary Israel is Prof. Haim Sompolinsky, who was mentioned already in the introduction of this book. Sompolinsky is a believing and religiously observant Jew who heads the Safra Center for Brain Sciences at Hebrew University, and he declares his materialist and deterministic views in various media such as TV and online. Sompolinsky is an important physicist and a very talented person, and is unquestionably one of the most prominent neuroscientists in the world today. His scientific successes and capabilities are impeccable. Precisely because of this it’s important for me to figure out if the religious worldview he presents is consistent, and that the loyalty he declares to the sources of halacha and faith which he openly declares fit with materialist approaches. Sompolinsky says that his motivation for creating such an approach lies in the findings of scientific research. He finds himself forced to create an innovative and highly irregular worldview in the Orthodox Jewish world, as to his understanding, these findings force him to adopt a materialist and deterministic position. My criticism of this worldview, which combines determinism and Jewish religious faith, is on two levels: (1) The question of motivation: As I showed throughout the book, scientific research does not require such a position at all. The motivation for creating such a philosophical position does not actually exist. (2) As I will briefly show in this appendix, such an approach does not sit with Jewish tradition, at least not in its accepted meaning.

Deterministic Approaches in the Sources I will not conduct an extensive survey of the sources, and will instead only touch on some of the central sources which are raised in these discussions. First, no-one disputes that most if not all of the Jewish thinkers held to a non-deterministic worldview, and many even saw it as a Principle of Faith. At the head of these is Maimonides, who wrote as follows in Chapter five, halachas 1-5 in his Hilchot Teshuva: Free will is granted to all men. If one desires to turn himself to the path of good and be righteous, the choice is his. Should he desire to turn to the path of evil and be wicked, the choice is his…This principle is a fundamental concept and a pillar [on which rests the totality] of the Torah and mitzvot as states: ’Behold, I have set before you today life [and good, death and evil].’ Similarly, states, ’Behold, I have set before you today [the blessing and the curse],’ implying that the choice is in your hands. Any one of

the deeds of men which a person desires to do, he may, whether good or evil…However, this is known without any doubt: That man’s actions are in his [own] hands and The Holy One, blessed be He, does not lead him [in a particular direction] or decree that he do anything…Consequently, the prophets taught that a person is judged for his deeds, according to his deeds - whether good or bad. This is a fundamental principle on which is dependent all the words of prophecy. The Maharal of Prague, who lived in the sixteenth century, and was one of the greatest Jewish thinkers throughout the generations (and who actually had more important and serious things to do than create Golems attributed to him in later stories). He writes in his book Derekh Hayyim [Path of Life], p. 148: Man was created in the image of God and he is similar in this to God, May He Be Blessed. Man created in the Image of God has a virtue that he is his own authority, and like God May He Be Blessed he does what he wills. And thus Man has permission to do what he wills and has choice. Meaning, the Maharal sees choice as the supreme virtue of Man, the attribute which makes him like God. However, we cannot deny that there are a number of Jewish sources — sources written by Jews — which express deterministic views. I already mentioned R. Hisdai Crescas and his book Or Hashem, where he devotes two whole sections to the question of choice. Jewish Studies scholar Aviezer Ravitzki claims that he expresses a more moderate view in the first section, according to which what is given to man is how he addresses events — but not the conducting of events, including his own actions. But in the second section, which Ravitzki claims was written later, he already expresses an entirely deterministic view. We could add a number of Hassidic sources to this pile, primarily from the school of the author of Mei Hashiloach from the first half of the nineteenth century and his student R. Tzadok Hacohen of Lublin, who ostensibly also espouses determinism (and indeed, Sompolinsky mentions them as sources for his worldview). The Mei Hashiloach is known for saying “All is in Heaven even the fear of Heaven” (ibid., v. 1, 4b), a paraphrase of the famous saying of the Sages that “All is in Heaven except for fear of Heaven (Babylonian Talmud Berachot 33b, and other sources). R. Tzadok follows suit in his book Tzidkat Hatzadik, section Mem (40): The core of repentance is that until God lights his eyes that damnations become merits, meaning that he should recognize and understand that all that he sinned was by the will of God May He Be Blessed just as [the Sages] of blessed memory ’And you turned their hearts‘ &c. He writes here and in a few other places that it is all from God — both the commandments and the transgressions. It would ostensibly appear that he thinks man has no choice at all, and that repentance is solely a matter of consciousness. However, the writings of R. Tzadok Hacohen contain quite a few sections which clearly point to his not being a determinist. This is also true when it comes to his teacher, the Mei Hashiloach. 243In fact, the continuation of the passage we brought above shows him writing the following: And the reason for the knowledge and the choice which the Ari of blessed memory at the end of the book four hundred shekels of silver that both are true each in place in and of themselves: In the place of choice there is no room for knowledge and in the place of knowledge there is indeed no room for choice. And when he conceives this enormous light then all his damnations returned from the depth of the knowledge of God, May He Be Blessed, and He and His knowledge and His will are all one. And since God May He Be Blessed wanted it then they are all merits and he receives the complete absolution of Yom Kippur… It is clear that he is speaking of two modes of thought which must be lived by, and not any kind of philosophical statement. Moreover, the writings of these Hassidic thinkers contains a radical conception of self-fulfillment and highly original individualism, something which doesn’t really sit well with determinism. Another example among many can be found in the same book, letter Mem-Gimel (43): Sometimes a person faces such a great test that it cannot be that he will not sin as they said (Berachot 32a) ‘What will the boy do if not sin’ &c and in this he is considered entirely compelled which God exempts. And also in the incitement of the inclination with great force which cannot be defeated

