The Gunung Padang archaeological site in western Java was built by a civilization 25,000 years ago      $1 million to anyone who solves one of the 7 hardest math problems in the world – The Riemann Hypothesis      Russian philosophy. Ancient Rus’. Romanticism. Slavophilism and Westernism. Philosophy and power      Tibetan Book of the Dead (Full text). “Great Liberation as a result of what was heard in the bardo”      History of the development of Buddhism in Russia     
Русский язык  English  French  Deutch  Spain

Category Archives: Library

Goswami’s book questions the existence of an “external”, real, objective reality. It is argued that the universe is self-aware and it is consciousness itself that creates the physical world and explains how a single consciousness appears to be so many separate consciousnesses.

Goswami’s book is an attempt to bridge the age-old gap between science and spirituality through a monistic idealism that resolves the paradoxes of quantum physics.

The author of the book is a physicist and professor at the Institute of Theoretical Sciences at the University of Oregon.

INTRODUCTION

When I was studying quantum mechanics as a graduate student, we used to spend hours discussing complex questions like, “Can an electron really be in two places at once?”
I could accept this – yes, an electron can be in two places at the same time: quantum mechanics gives an unambiguous answer to this question, although full of subtleties. But do ordinary objects—say, a chair or a table, those things we call “real”—behave the same way as an electron? Does such an object become a wave, inexorably beginning to spread in a wavelike manner when no one is looking at it?

Objects encountered in our everyday experience do not seem to behave in the strange ways typical of quantum mechanics. So it’s easy for us to unconsciously convince ourselves to think that macroscopic matter is different from microscopic particles—that its ordinary behavior is governed by Newton’s laws, which are called classical physics. Indeed, many physicists stop racking their brains over the paradoxes of quantum physics and surrender to this solution. They divide the world into quantum and classical objects – as I did myself, although I was not aware of what I was doing.

To have a successful career in physics, you can’t think too much about stubborn questions like quantum paradoxes. I’ve been told that the pragmatic way to do quantum physics is to learn to calculate. So I compromised, and the painful questions of my youth gradually faded into the background.

However, they did not disappear. Circumstances changed, and—after the nth bout of stress-induced heartburn that had characterized my entire career as a successful physicist—I began to remember the richness of the feelings I had once felt about physics. I knew there had to be a way to pursue this subject that brought joy, but I needed to rekindle my spirit of exploring the meaning of the universe and abandon the mental compromises dictated by career considerations. I found a lot of help in Thomas Kuhn’s book, which distinguishes between paradigm research and scientific revolutions that lead to paradigm shifts. I’ve already done my share of research within the paradigm; it was time to go to the forefront of physics and think about a paradigm shift.

My personal turning point roughly coincided with the publication of Fridtjof Capra’s book The Tao of Physics. Although my initial reaction to the book was one of suspicion and rejection, it nevertheless affected me deeply. After some time, I was able to understand that the book raises an issue that is not thoroughly explored in it. Capra touches on the parallels between the mystical worldview and the ideas of quantum physics, but does not explore the reasons for these parallels: are they more than a coincidence? I finally discovered where my research into the nature of reality should focus.

Capra approached questions about reality from the perspective of particle physics, but I intuitively felt that the key questions were most directly related to the problem of interpreting quantum physics. This is what I decided to explore. Initially, I did not expect it to be such an interdisciplinary project.

I was teaching a science fiction physics course (I’ve always had a soft spot for science fiction) and one student remarked, “You sound like my psychology professor, Caroline Keutzer!” This resulted in a collaboration with Keutzer, which, although it did not lead to any serious insights, nevertheless introduced me to a large amount of important psychological literature. Eventually, I became aware of the research of Mike Posner and his group in cognitive psychology at the University of Oregon, which was to play a critical role in my work.

In addition to psychology, my subject of research required significant knowledge in neurophysiology – the science of the brain. I met my neurophysiology teacher through the famous dolphinologist John Lily. Lily kindly invited me to participate in a week-long seminar he was teaching at the Esalen Institute; Among the participants was Dr. Frank Burr, MD. While quantum mechanics was my passion, Frank was passionate about brain theory. I could learn from him almost everything I needed to begin working on the mind-brain aspect of this book.

Another important component in the formation of my ideas were the theories of artificial intelligence. I was very lucky here too. One of the popularizers of the theory of artificial intelligence, Douglas Hofstadter, began his career as a physicist; he is a graduate student at the University of Oregon, where I teach. Naturally, when his book came out, I was especially interested in it, and I got some of my key ideas from Doug’s work.

The significant coincidences continued. I became familiar with research in parapsychology through many discussions with another of my colleagues, Ray Hyman, who is by nature a very open-minded skeptic. The last but not least important coincidence was my meeting in the summer of 1984 in Lone Payne, California, with three mystics: Franklin Merrell-Wolf, Richard Moss and Joel Morewood.

Since my father was a Brahmin guru in India, I, in a sense, grew up in an atmosphere of mysticism. However, at school I began my long departure from it through traditional training and practice in a separate field of science. This direction took me away from my childhood sympathies and made me believe that the only reality was the objective reality determined by conventional physics, and everything subjective was due to the complex dance of atoms that we would someday decipher.

In contrast, the mystics at Lone Payne spoke of consciousness as “primordial, self-sufficient and formative of all things.” At first, their ideas caused me considerable cognitive dissonance, but over time I realized that you can still do science, even if you consider consciousness rather than matter to be primary. Moreover, this way of doing science dispels not only the quantum paradoxes of my youth, but also the new paradoxes of psychology, the brain, and artificial intelligence.

So this book represents the end result of my circuitous journey. It took me ten to fifteen years to overcome my addiction to classical physics and then to conduct research and write a book. I hope that the fruit of my efforts deserves your attention. To paraphrase Rabindranath Tagore:

I listened and watched with an open mind,
I poured my soul out into the world,
seeking the unknown in the known,
And I scream out loud in amazement.

Obviously, many people besides those mentioned above contributed to the book, including Gene Varnet, Paul Ray, David Clarke, John David Garcia, Suprokash Mukherzhdi, Jacobo Ginberg, the late Fred Attneave, Ram Dass, Ian Stewart, Henry Stapp , Kim McCathy, Robert Tompkins, Eddie Oshins, Sean Bowles, Fred Wolf, Mark Mitchell and others. The encouragement and support of friends, including Susan Parker Barnett, Kate Wilhelm, Diamon Knight, Andrea Pucci, Dean Kisling, Fleetwood Bernstein, Sherry Anderson, Manoj and Deeptu Pal, Geraldine Moreno-Black and Edd Black, and my late colleague Mike Moravcsik, were essential. and especially our late beloved friend Frederika Leif.

I especially thank Richard Reed, who convinced me to submit the manuscript for publication and passed it on to Jeremy Tarcher. In addition, Richard provided important support, providing helpful criticism and assisting with editing. Of course, my wife Maggie contributed so much to both the development of ideas and the language in which they are expressed that without her this book would literally not be possible. My heartfelt thanks go to J. Tarcher’s editors, Aidin Kelly, Daniel Mulvin, and especially Bob Sheppherd, as well as Jeremy Tarcher himself, for believing in this project.

Thank you all.

PREFACE

Not long ago, we physicists believed that we had finally completed our search: we had reached the end of the road and discovered a mechanical universe, perfect in all its splendor. Things behave the way they do because they were that way in the past, They will be the way they will be because they are that way in the present, and so on. Everything fits perfectly within the narrow framework of Newton’s and Maxwell’s laws. There were mathematical equations that actually corresponded to the behavior of nature. There was a one-to-one correspondence between the symbol on the page of a scientific article and the movement of any objects – from the smallest to the largest – in space and time.

The nineteenth century was ending when the famous A. A. Michelson, speaking about the future of physics, declared that it would consist in “adding decimal places to the results already obtained.” In fairness, it should be noted that Michelson, when making this remark, believed that he was quoting the famous Lord Kelvin. In fact, it was Kelvin who said that essentially everything in the landscape of physics is perfect, except for two dark clouds blocking the horizon.

It turned out that these two dark clouds not only blocked the sun from Turner’s landscape of Newtonian physics, but turned it into a bewildering abstract painting of dots, spots and waves in the spirit of Jackson Pollock. These clouds were the harbingers of the now famous quantum theory of everything.

Now we have reached the end of a century again, this time the twentieth, and clouds are once again gathering to obscure the landscape of even the quantum world of physics. As before, the Newtonian landscape had and still has its fans. It remains suitable for explaining a wide range of mechanical phenomena, from spaceships to automobiles, from satellites to can openers; and yet, when quantum abstract painting eventually revealed that the Newtonian landscape is made up of seemingly random dots, many of us still believe that ultimately there must be some kind of underlying everything – and even quantum dots. that is a kind of objective mechanical order.

You see, science starts from a very fundamental assumption about how things are, or should be. It is this assumption that Amit Goswami, with the assistance of Richard E. Reed and Maggie Goswami, questions in the book you are about to begin reading. For this admission, like its cloudy predecessors in the last century, appears to signal not only the end of a century, but the end of science as we know it. This assumption is that there is an “external”, real, objective reality.

This objective reality is something fundamental: it consists of things that have attributes such as mass, electric charge, angular momentum, spin, position in space and continuous existence in time, expressed as inertia, energy, and even deeper in the microcosm – such properties like strangeness, charm and color. And yet the clouds still gather. For despite everything we know about the objective world, even taking into account all its unexpected twists and turns of space into time and into matter, and black clouds called black holes, even with all the power of our rational minds rushing forward at full speed, we are still left with many secrets, paradoxes and puzzle pieces that simply have nowhere to fit.

But we physicists are a stubborn lot, and we are afraid, as the saying goes, to throw the baby out of the bath with the dirty water. We still lather and shave our faces, being careful how we use Occam’s razor to ensure that we remove all unnecessary “dangerous assumptions.” What are these clouds that shadow the end of the twentieth century abstract art form? They boil down to one phrase: apparently, the universe does not exist without someone who perceives it.

Well, on some level it certainly makes sense. Even the word “universe” was invented by man. So in a sense we can say that what we call the universe depends on the ability of human beings to create the world. But is this observation something deeper than just a matter of semantics? For example, did the universe exist before human beings? It would seem, yes, it existed. Did atoms exist before we discovered the atomic nature of matter? Again, logic dictates that the laws of nature, forces and causes, etc., must surely exist, even though we knew nothing about such things as atoms and subatomic particles.

But it is precisely these assumptions about objective reality that have challenged our modern understanding of physics. Let’s take, for example, a simple particle – an electron. Is it a small piece of matter? The assumption that he is such and consistently behaves as such turns out to be clearly incorrect. After all, at times it appears to be a cloud consisting of an infinite number of possible electrons, which “looks” like a single particle if and only if we observe one of them. Moreover, when it is not a single particle, it appears as a wave-like, oscillating cloud capable of moving at speeds faster than the speed of light – in complete contradiction to Einstein’s concern that nothing material can travel faster than light. But Einstein’s concern is in vain, for when an electron moves in this way, it is not really a particle of matter.

Let’s take another example – the interaction between two electrons.
According to quantum physics, even though these two electrons may be vast distances from each other, the observations being made indicate that there must be some kind of connection between them that allows the message to travel faster than light. However, before these observations, before a conscious observer decided to make them, even the form of the connection was completely uncertain. And, as a third example, a quantum system such as an electron in a bound physical state appears to be in an uncertain state, and yet the uncertainty can be resolved into components of certainty that somehow add up to the original uncertainty. Then comes the observer who, like some giant Alexander cutting the Gordian knot, resolves the uncertainty into a single, definite but unpredictable state simply by observing the electron.

Moreover, the sword strike could occur in the future, determining what state the electron is in now. For now we even have the possibility that observations in the present legitimately determine what we can call the past.

Thus, we have again come to the end of the road. There is too much quantum supernaturalism around, too many experiments showing that the objective world is a world that moves forward in time like a clock, which says that action at a distance, especially instantaneous action at a distance, is impossible, which says that a thing cannot be in two or more places at the same time, represents an illusion of our thinking.

So what should we do? Perhaps this book has the answer. The author puts forward a hypothesis that is so alien to our Western mind that we want to immediately discard it as the delirium of an Eastern mystic. She argues that all of the above paradoxes are explainable and understandable if we abandon the dear assumption of the existence of an “external” objective reality independent of consciousness. She says even more – that the universe is “self-conscious” and that it is consciousness itself that creates the physical world.

By using the word consciousness, Goswami is implying something perhaps deeper than you or I would imply. In his understanding, consciousness is something transcendental, located outside of space-time, non-local and all-pervasive. It is the only reality, but we are able to gain some idea of ​​it only through action, which gives rise to the material and mental aspects of our processes of observation.

But why is it so difficult for us to accept this? Perhaps I am taking on too much by saying that it is difficult for you, the reader, to accept. Perhaps you find this hypothesis self-evident. Well, sometimes I’m quite happy with it, but then I bump into a chair and hurt my leg. That old reality intrudes again, and I “see” myself as different from the chair, cursing its position in space, so arrogantly separate from mine. Goswami addresses this issue brilliantly and gives several often amusing examples to illustrate his claim that the chair and I arise from consciousness.