compulsion is related [and if God May He Be Blessed turned his heart then this sin is not a sin at all, only that the will of God May He Be Blessed was so] and see what they said (Ketubot 51a) regarding it beginning in compulsion even if she later screams that if he had not let her be she would have hired him, she is permitted to her husband, for what reason? The inclination dominated her; after all, this is considered complete compulsion even though it was based on her desire, in any event such a great inclination cannot be subdued by man and he is completely compelled and there is no punishment involved even though he violated a prohibition since he was compelled. But man himself cannot testify to this because perhaps he still had strength to subdue the inclination [and as I heard on this matter from the case of Zimri who erred in this]. He explains here that if man has no choice, he is considered to have been compelled and is thus exempt from punishment. It is therefore likely that in all the cases where there are punishments for sins, he would also consider this to have been for actions done of free will. In general, it does not seem possible to extract a truly deterministic position from the writings of these two. I will also add that in my humble opinion, this deterministic approach and its like are full of logical contradictions and conceptual vagueness, and as we know, it’s hard to base any argument on a text which contradicts itself. However, our aim here is not to fully flesh out the positions of R. Tzadok or any other thinker, to say nothing of the fact that these really are esoteric anecdotes. What matters is that of the thousands of sages and thinkers we have had over the generations, one can find almost no sign of deterministic positions, and even those that are present are fairly questionable. For me, these are only examples which come to examine the possibility of such thought within the Jewish world of faith. And if we nevertheless find some source which makes a deterministic statement, does that mean that this position is free of contradictions or acceptable? Certainly not. But I have no intention of sufficing here with ad hominem arguments, and I will therefore move on to a more fundamental discussion.

A Discussion of Determinism in Light of Halachic Concepts The issue R. Tzadok raises in the last quote, regarding punishment and compulsion, is worth another discussion. Ostensibly, there is no room for punishing people in the deterministic picture, as their actions are not the result of choice. The very fact that halacha punishes violators ostensibly proves that one cannot reconcile loyalty to halacha and determinism. But as we saw in chapter three, the determinist explains the purpose of punishment as “repair” and not education; the fact that a punishment is imposed on the crime affects us in itself (deterministically) to not sin and do “bad deeds.” But R. Tzadok raises the question in a slightly different manner. Halacha does not just punish the criminals. It also exempts those who committed them under compulsion. This means that the condition for punishment is that the crime be done out of free will. Let us stress that halacha even recognizes internal compulsion as exempting from punishment, as this is the gemara R. Tzadok himself brings (Ketubot 51b) regarding forbidden sexual relations based on an irresistible urge (“Beginning in compulsion and ending in desire”). But determinism is after all based on man not really having choice when it comes to his actions. In the deterministic picture, all crimes are done out of compulsion, at the very least internal compulsion. According to the determinist, there is no room to exempt one from punishment in this situation, as he considers every criminal as “beginning and ending in compulsion.” This matter is of course related to the issue we dealt with in chapter two, in the discussion on the rule of R. Ilay. There we saw that if a man sees that his inclination is overcoming him and he is being overwhelmed by an urge which cannot be resisted, he can go somewhere distant, wear black and commit the sin there. For the determinist, everyone who commits a sin is in this position. The determinist could of course explain that the rule of R. Ilay deals with a case where a person’s estimation is that they are about to sin and cannot “overcome” — meaning that he is aware of what is about to necessarily happen (as opposed to other cases, in which the future also necessarily happens, but the person either doesn’t know or cannot assess the future). This is a very strained explanation, of course. In addition, there is a whole range of psychological terminology in halacha which is involved in evaluating crimes. For instance, a man who mistakenly committed the sin is exempt. If the purpose is deterrence, why should he be? After all, there are no considerations of fairness in punishment, as punishment is not moral retribution but a mechanical means of correction. So why impose a punishment on someone who mistakenly sins? Another example is the exemption for someone who unintentionally sinned. In halacha, such a person is also exempt from punishment. This is not necessarily someone who didn’t intend to violate the commandment but one who was not interested in the sin. For instance, a person dragged a bench on

Shabbat, while creating a line in the ground (forbidden because of the prohibition on ploughing). He knew that he was doing a sin (or that it was likely to happen), but this was not the goal of his action. Since he did not intend to plough the field but move the bench, he is exempt. Another exemption is “not convenient for him.” When a person does an action without intent like dragging the bench, when the creation of a line in the ground is certain in advance — he is liable. But what if he doesn’t want that line in his backyard? Most Jewish religious rulers say that he’s exempt. Why is it matter if he’s comfortable with the result or not? After all, what matters is the sin he violated. Another concept in this conceptual family is “fiddling,” a person handling something without even really noticing. For instance, a person walks by a tree branch on Shabbat and removes the branch while walking into it. This case is also exempt from punishment. All these concepts raise serious questions in the deterministic picture, as it is not clear why halacha cares about intent if the aim is to simply avoid deeds. The mental epiphenomenon should not be part of the punishment calculus, as the punishment is only coming to correct matters at the neural level, and it is not clear why halacha even sees fit to address it. Thus far, sins. But the question of intent also refers to positive commandments. Per all religious opinions, doing the commandment is not sufficient. Intent is also needed. Some believe that this is not disqualifying and that fulfilling commandments without intent also has value, but the unanimous position is that there should be intent from the outset. Another world of terminology where we see the significance of the mental dimension and of freedom of choice is the field of acquisition. Halacha requires, as a condition for a person acquiring an object, that there be complete intent on the part of both sides, with both clearly deciding that they wish to acquire and grant the object. Without this, there is no acquisition. Jewish halachic rulers are split on the question of whether the act creates the acquisition and the decision is a condition therefor, or the decision creates the acquisition and the action is only needed to form the decision. In the deterministic word, this entire discussion is obviously entirely meaningless. What’s the difference between a person who makes an action of granting without a clear decision to grant, and someone who does so with clear intent? After all, the decision is the deterministic result of a brain state, and in effect every such decision means nothing more than the brain state which led to the action of acquisition. What additional meaning can a decision, which is just epiphenomena, have in the deterministic world? This is true of any other halachic relationship where the person’s opinion is required in addition to his action: conversion, marriage and divorce, and many other fields — all these lose their meaning in the deterministic world. Even halachic concepts which do not belong to the legal sphere involve intent and decision. Concepts like repentance and regret, which are a condition for being forgiven for sins, also lose their significance in the deterministic view, as these actions are nothing more than the result of neural physics. What does it mean to regret? To feel an illusion of regret?! After all, I did not decide to regret; this is a situation I found myself in. Even if we accept the existence of a psychological state of regret as a kind of epiphenomenon, why does this state even matter? Why is this the condition for repentance and exemption from sin? And what of a person who makes a vow? Halacha allows him to go to a Rabbi who will release him from the vow. This release requires regret and an opening, meaning the person is examined as to whether he mistakenly made the vow and then the vow is removed. But if the vow is always done due to internal compulsion, what does it mean to release the vow? What is the regret? Nothing more than an epiphenomenal illusion which accompanies this or that neural activity. In light of this, it’s really not clear what the significance of the Rabbi releasing the vow is, as he is also compelled to do what he does. This entire halachic picture appears entirely surreal when viewed through the deterministic prism. And above all, what does it mean to repent in the deterministic world? To have an illusory feeling of repentance which accompanies this or that neural activity? What does it mean to feel regret for my past deeds? After all, if I am compelled to act, I am also compelled to regret now. So why does the Torah even command it? Even if we understand the Torah’s commandments to be aimed at influencing me so that I act towards practical goals, what is the aim of the commandment (or recommendation) that I repent? R. Hisdai Crescas addresses this question in Or Hashem, but there he offers (article 3, section 2) that repentance is indeed not possible but that God’s grace allows us this option. What does this even mean? It would appear to me that the accepted ways of dealing with these issues, both by R. Crescas and the aforementioned Hassidic literature, are based on conceptual vagueness and nothing more. This vagueness allows one to say something and its opposite without feeling the contradiction. If we’re already dealing with punishment, we could also ask about the ultimate reward — the World to Come, and subsequently the meaning of the belief in the immortality of the soul. Why impose punishments on the person who sinned, if he is compelled to? As I already noted, normal legal systems can find logic