Goswami’s book is an attempt to bridge the age-old gap between science and spirituality, which he believes is achieved by his hypothesis. He has a lot to say about monistic idealism and how it alone resolves the paradoxes of quantum physics. He then looks at the age-old problem of mind and body, or mind and brain, and shows how his overarching hypothesis that consciousness is everything heals the Cartesian divide, and in particular – in case you were wondering about it – even , how one consciousness appears to be so many separate consciousnesses. Finally, in the last part of the book, he offers a glimmer of hope as we move through the clouds into the twenty-first century, explaining how this hypothesis will, in fact, lead to a return to man’s fascination with his environment, which we certainly need. He explains how he experienced his own theory when he realized the mystical truth: “for true understanding, nothing-but-consciousness must be experienced.”

While reading this book, I began to feel this too. Provided the hypothesis is true, you will also have this experience.
ed., Alain Wolf, Ph.D.,
author of the books “The Dreaming Universe”,
“Making the Quantum Leap”, etc.
La Conner, Washington

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

PART I. Integrating Science and Spirituality

The turmoil in today’s world is reaching critical levels. Our faith in the spiritual components of life—in the living reality of consciousness, values, and God—is crumbling under the relentless attacks of scientific materialism. On the one hand, we welcome the benefits that science provides, which presupposes a materialistic worldview. On the other hand, this dominant worldview cannot satisfy our intuitions about the meaning of life.

Over the past four centuries we have gradually come to believe that science can only be built on the idea that everything is made of matter – so-called atoms in the void. We have come to dogmatically accept materialism, despite its inability to explain the most common experiences in our daily lives. In short, we have an inconsistent worldview. Our situation has created the need for a new paradigm—a unifying worldview that integrates mind and spirit into science. However, no new paradigm has emerged.

This book offers such a paradigm and shows how we can develop a science that accepts the world’s religions, working with them to understand the whole of human existence. The basis of this paradigm is the recognition that modern science confirms the ancient idea that the basis of all things is not matter, but consciousness.

The first part of the book introduces new physics and the modern version of the philosophy of monistic idealism. On these two pillars I will try to build the promised new paradigm – a bridge across the gap between science and religion. Let communication be possible between them.

CHAPTER 1. THE Abyss and the Bridge

I see a strange, torn caricature of a man beckoning me towards him. What is he doing here? How can it exist in such a fragmented state? What should I call him?

As if reading my thoughts, the distorted figure says: “What significance does a name have in my condition? Call me Guernica. I’m searching for my consciousness. Don’t I have the right to consciousness?”

I recognize this name. “Guernica” is a brilliant painting created by Pablo Picasso as a protest against the fascist bombing of a small Spanish town of the same name.

“Okay,” I answer, trying to calm him down, “if you tell me what exactly you need, then maybe I can help.”

“You think? — His eyes light up. “Perhaps you can speak in my defense?” He looks at me hungrily.

“In front of whom? Where?” – I ask, puzzled.

“Inside. They gathered there while I was left here unconscious. Perhaps if I find my consciousness, I will be whole again.”

“Who are they?” – I ask.

“Scientists, those who decide what is real.”

“Yes? Then the situation can’t be that bad. I am a scientist myself. Scientists are open-minded people. I’ll go talk to them.”

The people at the party are divided into three separate groups, similar to the islands of the Bermuda Triangle. After a moment’s hesitation, I resolutely head towards one of these groups – they don’t go to someone else’s monastery with their own rules, and all that. There is a lively conversation between them. They talk about quantum physics. They must be physicists.

“Quantum physics makes predictions about experimentally observed events, and nothing more,” says an aristocratic-looking man with barely noticeable gray hair. “Why make unfounded assumptions about reality when talking about quantum objects?”

“Aren’t you tired of this position? It seems that a whole generation of physicists has been conditioned to think that an adequate philosophy of quantum physics was developed sixty years ago. This is simply not true. Nobody understands quantum mechanics,” says another with noticeable sadness.

These words go almost unheard when another gentleman with a wild beard declares with unquestionable authority: “Listen, let’s speak directly. Quantum physics states that objects appear as waves. Objects are waves. And waves, as we all know, can be in two (or more) places at the same time. But when we observe a quantum object, we find it entirely in one place—here, not there, and certainly not here and there at the same time.”

A bearded man waves his arms excitedly. “So what does this mean in layman’s terms? “Here you are,” he says, looking at me, “what do you say?”

I get lost for a moment, but quickly come to my senses. “Well, apparently our observations, and thus ourselves, have a profound influence on quantum objects.”

“No. No. No, the questioner exclaims angrily. – When we observe, there is no paradox. When we are not observing, the paradox of an object being in two places at the same time returns. Obviously, the way to avoid the paradox is to vow never to talk about the location of the object between observations.”

“But what if we, our consciousness, actually influence quantum objects?” – I persist. Somehow it seems to me that the consciousness of Guernica has something to do with this hypothesis.

“But this means mind over matter,” all the people in the group exclaim in unison, looking at me as if I had uttered heresy.

“But, but…” I stammer, refusing to be tamed, “suppose there is some way to reconcile the primacy of mind over matter.”

I tell them about Guernica’s predicament. “Look, you have a responsibility to society. You have known for sixty years that the normal, objective way of doing physics is not suitable for quantum objects. We get paradoxes. And yet, you pretend to be objective, and the rest of society misses the opportunity to learn that we – our consciousness – are intimately connected to reality. Can you imagine how it would affect the worldview of the average person if physicists explicitly recognized that we are not separate from the world, but on the contrary, we are the world and must be responsible for it? Perhaps only then could Guernica—indeed, all of us—return to wholeness.”

The important gentleman intervenes: “In the dead of night, and when no one is around, I would admit that I have doubts. But my mother taught me that when in doubt, it is much better to feign ignorance. We don’t know anything about consciousness. Consciousness belongs to psychology, to those guys over there.” — He gestures to the corner.

“But,” I persist, “suppose we define consciousness as the factor that acts on quantum objects, making their behavior perceptible. I am sure that psychologists would take this possibility into account if you agreed with me. Let’s try to change our separatist worldview right now.” Now I am completely confident that Guernica’s chance of gaining consciousness depends on whether I can unite these people.

“To say that consciousness has a causal effect on atoms is to open Pandora’s box. This would turn objective physics on its head; physics would cease to be self-sufficient, and we would lose all confidence.” There is finality in the speaker’s voice. Someone else says in a voice I’ve heard before: “Nobody understands quantum mechanics.”

“But I promised Guernica that I would ask for his consciousness! Please listen to me.” I protest, but no one pays any attention. I became non-existent for this group – a non-consciousness, like Guernica.

I decide to try my luck with psychologists. I recognize them by the cluster of rat cages and computers in their corner.

A woman, looking knowledgeable, explains something to a young man. “By assuming that the brain-mind is a computer, we hope to break the vicious circle of behaviorism. The brain is the hardware of the computer. In reality, there is only the brain; this is what is real. However, brain hardware states perform independent functions over time, similar to computer software. It is these states of hardware that we call mind.”

“Then what is consciousness?” – the young man asks.

Oh, how timely. That’s exactly what I came here to find out – what psychologists think about consciousness! They must be the ones who control the consciousness of Guernica.

“Consciousness is like a central processing unit, the command center of a computer,” the woman answers patiently.

The questioner, not satisfied with this answer, energetically continues: “If we, at least in principle, can explain all the relationships of our inputs and outputs in terms of the activity of computer circuits, then consciousness seems absolutely unnecessary.”

I can’t help myself: “Please don’t refuse to acknowledge consciousness just yet. My friend Guernica needs it.” I tell them about the Guernica problem.

As if echoing my recent physicist acquaintance, the neatly dressed gentleman casually remarks: “But cognitive psychology is not yet ready for consciousness. We don’t even know how to define it.”

“I can tell you how a physicist defines consciousness. It has to do with quantum.”

That last word gets their attention. First I explain that quantum objects are waves that can exist in more than one place, and how consciousness can be a factor in focusing the waves so that we can observe them in one place. “And this is the solution to your problem,” I say. —You can take the definition of consciousness from physics! And then you may be able to help Guernica.”

“Aren’t you confused? Don’t physicists say that everything is made of atoms – quantum objects? If consciousness also consists of quantum objects, then how can it causally influence them? Think for yourself.”

I feel a slight panic. If these psychologists know what they are saying, then even my consciousness is an illusion, not to mention the consciousness of Guernica. But psychologists are right only if everything that exists, including consciousness, really consists of atoms. Suddenly another possibility occurs to me! And I blurt out: “You are completely wrong! You cannot be sure that everything in existence is made of atoms – this is just an assumption. Suppose instead that everything that exists, including atoms, is made of consciousness!”

My listeners seem stunned. “Look, there are some psychologists who think so. I admit it’s an interesting possibility. But it is not scientific. If we want to elevate psychology to the status of a science, we must eschew consciousness—and especially the idea that consciousness might be the ultimate reality. Sorry, buddy.” The woman who says this actually sounds quite sympathetic.

But I still have not moved towards Guernica consciousness. In desperation, I turn to the last group – the third vertex of the triangle. They happen to be neuroscientists (brain researchers). Perhaps their opinion actually means something.

Brain researchers are also debating consciousness, and my hopes are rising. “I argue that consciousness is the causal entity that gives meaning to existence,” says one of them, turning to an older and rather thin man. “But it must be an emergent phenomenon of the brain, and not separate from it.” After all, everything is made of matter; there is nothing besides her.”

The thin gentleman, speaking with an English accent, objects: “How can something made of something else cause a causal effect on the thing of which it is made? It’s the same as if television advertising is repeated, affecting the electronic circuits of the television. God forbid! No, to have a causal effect on the brain, consciousness must be a separate entity from it. It belongs to a separate world outside the material world.”

“But then how do these two worlds interact? The spirit cannot influence the machine.”

Rudely interrupting them, a third man with his hair tied in a ponytail laughs and says, “Both of you are talking nonsense. All your problems arise from trying to find meaning in an inherently meaningless material world. Look, physicists are right when they say that there is no meaning, no free will, and everything that exists is a random play of atoms.”

The English proponent of a separate world of consciousness responds sarcastically: “And you think that what you said makes sense? You yourself are a game of random, meaningless movement of atoms, yet you make up theories and think that your theories mean something.”

I’m interfering in the dispute. “I know how meaning can be made even in the play of atoms. Suppose that everything is not made of atoms, but of consciousness. What then?

“Where did you get this idea?” – they demand.

“From quantum physics,” I say.

“But there is no quantum physics at the macroscopic level of the brain,” they all say authoritatively, united in their objection. — Quantum physics — for the micro level, for atoms. Atoms form molecules, molecules form cells, and cells form the brain. We work with the brain every day; there is no need to invoke quantum mechanics of atoms to explain the macroscopic behavior of the brain.”

“But you don’t pretend to have a complete understanding of the brain, do you? The brain is not that simple! Didn’t someone say that if the brain were so simple that we could understand it, then we would be so simple that we wouldn’t be able to do it?”

“That may be true,” they concede, “but how does the idea of ​​quantum help us understand consciousness?”

I tell them that consciousness affects the quantum wave. “You see, it’s a paradox if consciousness consists of atoms. But if we reverse our idea of ​​what the world consists of, then this paradox is very satisfactorily resolved. I assure you, the world is made of consciousness.” I can’t hide my excitement and even pride – this is such a big idea. I ask them to join me.

“The sad thing,” I continue, “is that if ordinary people really knew that the link connecting them to each other and to the world is consciousness, and not matter, then their views on war and peace, environmental pollution, social justice, religious values ​​and all other human aspirations would undergo a radical change.”

“This sounds interesting, and believe me, I share your feelings. But your idea also sounds like something from the Bible. How can we accept religious ideas as science and continue to have credibility?” The questioner’s voice sounds as if he is talking to himself.

“I ask you to give consciousness its due,” I reply. “My friend Guernica needs consciousness to become whole again.” And from what I heard at this party, he’s not the only one. How can you still argue whether consciousness even exists? Surely the existence of consciousness cannot be disputed, and you know it.”

“I see,” says the man with the ponytail, shaking his head. – My friend, there was a misunderstanding. We are all chosen to be Guernica; you have to be if you want to do science. We have to assume that we are all made of atoms. Our consciousness has to be a secondary phenomenon – an epiphenomenon of the dance of atoms. This is required by the obligatory objectivity of science.”

I return to Guernica and tell him about my experience. “As Abraham Maslow once said, “If the only tool you have is a hammer, you start treating everything as if it were a nail.” These people are accustomed to perceiving the world as consisting of atoms and separate from themselves. They consider consciousness an illusory epiphenomenon. They can’t give you consciousness.”

“But what about you? — Guernica is looking at me intently. “Are you also going to hide behind scientific objectivity or are you going to do something to help me restore integrity?” Now he’s shaking me.

His persistence awakens me from my sleep. Gradually the desire to write this book is born.

* * *

Today we are faced with a great dilemma in physics. In quantum physics—the new physics—we have found a theoretical framework that works; it explains a myriad of laboratory experiments. Quantum physics has led to extremely useful technologies such as transistors, lasers and superconductors. And yet, we cannot understand the meaning of the mathematics of quantum physics without offering an interpretation of the experimental results, which many people can only look at as paradoxical and even impossible. Take a look at the following quantum properties:

• A quantum object (such as an electron) can be in more than one place at the same time (
wave property).

• A quantum object cannot be said to manifest itself in ordinary space-time reality until we observe it as a particle (
wave collapse).

• A quantum object ceases to exist here and simultaneously begins to exist somewhere else; however, we cannot say that he passed through the space separating these places
(quantum leap).

• The manifestation of a quantum object caused by our observation simultaneously affects its correlated counterpart object – no matter how far apart they are
(quantum action at a distance).