even in the deterministic picture, according to which the legal system is trying to correct human behavior, with the punishment being nothing more than an additional constraint whose aim is to make him behave differently. But punishments in the World to Come — or punishments like death from Heaven, which no-one knows when and if they happen — do not fix anything, as they are after all imposed after our death and noone knows about them. So what is their role? At most, we could justify talk of punishments and the World to Come, as such talk may change our behavior in this world. But the World to Come itself, and certainly punishment in the World to Come, do not seem reasonable, and there is no reason at all that God will actually carry out the punishments. Thus, logic says to speak of the World to Come but not carry out the punishment, meaning to lie. But now even such false talk doesn’t achieve its aim, as I know from logical reasoning that it isn’t true. Moreover, all this emerges only if we can even fathom the nature of the entity being punished there, and after all — our soul doesn’t exist as far as the determinist is concerned (as we saw, he is usually a materialist, as well). So who exactly is being punished after our death? Alternatively, we could of course allegorize it all, and tell ourselves a surreal story of all our actions being but shadows of the real world. In truth, everything is mechanical and illusions are epiphenomena, and the story told to us is a lie which appeals to that epiphenomenal illusion, meaning appealing to the feelings and insights and experiences even though they have no effect on effective behavior. This is but a logical conclusion from that approach which sees the whole world and all human feelings and insights as a collection of illusions. It’s particularly surprising that all this is done in the name of science and reason.

The Concept of Choice in the Written and Oral Torah Thus far, we dealt with the teachings of the Sages and the Medieval Jewish commentators, which serve as the core of halacha. All these are people, and we can therefore perhaps say of them that when they speak of mental events and experiences, they are speaking in their own language and based on their own world of conceptions. They thought we have choice, but were wrong in this assumption due to the lack of scientific knowledge in their era. They thought that the soul was immortal and that there is a World to Come, and they were wrong on this as well. So how does the Torah itself deal with these questions? The Bible requires us to choose in a number of places. For instance, chapter 30 of Deuteronomy states: If you obey the Lord your God and keep his commands and decrees that are written in this Book of the Law and turn to the Lord your God with all your heart and with all your soul. Now what I am commanding you today is not too difficult for you or beyond your reach. It is not up in heaven, so that you have to ask, “Who will ascend into heaven to get it and proclaim it to us so we may obey it?” Nor is it beyond the sea, so that you have to ask, “Who will cross the sea to get it and proclaim it to us so we may obey it?” No, the word is very near you; it is in your mouth and in your heart so you may obey it. See, I set before you today life and prosperity, death and destruction. For I command you today to love the Lord your God, to walk in obedience to him, and to keep his commands, decrees and laws; then you will live and increase, and the Lord your God will bless you in the land you are entering to possess. But if your heart turns away and you are not obedient, and if you are drawn away to bow down to other gods and worship them, I declare to you this day that you will certainly be destroyed. You will not live long in the land you are crossing the Jordan to enter and possess. This day I call the heavens and the earth as witnesses against you that I have set before you life and death, blessings and curses. Now choose life, so that you and your children may live and that you may love the Lord your God, listen to his voice, and hold fast to him. For the Lord is your life, and he will give you many years in the land he swore to give to your fathers, Abraham, Isaac and Jacob. These verses speak of repentance and on the granting of two options and of the duty to choose a life of commandments and avoidance of sin. But according to the determinist, we don’t really have the option of either. For him, we don’t have two paths before us but one, and it is not we who choose it. Moreover, the Torah is trying to convince us that this is not some distant thing, neither high as the sky nor beyond the sea, and tries to aid us in this choice, clearly expressing a position that this is indeed in our hands. How does a person who believes in Torah from Heaven address the verses in the Torah which speak in these terms? The assumption of the believing Jew is that the entire Torah was written by God, so does even God not know the truth? Or is His aim perhaps to confuse us, or manipulate us in one way or another? Even if we accept that these words are not meant to reflect truth but only influence us (meaning “correct” us, like one fixes a drill or car), and even if we would be willing to accept God as being such a

manipulator, it’s still not clear why all this is required, anyway. After all, according to religious belief, God created us as well as the world and its rules. Why? According to the determinist, He certainly did not want us to choose, as we lack choice. Thus, He simply wanted us to do a set of actions. In such a state of affairs, there is a much simpler way to achieve this instead of strange manipulations of our mind — instead of appealing to us in such indirect ways and presenting an illusory picture as though we have choice, why not just directly program us to follow His will? Moreover, the way which God chose keeps failing. After all, we all sin every so often, and don’t always do the will of God. So why does an entity with infinite knowledge and ability choose such an unsuccessful method? Are people indeed punished and rewarded according to their deeds (if at all), and all of this is but a manipulation? Is the serious punishment of the people of Sodom only imposed to educate their surroundings (since they already died and cannot learn the lesson), even though they bear no responsibility or guilt for their actions? Beyond the injustice involved, it’s not clear what the purpose of the commandments are in this scenario; why shouldn’t I just sit idle instead of observing the commandments? 244 Put differently: Is there room for condemning me if I don’t? Why, then, do the Torah and the Sages condemn the sinners and reprimand them for not overcoming their inclination and sinning? Why does the Torah call some sins words like “abomination” and define someone who committed murder as one who defiled the land which will not be forgiven except with the blood of the murderer (See Num. 35:33), when this murder is in effect the direct result of the actions of God Himself? The believing determinist would appear to think that all this is nothing more than a collection of manipulative schemes whose aim is to “fix” us and “re-educate” us, instead of just creating us in the correct way in advance. We should remember that after all this, we are still supposed to accept the characterization of God as being merciful and patient, and whose signature is truth (see Ex. 34:6 and others). But the problem is not just God’s good and truth, but also His intelligence. As I already noted, the main manager of this circus, the one who holds all the strings of all the puppets, cannot create a world which will do what He Himself wants. After all, we sin every chance we get, and this contrary to His explicit will (unless His will is also a manipulation). Moreover, He is even sufficiently unsuccessful to fail at hiding this particular scheme. Why would He allow us to reach the understanding that our world is an illusion and that He is manipulating us? After all, He is the one who enabled us to use the tools of science and our reasoning to arrive at the deterministic conclusion, and thus bring down the tent on the whole show. Sounds a little strange to attribute infinite intelligence to such an entity, no? The determinist is effectively continuing in the same path we encountered throughout the book. He continues to speak of illusions and shadows, on mistaken beliefs and hidden manipulations. For him, just as our insights about ourselves and life are untrue, the same is true when it comes to Torah and its commandments. Moreover, the Torah itself aids in creating this grand illusion, but with the generous aid of neuroscience (which alone is of course nothing but the hard truth), we finally know all this. The picture presented by the determinist is a kind of caricature of the world, of man, of religion, and effectively of God. For him, God built Himself is a virtual game in which programmed creatures play before Him, live in the illusion that they decide how to act, as though fateful issues hinge on what they choose to do, while all this is but a rigged game. The results were set from the first moment. Like any good game or book, the good guys always win in the end.

Can God Even Be Present in a Deterministic Picture? We saw throughout the book that causal determinism is based on physicalism. At its foundation is the assumption that the entire world is nothing but the material, and it is thus entirely controlled by the laws of physics. The physicalist usually argues that there is no evidence for the existence of anything outside physics, and there is therefore no reason to assume its existence. As we saw in chapters nine and ten, there are no gaps in the physicalist framework, and it therefore dictates determinism. However, if we look at the world from a materialist angle, then it is not clear how it is possible to believe in an abstract spiritual entity like God. What are the signs of His existence, and are those signs not illusions? Why are they fundamentally different than the signs of the existence of the soul? Even direct revelation, like that which occurred at Mount Sinai (for those who believe this occurred), should be interpreted by physicalism in terms of the laws of physics, as sounds and lightning and fire are also physical phenomena, and such phenomena are formed solely by physical causes. That is the essence of physicalism. Moreover, it’s not clear how the physicalist understand interaction between God and ourselves and how this is even possible. After all, one of the fundamental assumptions which leads physicalists to their approach is that of the inability to connect the spiritual and the material, otherwise known as the