We cannot relate quantum physics to experimental data without using some kind of interpretive scheme, and interpretation depends on the philosophy we apply to the data. For centuries, science has been dominated by the philosophy of physical, or material, realism, which assumes that only matter—composed of atoms or, at most, elementary particles—is real; everything else is secondary phenomena of matter, simply the dance of the atoms that form it. This worldview is called realism because objects are assumed to be real and independent of the subjects—us—or how we observe them.

However, the idea that everything that exists is made of atoms is an untested assumption; it is not based on any direct evidence for all things. When new physics confronts us with a situation that appears paradoxical from the point of view of material realism, we tend to overlook the possibility that the paradoxes may arise because our untested assumption is wrong. (We tend to forget that a long-held assumption does not thereby become a fact, and we even become indignant when we are reminded of this.)

Today, many physicists suspect that something is wrong with material realism, but are afraid to rock the boat that has served them so well for so long. They do not realize that their boat is adrift and needs new navigation under the guidance of a new worldview.

Is there an alternative to the philosophy of material realism? Material realism, despite all its efforts and all its computer models, fails to explain the existence of our minds, especially the phenomenon of causally efficient self-awareness. “What is consciousness?” Material realism tries to shrug off this question by arrogantly answering that it doesn’t matter. However, if we take any seriously all the theories that the conscious mind constructs (including those that deny it), then consciousness still matters.

Ever since René Descartes divided reality into two distinct realms—mind and matter—many people have attempted to rationalize the causal efficacy of conscious minds within the framework of Cartesian dualism. However, science provides compelling reasons to doubt the validity of dualistic philosophy: for the worlds of mind and matter to interact, they must exchange energy, but we know that the energy of the material world remains constant. This means, of course, there is only one reality. It becomes a catch-22: if the only reality is material reality, then consciousness cannot exist except as an anomalous epiphenomenon.

So the question arises: is there an alternative to material realism, in which mind and matter are integral parts of one reality, but a reality that is not based on matter? I am convinced that there is. The alternative I propose in this book is monistic idealism. This philosophy is monistic, not dualistic, and it is idealism, since the main elements of reality are considered ideas (not to be confused with ideals) and their awareness, and matter is considered secondary. In other words, instead of asserting that everything (including consciousness) is made of atoms, this philosophy postulates that everything (including matter) exists in consciousness and is controlled from consciousness. Note that this philosophy does not say that matter is unreal, but only states that the reality of matter is secondary to the reality of consciousness, which itself is the basis of all that exists – including matter. In other words, in response to the question: “What is matter?”, a monistic idealist would never say, “It is immaterial.”

This book shows that the philosophy of monistic idealism provides a paradox-free and logically consistent satisfactory interpretation of quantum physics. Moreover, when the mind-body problem is reformulated in the general context of monistic idealism and quantum theory, mental phenomena such as self-awareness, free will, and even extrasensory perception are given simple and satisfactory explanations. This reformulated picture of the mind-brain allows us to understand ourselves entirely in a manner consistent with what the great spiritual traditions have taught us for thousands of years.

The negative impact of material realism on the quality of modern human life is simply amazing. Material realism depicts a universe devoid of any spiritual meaning: mechanical, empty and lonely. For us in the cosmos, this is perhaps all the more alarming because it has become alarmingly common that material realism has triumphed over theologies that posit a spiritual component of reality in addition to the material.

The facts prove otherwise; science shows the advantage of monistic philosophy over dualism – over spirit separated from matter. This book makes a strong case—supported by existing evidence—that the monist philosophy needed in the modern world is not materialism but idealism.

In idealistic philosophy consciousness is fundamental; therefore our spiritual experiences are recognized and affirmed as essential. This philosophy accommodates many of the interpretations of human spiritual experience that have animated various world religions. From this perspective, we see that some of the concepts of various religious traditions become as logical, elegant and satisfying as the interpretation of quantum physics experiments.

Know yourself. This advice has been given throughout the ages by philosophers who fully realized that it is our self that organizes the world and gives it meaning; their overarching goal was self-knowledge along with knowledge of nature. All this changed as a result of modern science’s acceptance of material realism; Instead of unity with nature, consciousness became separated from nature, which led to the separation of psychology from physics. As Morris Berman notes, this material-realist worldview banished us from the magical world in which we lived in the past and doomed us to a world alien to us. Now we live as exiles in this strange land; Who but the exiles would risk destroying this beautiful land through nuclear war and environmental pollution? This feeling of exile undermines our motivation to change our perspective. We are conditioned to believe that we are machines—that all our actions are determined by perceived stimuli and prior conditioning. As exiles, we have no responsibility and no choice; our free will is a mirage.

This is why it has become so important for each of us to carefully examine our worldview. Why am I threatened with nuclear destruction? Why does war continue to be a barbaric way to resolve world disputes? Why is there constant hunger in Africa when only in the US can we grow enough food to feed the world? How did I acquire a worldview (and, more importantly, was it imposed on me?) that dictates such separation between me and my fellow humans when we all share similar genetic, mental, and spiritual endowments? If I abandoned the outdated worldview based on material realism and explored the new/old worldview that quantum physics seems to require, could I be one with the world again?

We need to know more about ourselves; we need to know whether we can change our points of view – whether our mental constitution allows it. Can new physics and idealistic philosophy of mind give us new contexts for change?

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE

Several decades ago, American psychologist Abraham Maslow formulated the idea of ​​a hierarchy of needs. Once human beings have satisfied their basic survival needs, it becomes possible for them to strive to satisfy higher-level needs. According to Maslow, the highest of these needs is spiritual: the desire for self-fulfillment, to know oneself at the deepest possible level. Since many Americans, in fact many Westerners, have already moved beyond the lower rungs of Maslow’s ladder of needs, it would be expected that Westerners would enthusiastically climb to the upper rungs, moving toward self-actualization or spiritual fulfillment. However, this does not happen. What is wrong with Maslow’s arguments? As Mother Teresa noted during her visit to the United States in the eighties, Americans are rich materially, but impoverished spiritually. Why should this be so?

Maslow did not consider the consequences of the unquestioned materialism that dominates Western culture today. Most Westerners accept as scientific fact that we live in a materialistic world—a world where everything is made of matter and where matter is the fundamental reality. In such a world, material needs increase endlessly, leading not to the desire for spiritual growth, but to the desire to have more, bigger and better things: more powerful cars, better housing, the most fashionable clothes, amazing forms of entertainment and a dazzling extravaganza of present and future technological wonders. In such a world, our spiritual needs are often unrecognized, denied, or sublimated when they come to the surface. If only matter is real, as materialism has taught us to believe, then the only rational basis for happiness and well-being can only be material possessions.

Of course, our religions, our spiritual teachers and our artistic and literary traditions teach us that this is not so. On the contrary, they say that materialism leads, at best, to sickening satiety, and at worst to crime, disease and other troubles.

Most Westerners hold both of these opposing beliefs and live dual lives, embracing the predatory materialistic consumer culture while secretly despising it. Those of us who still consider ourselves religious cannot completely ignore the fact that although we adhere to religion in words and thoughts, our actions too often run counter to our intentions; we fail to convincingly embody even the most basic teachings of religions, such as treating our fellow humans kindly. Others overcome this cognitive dissonance by embracing religious fundamentalism or equally fundamentalist scientism.

In short, we are experiencing a crisis—not so much a crisis of faith as a crisis of confusion. How did we get to this unfortunate state? By accepting materialism as the so-called scientific view of the world. In our belief that we must be scientific, we are like the owner of the old gift shop in this story:

The buyer, having discovered a device unfamiliar to him, brings it to the owner and asks what it is intended for.
“Oh, it’s a barometer,” the owner replies. “It tells you whether it will rain.”
“How does it work?” – the buyer is surprised.
The owner doesn’t really know how the barometer works, but admitting it would mean missing out on the deal. So he says, “You put it out the window, and then you bring it back in. If the barometer is wet, you know it’s raining.”
“But I could do it with my bare hand, so why use a barometer?” – the buyer objects.
“That would be unscientific, my friend,” the owner replies.

I argue that in our acceptance of materialism we are like this shopkeeper. We want to be scientific; we consider ourselves scientific, but we are not. To be truly scientific, we must remember that science, while making new discoveries, has always changed. Is materialism a correct scientific worldview? I believe that it is possible to show the validity of a negative answer to this question, although the scientists themselves answer it vaguely.

Scientists’ confusion stems from a hangover from over-indulgence in the nearly four-century spree called “classical physics,” which was started by Isaac Newton around 1665. Newton’s theories set us on a course that would lead to the materialism that dominates Western culture. The worldview of classical physics, variously called material, physical, or scientific realism, corresponds to the philosophy of materialism, which dates back to the ancient Greek philosopher Democritus (r. 460-370 BC). Although classical physics has been formally replaced in this century by a new scientific discipline – quantum physics, the old philosophy of classical physics – the philosophy of materialism – is still generally accepted.

Classical physics and material realism

While visiting the Palace of Versailles, French mathematician and philosopher René Descartes was fascinated by the huge collection of automata in the palace gardens. As a result of the work of invisible mechanisms, fountains flowed, music played, sea nymphs frolicked, and the figure of the mighty Neptune rose from the depths of the pond. Observing this spectacle, Descartes came to the idea that perhaps the world is such an automaton – a world machine.

Later, Descartes put forward a significantly modified version of his picture of the world as a machine. His famous philosophy of dualism divided the world into the objective sphere of matter (the purview of science) and the subjective sphere of mind (the purview of religion). Thus, Descartes liberated scientific research from the orthodoxy of the powerful church. Descartes borrowed the idea of ​​objectivity from Aristotle. Its basic premise is that objects are independent and separate from the mind (or consciousness). We will call this the principle
of strict objectivity.

In addition, Descartes made contributions to the laws of physics that were to scientifically perpetuate his idea of ​​the world as a machine. However, it was Newton and his followers in the 18th century. firmly established materialism and its corollary, the principle of causal determinism, which states that any movement can be accurately predicted based on knowledge of the laws of motion and the initial conditions of objects (where they were and how fast they were moving).

To better understand the Newtonian-Cartesian view of the world, imagine the universe as a vast array of billiard balls—big and small—on a three-dimensional billiard table we call space. If we know all the forces acting on each of these balls at all times, then simply knowing their initial conditions – their positions and velocities at some initial moment in time – allows us to calculate where each of these bodies will be at any future moment in time ( or, for that matter, where they were at any point in time in the past).

The philosophical significance of determinism was best summarized by an 18th century mathematician. Pierre Simon Laplace: “An intellect that at any given moment became familiar with all the forces that set nature in motion, and with the state of the bodies of which it is composed, could – being extensive enough to subject the data to analysis – embrace in one and the same the same formula for the movement of the largest bodies in the universe and the movement of the lightest atoms; for such an intellect nothing would be uncertain, and the future, like the past, would be open to its gaze.”

In addition, Laplace wrote a successful book on celestial mechanics, which made him famous – so famous that Emperor Napoleon invited him to the palace.

“Mr. Laplace,” said Napoleon, “in your book you never mentioned God. Why?” (Custom at the time required that any serious book should contain several references to God, so Napoleon obviously had reason to be curious. What kind of a brave man was Laplace to break such a respectable custom?)
Laplace’s answer became a classic:
” Your Majesty, I did not need this hypothesis.”
Laplace correctly understood the significance of classical physics and its causal-deterministic mathematical structure. In Newton’s universe there is no need for God!

We have now become familiar with the two fundamental principles of classical physics: strict objectivity and determinism. The third principle of classical physics was discovered by Albert Einstein. Einstein’s theory of relativity—an extension of classical physics to bodies moving at high speeds—required that the highest speed in nature be the speed of light. This speed is enormous – 300,000 kilometers per second – but still limited. A consequence of this speed limit is that all interactions between material objects in spacetime must be local: they must be transmitted through space at a finite speed. This is called the principle
of locality.

When Descartes divided the world into matter and mind, he counted on a tacit agreement not to attack religion, which would have the highest authority where mind is concerned, in exchange for the supremacy of science in the material realm. This agreement was respected for more than two centuries. Eventually, the advances of science in predicting and controlling natural phenomena led scientists to question the validity of all religious teachings. In particular, scholars began to deny the mental, or spiritual, side of Cartesian dualism. Thus, the principle of materialistic monism was added to the list of postulates of material realism: everything in the world, including the mind and consciousness, consists of matter (and such generalizations of matter as energy and force fields). Our world is completely material.

Of course, no one yet knows how to derive mind and consciousness from matter, and therefore another mandatory postulate was added: the principle of
epiphenomenalism. According to this principle, all mental phenomena can be explained as epiphenomena, or secondary phenomena of matter, by appropriate reduction to antecedent physical conditions. The basic idea is this: what we call consciousness is simply a property (or group of properties) of the brain, if the brain is considered at a certain level.

Thus, these five principles constitute the philosophy of material (or materialist) realism:
1. Strict objectivity
2. Causal determinism
3. Locality
4. Physical or material monism
5. Epiphenomenalism

This philosophy is also called scientific realism, which implies that material realism is necessary for science. Most scientists, at least unconsciously, still believe this, despite well-established evidence that contradicts the five principles.

It is important to be aware from the very beginning that the principles of material realism are metaphysical postulates. These are assumptions about the nature of existence, not conclusions from experiments. If experimental data are found that contradict any of these postulates, then this postulate should be abandoned. Likewise, if rational evidence shows the weakness of a particular postulate, then the validity of that postulate should be questioned.

The main weakness of material realism is that its philosophy seems to completely exclude subjective phenomena. If we adhere to the postulate of strict objectivity, then many compelling experiments performed in the cognitive laboratory will be unacceptable as data. Material realists are fully aware of this shortcoming: thus, much attention has been given in recent years to the question of whether mental phenomena (including self-awareness) can be understood on the basis of material models—particularly computer models. We explore the basic idea behind such models: the idea of ​​machine intelligence.

Can we build a conscious computer?

The task of science after Newton, of course, was to approach as closely as possible the omniscient intellect of Laplace. The insight of Newton’s classical physics was quite impressive, and important steps were being taken towards this kind of approximation. Scientists have gradually unraveled, at least partially, some of the so-called eternal mysteries – how our planet came to be, where stars get their energy, how the universe was created and how life reproduces.

Eventually, Laplace’s followers took on the problem of explaining the human mind, self-awareness, and so on. With their deterministic insight, they had no doubt that the human mind is also a Newtonian classical machine, like the world machine of which it is a part.

One of the staunch supporters of the understanding of the mind as a machine, Ivan Pavlov, was very happy that his dogs confirmed this belief. When Pavlov rang the bell, his dogs salivated, even though they were not offered any food. Pavlov explained that dogs have developed a conditioned reflex to expect food whenever the bell rings. In reality, it’s quite simple. Give a stimulus, observe the reaction, and if it is the reaction you want, reinforce it with a reward.

Thus was born the idea that the human mind is a simple machine with simple inputs and outputs connected by one-to-one correspondence, which operates on the basis of stimulus-response-reinforcement. This idea has been much criticized on the grounds that such a simple behavioral machine could not carry out mental processes such as thinking.

You need thinking – you have it, answered the smart classical mechanists who came up with a complex machine with internal states. Look at how even a simple mobile behaves, they said. He is so fun to watch because his reactions to the wind patterns are endlessly varied. And why? Because each reaction, in addition to the specific stimulus, depends on many combinations of different internal states of the mobile branches. In the case of the brain, these states are synonymous with thinking, feeling, and so on, which are epiphenomena of the internal states of the complex machine that is the human brain.

Opposition voices continued to object: what about free will? Human beings have freedom of choice. The mechanists responded that free will was simply an illusion; they added the interesting argument that there was a possible physical model of illusory free will. The ingenuity of machine intelligence researchers is truly admirable. Now there is an idea that, although classical systems are ultimately deterministic and exhibit mostly deterministic behavior, chaos is also possible: at times very small changes in initial conditions can lead to very large differences in the final outcome for the system. This creates uncertainty (an example of this chaotic behavior is the uncertainty of weather systems), and predictive uncertainty can be interpreted as free will. Since chaos is ultimately deterministic, it follows that it is an illusion of free will. So, is our free will an illusion?

An even more convincing argument in favor of a mechanical picture of man was proposed by the English mathematician Alan Turing. In his opinion, someday we will construct a machine that obeys classical deterministic laws – a semiconductor computer that will be able to conduct a conversation with any person with so-called free will. Moreover, he proves that impartial observers will not be able to distinguish the conversation of a computer from the conversation of a human being. (I propose a motto for the new society: DRCHII – Movement for the equality of human and artificial intelligence.)

While I admire many of the advances in artificial intelligence, they do not convince me that my consciousness is an epiphenomenon and my free will a mirage. I do not believe that I am inherent in the limitations imposed on the classical machine by locality and causality. I do not believe that these are real limits for any human being, and I fear that understanding them as such may become a self-fulfilling prophecy.

“We are reflections of the world in which we live,” said historian of science Charles Singer. The question is how big can the reflection be? The sky is reflected in small ponds and in the mighty ocean. Which reflection is bigger?

But we have come a long way towards creating an intelligent Turing machine, argue proponents of machine intelligence. Our machines are already capable of passing the Turing Test on a random unsuspecting human. Undoubtedly, with further training and development, they will have a mind similar to that of a human. They will understand, learn and behave like us.

The machine mind proponent continues in deterministic fashion: If we can create Turing machines that behave like humans in all known respects, isn’t that proof that our own mind is just a collection of fully deterministic classical computer programs? Human unpredictability is no obstacle to such a view, since determinacy is not the same as predictability. In itself, this argument is convincing. If our computers can imitate human behavior, great, it will make communication between us and our machines easier. If we learn something about ourselves by studying computer programs that mimic some of our behaviors, that’s even better. However, it is a long way from simulating our behavior on computers to proving that we are made up of programs that perform the imitation.

Of course, even one example of a program we have that a classical computer can never imitate would destroy the myth of the mind as a machine. Mathematician Roger Penrose shows that computer-like algorithmic reasoning is not enough to discover mathematical theorems and laws. (An algorithm is a sequential procedure for solving a problem: a strictly logical, rules-based approach.) So, Penrose asks, where does mathematics come from if we act like computers? “Mathematical truth is
not something we establish simply by means of an algorithm. I also believe that the decisive element in our comprehension of mathematical truth is
consciousness. We must “see” the truth of a mathematical proof in order to be convinced of its correctness. This “seeing” is the very essence of consciousness. It must be present
whenever we directly perceive mathematical truth.” In other words, our consciousness must exist before our algorithmic computer ability.

An even stronger argument against the idea of ​​the mind as a machine was put forward by Nobel laureate physicist Richard Feynman. A classical computer, Feynman notes, will never be able to simulate nonlocality (a technical term for the transfer of information or influence without local signals; such transfer is an action at a distance and occurs instantaneously). Therefore, if non-local information processing exists in humans, then this is one of our non-algorithmic programs that a classical computer will never be able to imitate.

Do we have non-local information processing? A very compelling argument for non-locality can be found in our spirituality. Another controversial argument for nonlocality comes from claims of paranormal experience. For centuries, people have claimed to have the ability to telepathy – transferring information from mind to mind without local signals, and now there seems to be some scientific evidence for this.

Alan Turing himself understood that telepathy provided one reliable way to distinguish a person from a computer in the Turing Test: “Let us conduct a simulation experiment using a telepathic person and a machine as a witness. An expert may ask, for example, questions such as: “What suit is the card I have in my hand?” A person, using telepathy or clairvoyance, correctly guesses 130 out of 400 cards. The machine can only give random guesses, and will probably get 104 correct answers, and thus the expert will make the correct determination.”

Extrasensory perception (ESP), while undoubtedly controversial, is just one argument against the power of the classical computer. Another important ability of the human mind that seems unattainable by a computer is creativity. If creativity involves discontinuity, abrupt departures from past patterns of thought, then the computer’s ability to be creative is certainly questionable, since the classical computer operates with continuity.

However, at the end of the day, it’s all about consciousness. If machine intelligence experts can create a classical computer that is conscious in the same sense that you and I are conscious, things will take a different turn, despite all the secondary considerations listed above. Will they be able to? Who knows. Suppose we provided a Turing machine with a myriad of programs that perfectly imitated our behavior; Would the machine then become conscious? Of course, its behavior would demonstrate all the complexities of the human mind, and, like a Turing machine, it would be a flawless imitation of a person (with the exception of a few specifically human characteristics, such as ESP and the ability to be mathematically creative, which are considered dubious by machine intelligence enthusiasts anyway) , but would it be truly conscious?

While in college in the 1950s. I was introduced to the idea of ​​a conscious computer by reading Robert Hanlein’s science fiction novel The Moon is a Harsh Mistress. Hanlein conveys the idea that a computer’s consciousness depends on its size and complexity: once the machine in the novel passes the limit of complexity and size, it becomes conscious. Apparently, this view is widespread among artificial intelligence researchers.

It seems to me that computer consciousness is not defined by complexity. Of course, a high level of complexity may ensure that a computer’s responses to a given stimulus are no more predictable than those of humans, but no more than that. If we can trace the input/output characteristics of a computer to the activity of its internal circuits quite unambiguously (and this, at least in principle, should always be possible for a classical computer), then why do we need consciousness? Apparently it won’t have any function. I think proponents of artificial intelligence dodge the question by saying that consciousness is just an epiphenomenon, or an illusion. Apparently, Nobel laureate neurophysiologist John Eccles agrees with my point of view. He asks, “Why should we be conscious at all? In principle, we can explain all our input/output characteristics in terms of the activity of neural circuits; and therefore consciousness seems absolutely superfluous.”

Not everything that is superfluous in nature is prohibited, but it is unlikely. For a classical Turing machine, consciousness seems superfluous, and this is reason enough to doubt that these machines, no matter how complex, will ever be conscious. The fact that we ourselves have consciousness only suggests that our input/output characteristics are not completely determined by the algorithmic programs of classical computer machinery.

Proponents of machine intelligence sometimes make another argument: we freely attribute consciousness to other human beings because they report mental experiences—thoughts, feelings, etc.—that are similar to our own. If we program an android to report thoughts and feelings similar to our own, yours, could you distinguish his consciousness from the consciousness of your friend? After all, you are no more capable of experiencing what’s going on in your friend’s head than what’s going on in the head of an android. So you can never know for sure anyway!

This reminds me of an episode of the TV series Star Trek. The fraudster is sentenced to an unusual punishment, which does not look like a punishment at all. He is exiled to a colony where he will be the only person surrounded by androids serving him, many of which are in the form of beautiful girls.

You, like me, can guess why this is a punishment. The reason I do not live in a solipsistic universe (where only I am real) is not because others like me logically convince me of their humanity, but because I have an inner connection with them. I could never have such a connection with an android.

I argue that the sense of inner connection we have with other people is due to a real connection of the spirit. I believe that classical computers could never be conscious like us because they do not have this spiritual connection.

Etymologically , the word
consciousness comes from the words scire (to know) and sit (with). Consciousness means “to know with.” To me this term implies non-local knowledge; we cannot “know with someone” without having a non-local connection with that person.

We should not be dismayed that we cannot build a model of ourselves based on classical physics and using a computer algorithmic approach. Since the beginning of this century we have known that classical physics is incomplete. Not surprisingly, it gives us an incomplete worldview. Let us examine the new physics born at the dawn of the 20th century, and, from the point of view of the end of this century, see what freedom its worldview brings with it.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM

Almost a century ago, a number of experimental discoveries were made in physics that required a change in our worldview. What these experiments revealed were, in the words of philosopher Thomas Kuhn, anomalies that classical physics could not explain. These anomalies opened the way to a revolution in scientific thought.

Imagine that you are a physicist on the threshold of a new century. One of the anomalies you and your colleagues want to understand concerns how heated bodies emit radiation. As a Newtonian physicist, you believe that the universe is a classical machine made up of parts that behave according to the laws of Newtonian mechanics, which are almost all completely known. You believe that with all the information about the parts and the few remaining difficulties regarding the laws, you will be able to predict the future of the universe forever. However, these few remaining difficulties are unpleasant. You are not ready to answer questions concerning, for example, what is the law of radiation of heated bodies.

Imagine that while you are racking your brains over this question, your wife is sitting comfortably next to you in front of a burning fireplace.
You (mumbling): I just can’t understand it.

She : Pass me the nuts.
You (passing nuts ): I just can’t understand why we’re not sunbathing now.
She (laughing): Well, that would be nice. We might even have reason to use the fireplace in the summer.
You: You see, the theory says that the radiation from the fireplace should be as rich in ultraviolet light as sunlight. But what is it that makes sunlight, and not fireplace light, rich in these high frequencies? Why don’t we tan now by taking an ultraviolet bath?
She: Wait, please. For me to listen to this seriously, you’ll have to slow down a little and explain. What is frequency? What is ultraviolet?
You: Sorry. Frequency is the number of cycles per second. This is a measure of how quickly a wave oscillates. For light, this means color. White light is made up of light of different frequencies, or colors. Red is low frequency light and violet is high frequency light. If the frequency is even higher, then it is an invisible black color, which we call ultraviolet.
She: Okay, so both the light from burning wood and the light from the sun must contain a lot of ultraviolet radiation. Unfortunately, the sun obeys your theory, but burning wood does not. Perhaps there is something special about burning wood…
You: Actually, it’s even worse. All light sources, not just the sun or burning wood, should produce large amounts of ultraviolet radiation.
She: Ah, this is getting interesting. Ultraviolet inflation is omnipresent. But isn’t every inflation followed by a recession? Doesn’t the song say that everything that rises must fall? ( She starts humming without words.)
You (annoyed ): But how?
She ( holding out a bowl of nuts): Do you want some nuts, dear?
(The conversation ends.)

Planck makes his first quantum leap

At the end of the 19th century. Many physicists were disappointed until one of them broke the general trend – it was Max Planck from Germany. In 1900, Planck made a bold conceptual breakthrough when he declared that the old theory needed a quantum leap (he borrowed the word
quantum, meaning “quantity,” from Latin). The emission of light from hot objects—such as burning wood or the sun—is caused by electrons, tiny oscillating electrical charges. These electrons absorb energy from a heated environment, such as a fireplace, and then release it back as radiation. This part of the old physics was correct, but then classical physics predicted that the emitted radiation should be rich in ultraviolet, which our observations contradicted. Planck (very bravely) announced that the problem of emitting different amounts of ultraviolet light could be solved by assuming that electrons emit or absorb energy only in certain discrete portions, which he called “quanta” of energy.

To understand the meaning of an energy quantum, consider this analogy. Compare the case of a ball rolling down a staircase with the case of a ball rolling down an inclined plane (Fig. 1), which can occupy any position on the inclined plane, and its position can change by any amount. So this is a continuity model that represents the way we think in classical physics. By contrast, a ball on a staircase can only be on one step or another; its position (and its energy, which is associated with the position) is “quantized”.