psychophysical problem. For them, the spiritual cannot operate on anything physical. But such a revelation is conceived in our senses, meaning the encounter with a spiritual, non-physical entity creates interaction with man, a physical entity, and drives electrons in his brain, and this is impossible in the physicalistic view. So how can a connection form between us, creatures made of matter, and God, who is all spirit? The conclusion is that not only is God’s existence not possible in the deterministic picture, even if He did exist — His revelation to us would not be possible. And what of philosophical arguments for God’s existence? These are certainly not stronger than the arguments for the existence of the soul (see the prologue, for instance, where we saw Descartes prove the existence of God based on his conclusion regarding the existence of the thinking dimension within us). I’m not talking here of demons and angels, as their existence is certainly not necessary for the believing person. It may be possible to speak of them in metaphorical terms — certainly when it comes to demons and perhaps also angels. But if God Himself is a metaphor, all that’s left is emotional religiosity and nothing more. If that’s what is being reconciled with deterministic materialism, it would appear to me that this is wasted effort. Emotions don’t need to fit facts or science. 245 It’s important to stress here that this difficulty does not exist in the thinking of the theological determinists we mentioned above. They base their determinist approach on God’s omnipotence, and not on the principle of causality or physicalism. There is therefore no reason for them not to accept God’s existence as a fact, as they do not deny the spiritual dimension of reality and are not materialists, nor do they deny the possibility of contact between spirit and material. To the contrary, it is their spiritual approach which makes them see the Great Spirit they call God as something which determines everything we do. Their approach is that of dualistic determinism, which as I noted a number of times throughout the book is possible at the logical level. My argument here is only that determinism of the sort espoused by Sompolinsky based on physicalism does not sit with belief in God.246 We saw that causal determinism based on physicalism empties morality, science, and effectively our most basic insights of any real content. Many of them thus become fictions and illusions. We saw here that such an approach even includes God Himself to this long list of fictions. How, then, can one argue that determinism can seriously dwell alongside religious faith?

Back to the Atheist As noted, the point of this appendix is just to briefly examine the problems involved in combining determinism and religious faith. But as I already noted in the introduction, the fact that this debate is presented as one between the believer and the atheist is itself part of the problem. As we saw throughout the book, determinism is entirely unreasonable even for the complete atheist, or at least without foundation, beyond the fact that it is also morally dangerous. But for the atheist, this is a position which is unreasonable but still possible in principle. This appendix is meant to show that this position is inconsistent for the believer and contains internal contradictions. For a believing Jew, determinism is entirely impossible, unless one pays the heavy price it demands, meaning the emptying of faith of any concrete content and its replacement with nothing more than religious sentiment and God, the subject of faith, into one big illusion (as though they are but epiphenomena of electrical currents and neural processes). Such a “faith” is indeed possible, and it sits well with the deterministic system. But I don’t think that I would accept such “faith” as religious faith. To sum up, everyone has the right to define his own outlook and worldview, and he also has the right to hold to delusional and inconsistent positions. But the use of the word “Faith” in reference to all these seems to me to deviate from accepted dictionary definitions, and it is therefore preferable to find another term. Those sociologists who are deluded to think themselves a mathematician have the right to do so. But I don’t think I’d be willing to have him join a conference of mathematicians, or rely on his findings in this field, or publish his articles in a mathematical journal. Those who think themselves Napoleon are usually not given command of an army as opposed to a padded cell. So why is someone who lives with the illusion of belief in God, to say nothing of a God which is an illusion and self-contradictory — omnipotent but utterly helpless, or an entity which is entirely true but which lies all the time — entitled to be included among the religious? Maybe the religious determinist has the answers… 242 Prof. Mordechai Rotenberg from Hebrew University’s School of Social Work criticizes Protestantism from the psychiatric direction. Rotenberg and his students deal with methods of diagnostics and psychological treatment, claiming that the philosophical foundation of diagnosis and treatment define its character and results. 243 This has been discussed in many places. To start, see Rabbi Shagar’s (Hebrew) book Shuvi Nafshi (p. 126 passim). 244 This is, in effect, the lazy man’s argument from chapter four. 245 This common confusion between belief and religious emotions of experience is discussed in some detail in my first article in the 2011 Ynet

series (Hebrew) “Is Belief in God Rational?” http://www.ynet.co.il/articles/0,7340,L-4095282,00.html 246 Indeed, the argument that we have no judgment in the deterministic picture and therefore the conclusions forced on us — determinism itself included — have no validity, also undermines determinism of that sort.