Img. 1.
Quantum leap. On an inclined plane, the classical motion of a ball is continuous; on the quantum ladder, movement occurs in the form of discrete stages (quantum leaps)

You may object – what happens when the ball falls from one step to another? Doesn’t he occupy intermediate positions during his descent? This is where the unusualness of quantum theory comes into play. For a ball on a ladder, the answer must obviously be positive, but for the case of a quantum ball (atom or electron), Planck’s theory gives a negative answer. A quantum ball can never be found in any intermediate position between two steps; he is either on one or the other. This is quantum discontinuity.

So why can’t you get a tan from a wood burning fireplace? Imagine a pendulum in the wind Usually in such a situation the pendulum will swing, even if the wind is not very strong. Suppose, however, that the pendulum can absorb energy only in discrete portions of large magnitude. In other words, it is a quantum pendulum. What then? It is clear that unless the wind is capable of producing the required high increase in energy in one step, the pendulum will not move. Absorbing small amounts of energy will not allow it to accumulate enough energy to overcome the threshold. So it is with oscillating electrons in a fireplace. Small quantum jumps produce low-frequency radiation, but high-frequency radiation requires large quantum jumps. A large quantum jump must be caused by a large amount of energy in the environment surrounding the electron; The energy of wood burning in a fireplace is simply not strong enough to create the conditions to produce large amounts of blue light, let alone ultraviolet light. This is the reason why you can’t get a tan while sitting by the fireplace.

As far as we know, Planck was a fairly traditional scientist and was reluctant to make his ideas about energy quanta public. He even did his mathematics while standing, as was customary in Germany at that time. He didn’t particularly like the consequences of his innovative idea; however, it was becoming clear to the scientists who were to take the revolution much further that they were pointing to an entirely new way of understanding our physical reality.

Einstein’s photons and Bohr’s atom

One of these revolutionaries was Albert Einstein. At the time he published his first research paper on quantum theory, he was working as a clerk in the patent office in Zurich (1900). Questioning the then-popular idea of ​​the wave nature of light, Einstein hypothesized that light exists in the idea of ​​a quantum—a discrete bundle of energy—that we now call a photon. The higher the frequency of light, the more energy each beam has.

An even greater revolutionary was the Danish physicist Niels Bohr, who in 1913 used the idea of ​​the quantum of light to formulate the hypothesis that the entire atomic world is full of quantum leaps. We have all been taught that the atom is like a miniature solar system, that electrons orbit the nucleus much like the planets orbit the sun. You may be interested to know that this model, proposed by the English physicist Ernest Rutherford, had a crucial flaw that Bohr’s work corrected.

Imagine a swarm of orbiting satellites that are launched quite regularly from Earth using space rockets. These satellites do not exist forever. Due to collision with the earth’s atmosphere, they lose energy and slow down their movement. Their orbits narrow and they eventually fall to Earth (Fig. 2).

Img. 2.
The orbits of satellites revolving around the Earth are unstable. The orbits of electrons behave in the same way in the Rutherford model of the atom.

According to classical physics, the electrons surrounding the atomic nucleus would also lose energy due to the continuous emission of light and eventually fall into the nucleus. Therefore, the planetary model of the atom is unstable. However, Bohr (who supposedly saw the planetary system of the atom in a dream) created a stable model of the atom using the idea of ​​a quantum leap.

Suppose, Bohr said, that the orbits of electrons are discrete, like Planck’s energy quanta. The orbits can then be thought of as forming an energy ladder (Fig. 3). They are stationary – the amount of their energy remains unchanged. While in these quantized orbits, electrons do not emit light. An electron emits a quantum of light only when it jumps from a higher energy orbit to a lower orbit (from a higher energy rung of the ladder to a lower rung). Thus, if an electron is in the lowest energy orbit, it has no lower level to jump to. This base level configuration is stable and the electron has no chance of falling into the nucleus. All physicists greeted Bohr’s atomic model with a sigh of relief.

Img. 3.
Bohr’s orbit and quantum leap: a – quantized Bohr orbits. Atoms emit light as electrons jump from orbit to orbit; b – for quantum leaps along the energy ladder there is no need to pass through the intermediate space between steps

Bohr cut off the head of the Hydra of Instability, but another grew in its place. According to Bohr, an electron can never occupy any position between orbits; thus, when it makes a jump, it must somehow directly transfer to another orbit. This is not an orbital jump through space, but something radically new. Although it might be tempting to picture an electron’s jump as jumping from one rung of a ladder to another, the electron makes the jump without crossing the space between the rungs. Instead, it seems to disappear on one step, reappearing on another – without any continuous transition. Moreover, it is impossible to say where he is going to jump if there is more than one lower step between which he can choose. Only probabilistic predictions can be made.

Wave-particle duality

You may have noticed something strange about the quantum concept of light. To say that light exists in the form of quanta, photons, is to say that light consists of particles like grains of sand. However, such a statement largely contradicts the everyday experience that we gain when dealing with light.

Imagine, for example, looking at a distant street lamp through the fabric of a cloth umbrella. You won’t see a continuous stream of light passing through, as you would expect if the light was made up of tiny particles (put sand in a sieve and you’ll see what I mean). Instead, you’ll see a pattern of alternating dark and light borders, which is technically called an interference pattern. Light bends in and around the threads of fabric, creating a pattern that only waves can create. Thus, even our everyday experience shows that light behaves like a wave.

However, quantum theory insists that light also behaves as a beam of particles, or photons. Our eyes are such a wonderful instrument that we can observe for ourselves the quantum, grainy nature of light. The next time you part with a loved one at dusk, pay attention to how you see a retreating figure. Notice that the outline of the receding object appears fragmented. If the light energy bouncing off that object and hitting the optical receptors on your retina had wave-like continuity, then at least some light from every part of the object should always excite your optical receptors. You would always see the full image. (Admittedly, in low light, the contrast between dark and light would not be very clear, but this would not affect the clarity of the outlines.) However, what you see instead is not clear outlines at all, because the receptors in your eyes respond to individual photons. Dim light has fewer photons than bright light; so in this hypothetical twilight situation, only a few of your receptors will be stimulated at any given time—too few to detect the outline of a dimly lit figure. Therefore, the image you see will be fragmented.

Another question you may be wondering is: why can’t receptors store data indefinitely until the brain has collected enough information to put all the fragmented pictures together? Fortunately for quantum physicists, who are always desperate for everyday examples of quantum phenomena, optical receptors can only store information for fractions of a second. In dim light, at any given moment, not enough receptors in your eyes will be stimulated to create a complete image. The next time you say goodbye to the shadowy, retreating figure of a loved one at dusk, remember to think about the quantum nature of light; this will surely reduce the pain of your separation.

When light is considered as a wave, it is capable of being in two (or more places) at the same time – as is the case when it passes through the holes in the fabric of an umbrella, and forms a diffraction pattern; however, when we capture it on photographic film, it appears discretely, in separate specks, like a stream of particles. Thus light must be both a wave and a particle. Paradoxical, isn’t it? The case concerns one of the bastions of old physics: unambiguous description in natural language. Moreover, the very idea of ​​objectivity is at stake: does the nature of light—what light is—depend on how we observe it?

And as if the paradoxes surrounding light weren’t challenging enough, another question inevitably arises: can a material object, such as an electron, be both a wave and a particle? Can he have a duality like the duality of light? The physicist who first posed this question and persistently gave a positive answer to it, which shocked all his colleagues, was the French aristocrat Louis Victor de Broglie.

Waves of matter

When de Broglie wrote his PhD thesis around 1924, he drew a parallel between the discreteness of the stationary orbits of the Bohr atom and the discreteness of sound waves produced by a guitar. The parallel turned out to be fruitful.

Imagine the movement of a sound wave in some medium (Fig. 4). The vertical displacement of the particles of the medium changes from zero to a maximum (ridge), back to zero, to a negative maximum (trough), again to zero, and so on with increasing distance. The maximum vertical displacement in one direction (from zero to a crest or trough) is called amplitude. Individual particles of the medium move back and forth relative to their resting position. However, a wave passing through a medium spreads. A wave is a propagating disturbance. The number of crests passing through a given point per second is called the wave frequency, and the distance from crest to crest is called the wavelength.

Img. 4.
Graphical representation of the wave

Plucking a guitar string sets it in motion, but the resulting vibrations are called stationary (standing waves) because they do not propagate beyond the string. At any given location on the string, the displacement of the string particles changes over time: waviness occurs, but the waves do not propagate in space (Fig. 5). The propagating waves that we hear are driven by standing waves of vibrating strings.

Img. 5.
The first few harmonics of a stationary, or standing, wave in a guitar string

A musical note on a guitar consists of a range of sounds—a spectrum of frequencies. De Broglie was interested in the fact that the standing waves of a guitar string create a discrete spectrum of frequencies called harmonics. The lowest frequency sound is called the first harmonic, which determines the tone we hear. Higher harmonics—the musical sounds that give a note its characteristic quality—have frequencies that are multiples of the first harmonic.

Stationarity is a property of waves in a confined space. Such waves can easily be produced in a cup of tea. De Broglie asked whether the electrons of an atom are localized (confined) waves? If so, do they form discrete stationary wave patterns? For example, perhaps the lowest atomic orbit is the one in which a single electron produces a stationary wave of the lowest frequency—the first harmonic—and higher orbits correspond to stationary electron waves of higher harmonics (Figure 6).

Img. 6.
De Broglie’s idea: couldn’t electrons be stationary waves in the limited space of an atom?

Of course, de Broglie used much more complex arguments to support his idea, but he still had difficulty getting his dissertation approved. Eventually it was sent to Einstein for review. Einstein, who first recognized the dual nature of light, had no difficulty in realizing that de Broglie might well be right: matter could well be as dual as light. De Broglie was awarded his degree when Einstein commented on his thesis: “It may look crazy, but it is actually logical.”

In science, experiment is always the final arbiter. The correctness of de Broglie’s idea about the wave nature of the electron was brilliantly demonstrated by an experiment in which a beam of electrons was passed through a crystal (a three-dimensional “umbrella” suitable for electron diffraction) and photographed. The result was a diffraction pattern (Fig. 7).

Img. 7.
Concentric diffraction rings show the wave nature of electrons

If matter is a wave, one physicist quipped to another at the end of a 1926 seminar on de Broglie waves, then there must be a wave equation that describes the wave of matter. The physicist who made this remark immediately forgot about it, but the one who heard it, Erwin Schrödinger, went on to discover the wave equation for matter, now known as the Schrödinger equation. It is the cornerstone that replaced Newton’s laws in the new physics. The Schrödinger equation is used to predict all the amazing qualities of submicroscopic objects found in our laboratory experiments. Werner Heisenberg discovered this same equation even earlier, but in a less clear mathematical form. The mathematical formalism that grew out of the work of Schrödinger and Heisenberg is called quantum mechanics.

The idea of ​​matter waves proposed by de Broglie and Schrödinger gives rise to an amazing picture of the atom. It explains in simple terms the three most important properties of atoms: their stability, their identity with each other, and their ability to regenerate. I have already explained how stability arises – this was Bohr’s great contribution. The identity of atoms of a certain kind is simply a consequence of the identity of wave patterns in a limited space; the structure of stationary patterns is determined by the way in which the movement of electrons is limited, and not by their environment. The music of the atom, its wave pattern, remains the same no matter where it is – on Earth or in the Andromeda nebula. Moreover, a stationary pattern, depending only on the conditions of its limitation, has no trace of past history, no memory; it is restored again and again in the same form.

Waves of Probability

Electron waves are not like ordinary waves. Even in a diffraction experiment, individual electrons appear on a photographic plate as localized individual events; It is only by observing the pattern created by the entire beam of electrons that we discover evidence of their wave nature—the diffraction pattern. Electron waves are waves of probability, said physicist Max Born. They give us probabilities: for example, we are very likely to find a particle where the wave disturbances (or amplitudes) are large. If the probability of finding a particle is small, the amplitude of the wave will be weak. Imagine watching traffic from a helicopter hovering over the streets of Los Angeles. If cars were described by Schrödinger’s equation, we would say that the wave is strong in areas of traffic jams, and the wave is weak between traffic jams.

In addition, electron waves are usually represented as
wave packets . Using the concept of packets, we can make the wave amplitude large in certain regions of space, and small in all other places (Fig. 8). This is important because the wave must represent a localized particle. A wave packet is a packet of probability, and Born argued that for electron waves, the square of the wave amplitude—technically called the wave function—at some point in space gives us the probability of finding an electron at that point. This probability can be represented by a bell-shaped curve (Fig. 9).

Img. 8.
The superposition of many simple waves forms a typical local wave packet (From P. W. Atkins’s book Quanta: A Handbook of Concepts, Oxford: Clairdon Press, 1974)

Img. 9.
Typical probability distribution

Heisenberg Uncertainty Principle

Probability breeds uncertainty. For an electron or any other quantum object, we can only talk about the probability of its being in such and such a place, or that its momentum (mass times velocity) is equal to such and such, but these probabilities form a distribution described by a bell-shaped curve. The probability will be maximum for some position value, and this will be the most likely location of the electron. However, there will be a whole range of positions in which there is a significant chance of finding an electron. The width of this region corresponds to the uncertainty of the electron position. The same arguments allow us to talk about the uncertainty of the electron momentum.

Based on similar considerations, Heisenberg mathematically proved that the product of the uncertainties of the electron’s position and momentum is greater than or equal to a certain small number called Planck’s constant. This number, originally discovered by Planck, sets the quantitative scale at which quantum effects become usably large. If Planck’s constant were not so small, the effects of quantum uncertainty would even invade our everyday macroscopic reality.

In classical physics, any movement is determined by the forces that control it. Once we know the initial conditions (the position and momentum of an object at some initial time), we can calculate its exact trajectory using Newton’s equations of motion. Therefore, classical physics leads to the philosophy of determinism – the idea of ​​​​the possibility of completely predicting the movement of all material objects.

The principle of uncertainty undermines the philosophy of determinism. According to the uncertainty principle, we cannot accurately determine the position and speed (or momentum) of an electron at the same time; any attempt to accurately measure one makes knowledge of the other uncertain. Therefore, the initial conditions for calculating a particle’s trajectory can never be precisely determined, and the concept of a well-defined particle trajectory becomes unusable.

For the same reason, Bohr’s orbits do not provide a strict description of the location of the electron: the position of the actual orbits is uncertain. We really cannot say that an electron located at one energy level or another is located at such and such a distance from the nucleus.

Dubious fantasies

Let’s consider several fantastic scenarios, the authors of which did not realize the importance of the uncertainty principle or forgot about it.

In the science fiction book Fantastic Voyage and the movie based on it, objects were made miniature by compaction. Have you ever wondered if it is possible to compress atoms? After all, they are mostly made up of empty space. Is this possible? Decide for yourself based on the uncertainty principle. The size of an atom gives a rough idea of ​​the degree of uncertainty in the position of its electrons. Densifying an atom will place its electrons in a smaller volume of space, thereby reducing the uncertainty of their position; but the uncertainty of their momentum must increase. Increasing the uncertainty of the electron’s momentum means increasing its speed. Thus, as a result of compaction, the speed of electrons increases and they are more able to leave the atom.

In another example of science fiction, Captain Kirk (from the classic television series Star Trek) gives the command: launch! A button is pressed on the dashboard and whoops, the people standing on the platform disappear, appearing at a destination that is supposed to be an unexplored planet but looks a lot like a Hollywood set. In one of his novels based on the Star Trek series, James Blish tried to characterize this process as a quantum leap. Just as an electron jumps from one atomic orbit to another without crossing the space in between, so would the crew of the starship Enterprise. You can see what the problem is here. When and where the electron will make the leap is not subject to the law of causality and is unpredictable due to the laws of probability and uncertainty of the quantum leap. Such a quantum transport would force the heroes of the Enterprise, at least sometimes, to wait a very long time to get somewhere.

Quantum fantasies may be fun, but the ultimate goal of new physics and this book is serious. It is to help us deal with our daily reality.

Wave-particle duality and quantum measurement

The preceding background information helps explain a couple of puzzling questions. Does the quantum picture of an electron moving in waves around a nucleus imply that the electron’s charge and mass are spread throughout the atom? And does the fact that a free electron propagates as a wave should propagate according to Schrödinger’s theory mean that its charge is now spread throughout space? In other words, how can the wave pattern of an electron be reconciled with the fact that it has the properties of a localized particle? The answers to these questions are quite complex.

It would seem that at least wave packets make it possible to confine an electron to a small space. Alas, everything is not so simple. A wave packet that satisfies the Schrödinger equation at a given time must spread over time.

At some initial moment we can localize the electron to a tiny point, but within seconds the wave packet of the electron will spread throughout the entire city. Although initially the probability of finding an electron in a tiny spot is overwhelmingly high, after just a few seconds the probability of an electron appearing anywhere in the city becomes significant. And if we wait long enough, the electron could appear anywhere in the entire country, or even the entire universe.

It is this propagation of the wave packet that contributes to the ongoing jokes about quantum predestination among cognoscenti. For example, take this quantum mechanical way of materializing a Christmas turkey: prepare the oven and wait – there is a non-zero probability that a turkey from a nearby store will materialize in your oven.

Unfortunately for the turkey lover, for objects as massive as the turkey, propagation is extremely slow. To materialize even a small piece of turkey in this way might have to wait for the entire existence of the universe.

What about the electron? How to reconcile the propagation of an electron wave packet throughout the city with the picture of a localized particle? The answer is that we must take the act of observation into account in our calculations.

If we want to measure the charge on an electron, we must trap it in something like a cloud of vapor in a condensation chamber. As a result of this measurement, we must assume that the electron wave collapses, so that we are now able to see the path of the electron through the vapor cloud (Fig. 10). According to Heisenberg, “the path of an electron only comes into existence when we observe it.” When we make a measurement, we always find an electron localized as a particle. We can say that our measurement reduces the electron wave to the particle state.

Img. 10.
Electron track in a vapor cloud

When Schrödinger proposed his wave equation, he and others thought that they might have freed physics from quantum jumps—from discontinuity—since wave motion is continuous. However, the corpuscular nature of quantum objects had to be reconciled with their wave nature. Therefore, wave packets were proposed. Finally, with the recognition of the propagation of the wave packet and the realization that it is the measurement that must cause the instantaneous collapse of the packet dimensions, we see that the collapse must be intermittent (continuous collapse would take time).

It seems as if there can be no quantum mechanics without quantum leaps. Schrödinger once visited Bohr in Copenhagen, where he spent days protesting against quantum leaps. It is said that he finally gave up, exclaiming in irritation: “If I had known that it was necessary to recognize this damned quantum leap, I would never have gotten involved with quantum mechanics.”

Let’s go back to the atom: if we measure the position of an electron in an atomic stationary state, we again collapse its cloud of probability, finding it in a certain position, and not smeared all over the place. By making a large number of measurements in search of an electron, we will more often find it in places where the probability of finding it is high, in accordance with the prediction of the Schrödinger equation. Indeed, if after a large number of measurements we plot the measured positions, it will look exactly like the blurred orbital distribution that the solution of the Schrödinger equation gives (Fig. 11).

Img. 11.
Results of multiple measurements of the position of the electron in the hydrogen atom in the lowest orbit. It is obvious that the electron wave usually collapses where the predicted probability of finding it is high, resulting in a fuzzy orbit

What does a flying electron look like from this point of view? When we make an initial observation of any propagating submicroscopic object, we find it localized as a particle in a tiny wave packet. However, after observation, the packet dissipates, and the dispersion of the packet represents the cloud of our uncertainty about the packet. If we observe again, the packet is localized again, but between our observations it always dissipates.

According to physicist-philosopher Henry Margenau, seeing electrons is like seeing fireflies on a summer evening. You may see a flash here, and a flicker of light there, but have no idea where the firefly is between your sightings. You cannot determine its trajectory with any certainty. Even for such a macroscopic object as the moon, quantum mechanics predicts essentially the same picture – the only difference is that the scattering of the wave packet is immeasurably small (but non-zero between observations).

Now we come to the heart of the matter. Whenever we measure a quantum object, it appears in one place as a particle. The probability distribution simply identifies the place (or places) where it is likely to be found when we measure it – nothing more. When we don’t measure it, a quantum object scatters and exists at the same time in more than one place, just like a wave or a cloud – no less.

Quantum physics offers a new and exciting worldview that challenges old concepts such as deterministic trajectories and causal continuity. If initial conditions do not forever determine the movement of an object, if every moment we observe becomes a new beginning, then at a fundamental level the world is creative.

There was a Cossack who saw how almost every day, at approximately the same time, a rabbi crossed the city square. One day, out of curiosity, he asked, “Where are you going, Rabbi?”
The rabbi answered: “I don’t know for sure.”
“You pass this road every day at this time. Of course you know where you are going.”
When the rabbi insisted that he did not know this, the Cossack became angry, then became suspicious and finally took the rabbi to prison. Just as he was locking the cell door, the rabbi looked at him and said softly, “See, I didn’t know.”
Before the Cossack stopped him, the rabbi knew where he was going, but after that he no longer knew. Stopping (we can call it measuring) opened up new possibilities. This is the meaning of quantum mechanics. The world is not determined once and for all by initial conditions. Every measurement event is potentially creative and can open up new possibilities.

The principle of complementarity

A new way of understanding the paradox of wave-particle duality was proposed by Bohr. According to him, the wave and particle nature of the electron are not dual, but simply polar opposite qualities. These are complementary qualities that are revealed to us in complementary experiments. When we take the diffraction pattern of an electron, we discover its wave nature; when we trace it in a cloud of vapor, we see its corpuscular nature. Electrons are neither waves nor particles. They can be called “wave particles” because their true nature transcends both descriptions. This is the principle of complementarity.

Since contemplating the fact that the same quantum object has such seemingly contradictory properties as wave and particle can be dangerous for the human psyche, nature has provided a shock absorber. Bohr’s principle of complementarity assures us that although quantum objects have both wave and particle properties, we can, in any experimental setting, measure only one aspect of a wave particle at any given time. We choose which aspect of the wave particle we want to see by choosing the appropriate experimental setting.

Principle of correspondence

Having understood the revolutionary ideas of the new physics, it would be completely wrong to think that Newton’s physics was completely wrong. The old physics lives on in the realm of most (but not all) gross matter as a special case of the new physics. An important feature of science is that when a new order replaces an old one, it usually expands the scope of its application. In the old region, the mathematical equations of old physics remain valid (confirmed by experimental data). Therefore, in the realm of classical physics, the conclusions of quantum physics about the motion of objects correspond clearly to those made using Newtonian mathematics under the assumption that the bodies with which we are dealing are classical. Bohr formulated this principle of correspondence. The relationship between classical and quantum physics is in some ways like a visual illusion. “My wife and my mother-in-law” (Fig. 12). What do you see in this picture? First you see either the wife or the mother-in-law. I always see my wife first. In fact, it may take you a while to spot the second image in the drawing. If you look closely at it, suddenly a second image appears. The wife’s chin becomes the mother-in-law’s nose, her neck becomes the chin of an older woman, and so on. You may be wondering – what’s going on? The lines of the drawing remain the same, but suddenly a new way of perceiving the painting becomes possible for you. Very soon you discover that you can easily move from the old picture to the new one and back again. At any moment you still see only one of the two images, but your consciousness has expanded so that you are aware of their duality. In this expanded state of awareness, the strangeness of quantum physics begins to become clear. It even becomes exciting. To paraphrase Hamlet’s words addressed to Horatio, we can say that there are many things in heaven and on earth that classical physics never dreamed of.

Img. 12.
My wife and my mother-in-law

Quantum mechanics gives us a broader perspective, a new context that expands our perception into a new area. We can see nature as separate forms – waves or particles – or we can detect complementarity: the idea that the same thing has both wave and particle properties.

Copenhagen interpretation

According to the so-called Copenhagen interpretation of quantum mechanics, developed by Bohr, Heisenberg and Born, we calculate quantum objects as waves and interpret the waves in a probabilistic manner. We define their attributes, such as position and momentum, somewhat vaguely and understand them in terms of the principle of complementarity. In addition, lack of continuity and quantum jumps—for example, the collapse of an expanding wave packet upon observation—are considered fundamental aspects of the behavior of a quantum object. Another aspect of quantum mechanics is inseparability. Talking about a quantum object without talking about how we observe it makes no sense, since one is inseparable from the other. Finally, for massive macroobjects, the predictions of quantum mechanics coincide with the predictions of classical physics. This prohibits the manifestation of such quantum effects as probability and discontinuity in the macroscopic sphere of nature, which we observe directly with the help of our senses. Classical correspondence masks quantum reality.

Overcoming material realism

The principles of quantum theory make it possible to abandon the unfounded assumptions of material realism.

Assumption 1: Strict objectivity . The basic assumption of materialism is that there is a material universe independent of us. This assumption has some obvious operational validity and is often considered necessary for the meaningful pursuit of science. Is this assumption really justified? Quantum physics shows that we choose which aspect – wave or particle – a quantum object will demonstrate in a given situation. Moreover, our observation collapses the quantum wave packet into a localized particle. Subjects and objects are inseparably linked together. If this is so, how can one adhere to the assumption of strict objectivity?

Assumption 2: Causal determinism. Another assumption of classical science that underpins material realism is that the world is fundamentally deterministic: we only need to know the forces acting on each object and the initial conditions (the initial speed and position of the object). However, the principle of quantum uncertainty says that we can never simultaneously determine both the speed and position of an object with absolute accuracy. Our knowledge of the initial conditions will always contain error, and strict determinism is unacceptable. The very idea of ​​causality is equally suspect. Since the behavior of quantum objects is probabilistic, a strict cause-and-effect description of the behavior of a single object is impossible. Instead, when we talk about large groups of particles, we have statistical cause and statistical effect.

Assumption 3: Locality. The assumption of locality—that all interactions between material objects are mediated by local signals—is crucial to the materialist view that objects exist essentially separately and independently of each other. However, if the waves travel over vast distances and then suddenly collapse when we make a measurement, then the influence of our measurement is not transmitted locally. Thus, locality is excluded. This is another death blow to material realism.

Assumptions 4 and 5. Materialism and epiphenomenalism. Materialism maintains that subjective mental phenomena are merely epiphenomena of matter and can be completely reduced to the material brain. However, according to the principle of complementarity and the idea of ​​subject-object confusion, to understand the behavior of quantum objects, we seem to need to consider consciousness – our ability to make choices. Moreover, it seems absurd that an epiphenomenon of matter can affect matter: if consciousness is an epiphenomenon, then how can it “collapse” a dispersed wave of a quantum object into a localized particle when performing a quantum measurement?

Despite the principle of correspondence, the new paradigm of physics – quantum physics – contradicts the data of material realism. There is no way around this conclusion. We cannot say, invoking the correspondence principle, that classical physics is valid for macro objects for all practical purposes, and that since we live in the macro world, we will assume that quantum strangeness is limited to the submicroscopic sphere of nature. On the contrary, strangeness haunts us at the macro level. If we divide the world into the realms of classical and quantum physics, insoluble quantum paradoxes arise.

In India, people came up with a clever way to catch monkeys using a container of nuts. The monkey reaches into the container, grabbing a handful of nuts. Alas, having clenched the food in her fist, she can no longer pull her hand out – the neck of the vessel is too narrow. The trap works because the monkey’s greed prevents it from releasing the nuts. The axioms of material realism are materialism, determinism, locality, etc. – served us well in the past, when our knowledge was more limited than it is today, but now they have become a trap for us. We may have to give up the nuts of certainty to embrace the freedom that lies beyond the material arena.

If material realism cannot be an adequate philosophy for physics, then what philosophy can deal with all the strangeness of quantum behavior? This is the philosophy of monistic idealism that underlies all religions of the world.

Traditionally, only religions and the humanities recognize the value of human life beyond physical survival – a value derived from our love of beauty; our creative abilities in art, music and thought; and our spirituality in the intuition of unity. The natural sciences, locked within the framework of classical physics and its philosophical baggage of material realism, were a seducer of skepticism. Now the new physics is in dire need of a new, liberating philosophy suitable to the current level of our knowledge. If monistic idealism is the way to go, then the sciences and humanities, along with religions, will be able to go hand in hand in the search for all human truth for the first time since Descartes.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 4. PHILOSOPHY OF MONISTIC IDEALISM

Monistic idealism is the exact opposite of material realism. In this philosophy, consciousness, not matter, occupies a fundamental place. Both the world of matter and the world of mental phenomena are determined by consciousness. In addition to the material and mental realms (which together constitute immanent reality, or the world of manifestation), idealism posits a transcendent, archetypal realm of ideas as the source of material and mental phenomena. It is important to understand that monistic idealism, as its name suggests, is a unitary philosophy; any divisions, such as, for example, into immanent and transcendental, exist in consciousness. Only consciousness represents the ultimate reality.

In the West, the most influential formulation of the philosophy of monistic idealism comes from Plato, who gives his famous allegory of the cave in his Republic. As hundreds of generations of philosophy students have known, this allegory clearly illustrates the fundamental concepts of idealism. Plato depicts people sitting motionless in a cave and looking at the wall. The vast universe outside is projected onto the wall of the cave in the form of a shadow theater, and we humans observe these shadows. We observe the illusions of shadows, mistaking them for reality. True reality is behind us, in the light and archetypal forms casting shadows on the wall. In this allegory, the play of shadows corresponds to the unreal immanent manifestations in human experience of archetypal realities that belong to the transcendental world. In reality, the only reality is light, for light is all we see. In monistic idealism, consciousness is like the light in Plato’s cave.

The same basic ideas appear repeatedly in the idealistic literature of many cultures. In Indian Vedanta literature, the Sanskrit word
nama is used to denote the transcendental archetypes , and the word
rupa denotes their immanent form. Beyond
nama and
rupa shines the light
of Brahman – the universal consciousness, one without the other, the basis of all existence. “This entire universe that we talk and think about is nothing but
Brahman. Brahman is beyond
maya (illusion). There is nothing else.”

In Buddhist philosophy, the realm of matter and the realm of concepts are called, respectively,
nirmanakaya and
sambhogakaya, but beyond them there is the light of a single consciousness –
dharmakaya – illuminating them both. And, in fact, only
the dharmakaya exists. “Nirmanakaya is the outer appearance of the Buddha’s body and his inconceivable deeds.
The Sambhogakaya has enormous and limitless potential. The Buddha’s
Dharmakaya is free from any perception or concept of form.”

Perhaps the Taoist
yin-yang symbol is more widely known than the Indian symbols. Light
yang, considered a masculine symbol, defines the transcendental sphere, and dark
yin , a feminine symbol, defines the immanent.

Img. 13.
Yin-Yang symbol

Note their figure-ground relationship. “That which allows the dark and the light to manifest itself is
Tao ” – that which surpasses complementary manifestations.

Similarly, the Jewish Kabbalah describes two orders of reality: the transcendental one, which the Sephiroth represents as the Theogony, and the immanent one –
alma-de-peruda, “the world of separation.” According to the
Zohar, “When one contemplates things in mystical meditation, everything appears as one.”

In Christendom, the names of the transcendental and immanent realms—heaven and earth—are part of our everyday vocabulary. However, our everyday usage misses the origins of these concepts in monistic idealism. Beyond the realms of heaven and earth there is God, the King of these realms. Kingdoms do not exist separately from the King: the King is the kingdoms. As the Christian idealist Dionysius writes: “It [consciousness – the basis of being] is in our minds, souls and bodies, in heaven, on earth and everywhere, remaining one and the same in Itself. It is simultaneously in the world, around it and above it, super-heavenly and super-essential, the sun, star, fire, water, spirit, dew, cloud, stone, rock – everything that exists.”

Note that in all these descriptions it is stated that the unified consciousness is given to us through complementary manifestations: ideas and forms,
nama and
rupa, sambhogakaya and
nirmanakaya, yang and
yin, heaven and earth.

This complementary description represents an important aspect of idealist philosophy.

Usually, when we look around, we see only matter. Heaven is not a tangible object of ordinary perception. This is not only what leads us to call matter real, but also what leads us to accept the philosophy of realism, which declares matter (and its alternative form, energy) to be the only reality. However, many idealists argued that by going beyond the mundane, everyday experience, one could directly experience heaven. People who make such statements are called mystics. Mysticism offers experimental proof of monistic idealism.

Mysticism

Realism grows out of our everyday perception. Our daily experience of the world provides abundant evidence that things are material and separate from each other and from us.

Of course, mental experience does not agree with this formulation. Experiences of the mind such as thought do not appear to be material, and so we have come up with a dualistic philosophy that places the mind and body in different realms. The disadvantages of dualism are well known. In particular, it cannot explain how the separate, immaterial mind interacts with the material body. If this kind of mind-body interaction existed, then energy exchanges would have to occur between these two spheres. In numerous experiments we find that the energy of the material universe itself remains constant (this is the law of conservation of energy). There is also no data that would indicate the loss or acquisition of energy by the material sphere. How might this be if there were interactions between these two areas?

Idealism, although it considers consciousness to be the primary reality, and therefore attaches importance to subjective, mental experience, does not assume that consciousness is the mind. (Beware of possible semantic confusion:
consciousness is
a relatively new word in the English language. The word mind is often used to denote consciousness, especially in older literature
. In this book, the distinction between mind and consciousness is necessary and important .) Instead, idealism asserts that both material objects (such as a ball) and mental objects (such as the thought of a ball) are objects of consciousness. In addition, in experience there is a subject – the experiencer. What is the nature of this experiencer? In monistic idealism this question is of paramount importance.

According to monistic idealism, the consciousness of the subject in subject-object experience is the same consciousness that represents the basis of all existence. Therefore, consciousness is one. There is only one consciousness-subject, and we are this consciousness. “You are That,” says the sacred books of Hinduism known as the Upanishads.

Why then do we seem so separate in our everyday experience? As the mystics insist, this separateness is an illusion. If we meditate on the true nature of our self, we discover – as mystics of all times have discovered – that behind all diversity there is only one consciousness. This one consciousness-subject-self has many names. Hindus call it
atman; Christians call it the Holy Spirit, or in Quaker Christianity, the inner light. Whatever it is called, everyone agrees that the experience of this one consciousness is of inestimable value.

Buddhist mystics often refer to consciousness outside of man as the no-self, leading to the potential misconception that they may be denying consciousness entirely. The Buddha explained this misconception this way: “There is the Unborn, the Beginningless, the Uncreated, the Formless. If it were not for this Unborn, Beginningless, Uncreated, Formless, salvation from the world of the born, the beginning, the created, the formless would be impossible.”

Thus, mystics are those people who testify to this fundamental reality of unity in diversity. Comparison of mystical texts from different cultures and spiritual traditions speaks of the universality of the mystical experience of unity.

European mystic of the 15th century. Caterina Adorna from Genoa simply and beautifully formulated her knowledge: “My being is God, not by virtue of mere participation, but by virtue of a genuine transformation of my being.”

In China, 6th century. the great Huineng, an illiterate peasant whose sudden insight eventually led to the founding of Zen Buddhism, declared: “Our very nature-self is the Buddha, and besides this nature there is no other Buddha.”

Sufi mystic of the 12th century. Ibn al-Arabi, revered by the Sufis as the Sheikh of Sheikhs, said this: “You neither cease to be nor continue to exist. You are He, not bound by such limitations. Therefore, if you know that your own being is like this, then you know God; and if not, then no.”

In the XIV century. Kabbalist Moshe de Leon—the probable author of the main book of the Kabbalists,
the Zohar —wrote: “God… when he has just decided to begin his work of creation,
He is called. God in the full unfolding of His Being, Bliss and Love, in which He becomes capable of being perceived by the mind of the heart… is called
You. But God in His highest manifestation, where the fullness of His Being finds its full expression in the last and all-embracing of His attributes, is called Self.”

It is believed that the mystic of the 8th century. Padmasambhava brought Tantric Buddhism to Tibet. His wife, the divinely inspired Yeshe Tsogyal, expressed her wisdom this way: “But when you finally find me, one clear Truth arises from within: Absolute Awareness pervades the Universe.”

Meister Eckhart, a 13th-century Dominican friar, wrote: “In this breakthrough I realize that God and I are one. Then I am what I was, and neither decrease nor increase, for then I am the immovable cause that moves all things.”

Sufi mystic of the 10th century. Mansur al-Khalaj makes the statement: “I am the Truth!”

8th century Hindu mystic Shankara eloquently expressed his realization: “I am a reality without beginning, which has no equal. I do not participate in the illusion of “I” and “you”, “this” and “that”. I am Brahman, one without a second, bliss without end, eternal unchanging truth… I reside in all beings as soul, pure consciousness, the basis of all phenomena, internal and external. I am both the enjoyer and the enjoyed. In the days of my ignorance I considered all this to be separate from myself. Now I know that I am All.”

And finally, Jesus of Nazareth declared, “My Father and I are one.”

What is the meaning of the experience of oneness? For the mystic, it opens the door to a transformation of being that liberates love, universal compassion and freedom from the cohesion of life in acquired separateness and from the compensatory attachments to which we cling. (In Sanskrit this liberated existence is called
moksha.)

Idealist philosophy grew out of the experience and creative intuition of the mystics, who constantly emphasized the direct experiential aspect of fundamental reality. “The Tao that can be spoken of is not the absolute Tao,” said Lao Tzu. Mystics warn that all teachings and metaphysical writings should be considered fingers pointing to the moon, and not the moon itself.

As
the Lankavatara Sutra reminds us: “These teachings are but a finger pointing to the Noble Truth… They are intended to be considered and guided by the discerning minds of all men, but they are not the Truth itself, which man can comprehend only for himself, in the very depths of his own consciousness.”

Some mystics resort to paradoxical descriptions. Ibn al-Arabi writes: “Neither existence nor non-existence can be attributed to it (consciousness) … It is neither existing nor non-existent. He cannot be called either the First or the Last.”

Essentially, idealistic metaphysics itself can be considered paradoxical, since it includes the paradoxical concept of the beyond (transcendent). What is transcendental? Philosophy can only answer
neti, neti – neither this nor that. But what is it? Philosophy is silent. Or, as the
Upanishads say: “It is within all this / It is beyond all this.”

Is the transcendental sphere located within the immanent world? Yes. Is it outside the immanent world? Yes. This is very confusing.

Idealistic philosophy for the most part does not answer the following questions: “How is a holistic and indivisible consciousness divided into the reality of the subject-object? How does one consciousness become many?” We are not satisfied with the only answer that the observed multiplicity of the world is an illusion.

In this book we will argue that, given quantum physics, monistic idealism is the correct philosophy for science. In addition, the integration of science and mysticism helps resolve some of the difficult issues that mysticism raises.

The integration of science and mysticism should not be too confusing – after all, they have one important similarity: both science and mysticism grew out of empirical data interpreted in the light of theoretical explanatory principles. In science, theory serves both as an explanation of data and as a tool for prediction and guidance for future experiments. Idealistic philosophy can also be seen as a creative theory that acts as an explanation for the empirical observations of mystics as well as a guide for other seekers of Truth. Finally, mysticism, like science, appears to be universal. There is no parochialism in mysticism – it arises when religions simplify mystical teachings to make them more suitable for transmission to the masses.

Religion

To come to an understanding of the Truth, the mystic usually finds and uses one or another methodology. Methodologies, or spiritual paths, have both similarities and differences. Differences that are secondary to the mystical insight itself contribute to the differences in religions based on the teachings of the mystics. For example, Buddhism developed from the teachings of Buddha, Judaism from the teachings of Moses, Christianity from the teachings of Jesus, Islam from the teachings of Mohammed (although, strictly speaking, Mohammed is considered the last of a number of prophets, including Moses and Jesus), and Taoism from the teachings of Lao Tzu. However, there are no rules without exceptions. Hinduism is not based on the teachings of any one teacher, but rather includes many teachings and many paths.

Mysticism involves the search for truth about ultimate reality, but religion has a slightly different function. Followers of this or that mystic (most often after his death) may realize that the individual search for truth is not for everyone. Most people, lost in the illusion of the separateness of their ego and busy fulfilling its aspirations, do not feel the urge to discover the truth for themselves. How then can one share the light of mystical comprehension with these people?

The answer is by simplifying it. Followers simplify the truth to make it accessible to the average person. This person is usually consumed by the demands of everyday life. Lacking the time and commitment necessary to understand the subtleties of transcendence, he cannot appreciate the importance of direct mystical experience. Therefore, the disseminators of the truth discovered by the mystic replace the direct experience of a single consciousness with the idea of ​​God. Unfortunately, God, the transcendent creator of the immanent world, is transformed in the mind of the average person into the dualistic image of a powerful Lord in Heaven ruling over the Earth below. The mystic’s revelation inevitably becomes emasculated and distorted.

The mystic’s followers, acting with the best of intentions, unwittingly play the role of the devil in an old joke: One day God and the devil were walking together, and God picked up a piece of paper. “What does it say?” – asked the devil. “True,” God calmly answered. “Give it to me,” said the devil impatiently. “I’ll organize it for you.”

However, despite the difficulties and errors of systematization, religion still conveys the spirit of the mystic’s revelation – this is what gives it vitality. After all, for mystics, the significance of realizing the transcendental nature of Reality is that they are strengthened in a mode of being where virtues such as love become simple. How can you not love, knowing that there is one consciousness and that you and the other are not really separate from each other?

But how can an ordinary person, who is not aware of unity, be motivated to love others? The mystic clearly understands that ignorance of transcendental unity is an obstacle to love. The end result of the absence of love is suffering. To avoid suffering, the mystic advises us to turn inward and begin a journey of self-discovery. In a religious context, this teaching is transformed into a statement that if we want to be saved, we must turn to God as the highest value in our lives. The method of this salvation is a set of practices based on the original teachings that form the moral code of a particular religion – the Ten Commandments and the Golden Rule of Christian ethics, the precepts of Buddhism, the law of the Koran or Talmud, and so on.

Of course, not all religions introduce the concept of God. For example, in Buddhism there is no concept of God. On the other hand, there are many gods in Hinduism. However, even in these cases, the above considerations regarding religion are obvious. Thus we come to three universal aspects of all exoteric religions:

1. All religions start from the premise that our way of life is wrong. Wrongness has different names – ignorance, original sin, or simply suffering.

2. All religions promise a way out of this wrongness, provided that the “path” is followed. This way out is called salvation, liberation from the wheel of suffering in the world, enlightenment, or eternal life in the Kingdom of God – paradise.

3. The path consists of adherence to religion and the community of followers of religion, and adherence to a prescribed code of moral and social rules. In addition to the way different religions distort the esoteric teaching of transcendence, they differ from each other precisely in their codes of ethical and social rules.

Note the obligatory dualism of the first point: wrong and right (or evil and good). In contrast, the mystical path consists of transcending all dualities, including good and evil. Note also that the clergy turns the second point into a carrot and a stick—hell and heaven. On the other hand, mysticism does not contrast heaven and hell, considering both as natural accompanying circumstances of our way of life.

As you can see, when filtered through the religions of the world, the monism of monistic idealism becomes even more obscure and dualistic ideas predominate. In the East, thanks to the endless influx of people willing to study mysticism, monistic idealism in its esoteric form, at least in part, retained fame and respect among the general public. However, mysticism has had relatively little influence in the West. The dualism of Judeo-Christian monotheistic religions, supported by a powerful hierarchy of interpreters, prevailed in the mass consciousness. But like the Cartesian mind-body dualism, God-world dualism does not seem to stand up to scientific scrutiny. As scientific evidence undermines religion, there is a tendency to throw out the baby with the dirty bathwater—ethics and values ​​taught by religion—ethics and values ​​that continue to be valid and useful.

Exposing the illogicality of dualistic religions does not necessarily lead to a monistic philosophy of material realism. As we have seen, there is an alternative monism. Given how quantum physics refutes material realism, monistic idealism may be the only viable monistic philosophy of reality. Another option is to abandon metaphysics entirely, which has been the main focus of philosophy for some time. Currently, this trend appears to be reversing.

We must now pose the decisive question: is science compatible with monistic idealism? If not, then we should abandon metaphysics in the pursuit of science, further aggravating the threatening crisis of faith. If so, then we must reformulate science according to the demands of philosophy. In this book we argue that monistic idealism is not only compatible with quantum physics, but also necessary for its interpretation. The paradoxes of the new physics disappear when we consider them from the point of view of monistic idealism. Moreover, quantum physics, combined with monistic idealism, gives us a powerful paradigm with which to resolve some of the paradoxes of mysticism – for example, the question of transcendence and multiplicity. Our work points to the beginnings of idealistic science and the revival of religion.

An idealistic metaphysics for quantum objects

Quantum objects exhibit complementary aspects of wave and particle. Is quantum complementarity—the resolution of wave-particle duality—the same thing as the complementarity of monistic idealism?

The writer George Leonard clearly saw a parallel between these two types of complementarity when he wrote in The Silent Pulse: “Quantum mechanics is the ultimate koan of our time.” Koans are tools used in Zen Buddhism to break through apparent paradoxes to transcendental solutions. Let’s compare koans with complementarity.

In one koan, Zen student Daibei asks Zen master Baso, “What is Buddha?” Baso replies: “The mind is the Buddha.” When another monk asked the same question, Baso replied, “This mind is not the Buddha.”

Compare this with Bohr’s principle of complementarity. Ask Bohr: “Is an electron a particle?” Sometimes Bohr can answer: “Yes.” When we look at the trace of an electron in a condensation chamber, it makes sense to say that the electron is a particle. However, Bohr, puffing on his pipe, will say: “You must agree that the electron is a wave.” Bohr, like the Zen master, seems to be of two minds about the nature of electrons.

Quantum waves are waves of probability. To see the wave aspect, such as the diffraction pattern, it is necessary to experiment with many wave particles.
We can never experimentally see the wave aspect of a single quantum object: a single wave particle is always detected as a localized particle . Nevertheless, even a single wave particle has a wave aspect. Does the wave aspect of a single wave particle exist in transcendental space since it never appears in ordinary space? Does Bohr’s idea of ​​complementarity point to the same transcendental order of reality that the philosophy of monistic idealism speaks of?

Bohr never gave a definitive positive answer to such questions, and yet his Nobel coat of arms featured the Chinese
yin-yang symbol. Could it be that Bohr understood the complementarity of quantum physics in a similar way to monistic idealism, that he was a proponent of idealistic metaphysics as applied to quantum objects?

Let’s remember the uncertainty principle. If the product of the uncertainties of position and momentum is a constant, then decreasing the uncertainty of one increases the uncertainty of the other. Extrapolating this conclusion, one can see that if the position is known with complete certainty, then the impulse becomes completely uncertain, and vice versa – when the impulse is known with complete certainty, the position becomes completely uncertain.

Many newcomers to quantum physics object to these implications of the uncertainty principle, saying, “But surely the electron must be somewhere—we just don’t know where.” No, things are worse. We cannot even determine the position of an electron in ordinary space-time. Obviously, quantum objects exist in a completely different way from the familiar macro-objects of everyday life.

Heisenberg also recognized that a quantum object cannot occupy a given place and at the same time move in a predictable manner. Any attempt to take a snapshot of a submicroscopic object provides only its position, but information about its state of motion is lost. And vice versa.

This observation raises another question: what does the object do between snapshots? (This is analogous to the question about electrons making quantum jumps between the orbits of a Bohr atom: where does the electron go between jumps?) We cannot assign a specific trajectory to the electron. To do this, we would need to know its initial speed and position, which would violate the uncertainty principle. Can we attribute to the electron any apparent reality in space and time between observations? The Copenhagen interpretation of quantum mechanics gives a negative answer to this question.

Between observations, the electron, in accordance with the Schrödinger equation, spreads out – but, according to Heisenberg, probabilistically, in potency (Heisenberg took the term
potency from Aristotle). Where do these potencies exist? Since the electron wave immediately collapses upon observation, the potencies cannot be located in the material realm of space-time; As you remember, in space-time all objects must obey the speed limit established by Einstein. Therefore, the sphere of potency must be outside space-time. Potentialities exist in the transcendental realm of reality. Between observations, the electron, like Plato’s archetypes, exists as a form of possibility in the transcendental realm of potentialities. (The poet Emily Dickinson writes, “I dwell in Possibility.” If the electron could talk, this would probably be how it would describe itself.)

Electrons are too far removed from ordinary personal reality. Suppose we ask, “Does the moon exist when we are not looking at it?” To the extent that the moon is ultimately a quantum object (composed entirely of quantum objects), we must say no; So says physicist David Mermin. Between observations, the moon also exists as a form of possibility in the transcendental sphere of potentialities.

Perhaps the most important and most insidious assumption we learn as children is that there is a material world of objects outside of us, independent of the subjects who observe it. There is detailed evidence in favor of such an assumption. For example, whenever we look at the moon, we find it where we expect it to be, according to its classically calculated trajectory. Naturally, we assume that the moon is always there in space-time, even when we are not looking at it. Quantum physics says no. When we do not look at the moon, its wave of possibility blurs, although by an extremely small amount; When we look, the wave collapses; therefore, the wave could not be in space-time. It makes more sense to accept the assumption of idealistic metaphysics: no object exists in space-time without a conscious subject looking at it.

So, quantum waves are like Plato’s archetypes in the transcendental realm of consciousness, and the particles that appear as a result of our observation are immanent shadows on the cave wall. Consciousness is the factor that causes the collapse of the wave of a quantum object existing in potency, making it an immanent particle in the world of manifestation. This is the basic tenet of idealistic metaphysics that we will use in this book for quantum objects. We will see that in the light of this simple idea, all the famous paradoxes of quantum physics melt away like morning fog.

Note that Heisenberg himself came close to idealistic metaphysics when he proposed the concept of potency. An important new element is that the sphere of potency also exists in consciousness. There is nothing outside consciousness. This monistic view of the world is crucial.

Science discovers the transcendental

Before the modern interpretation of new physics, the word “transcendence” was rarely mentioned in the dictionary. The term was even considered heretical (and still is for adherents of science, which obeys classical laws in a deterministic and mechanistic universe of cause and effect).

For the philosophers of Ancient Rome, transcendence meant “the state of going beyond all possible experience and knowledge” or “being beyond comprehension.” In monistic idealism, transcendental also means “not this, and not anything known.” Today, modern science is invading areas that for more than four thousand years were the domain of religion and philosophy. Is the universe just an objectively predictable series of phenomena that man can observe and control, or is it much more elusive and even more amazing? Over the past three centuries, science has become the unrivaled touchstone of reality. We are fortunate to be part of this evolutionary and transcendental process in which science not only changes itself, but also changes our understanding of reality.

An exciting achievement – an experiment by a team of physicists in Orcy, France – not only confirmed the idea of ​​transcendence in quantum physics, but also clarified the very concept of transcendence. The experiment of Alain Aspect and his collaborators shows that when two quantum objects are “correlated”, then when one of them is measured (causing the collapse of its wave function), the wave function of the other also instantly collapses – even at a macroscopic distance, even in the absence of a signal in space – time mediating their connection. However, Einstein argued that all connections and interactions in the material world must be mediated by signals propagating in space (the principle of locality), and therefore must be limited by the speed of light. Where then is the instantaneous communication between correlated quantum objects that is responsible for their action at a distance without signaling? The short answer is: in the transcendental realm of reality.

In physics, instantaneous action at a distance that is not mediated by signals is called nonlocality. The correlation of quantum objects in the Aspect experiment is a non-local correlation. Once we recognize quantum nonlocality as an established physical aspect of the world in which we live, it becomes easier in science to speak of a transcendental realm beyond the manifest physical realm of spacetime. According to physicist Henry Stapp, quantum nonlocality indicates that “a fundamental process of Nature lies outside space-time, but produces events that can be detected in space-time.”

Warning: If the words “outside space” make you think of another “box” outside the dimensional “box” we are in, forget it. Another box, by definition, can be made as much a part of the universe of space as our own. In the case of non-local communication, we are forced to think about a realm of reality outside of space-time, since non-local communication cannot happen in space-time.

There is another paradoxical way to imagine non-local reality – as being everywhere and nowhere, always and never. It’s still paradoxical, but it’s thought-provoking, isn’t it? I like to play with the word
nowhere
, which I first read as “now here” as a child
. Nonlocality (and transcendence) is nowhere and now here.

About 2,500 years ago, Democritus proposed the philosophy of materialism, but shortly thereafter Plato gave one of the first clear formulations of the philosophy of monistic idealism. As Werner Heisenberg observed, quantum mechanics shows that of the two thinkers, Plato and Democritus, who most influenced Western civilization, Plato may ultimately emerge victorious. The success enjoyed in science by the atomism of Democritus in the last three centuries can only be a temporary delusion. Quantum theory, interpreted from the position of idealistic metaphysics, opens the way for idealistic science, in which consciousness takes the leading place, and matter recedes into the background.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES

Habits of thought are tenacious. Although quantum mechanics has replaced classical mechanics as the fundamental theory of physics, many physicists brought up with the old worldview still find it difficult to accept the idealistic implications of quantum mechanics. They don’t want to ask the difficult questions that quantum mechanics raises. They hope that if such problems are ignored, they will disappear. Once, at the beginning of a discussion of paradoxes in quantum mechanics, Nobel laureate Richard Feynman ridiculed this position in his inimitable ironic manner. He said: “Hush, hush. Close the doors.”

In the next five chapters, we will open doors and unashamedly expose the paradoxes of quantum physics. Our goal will be to demonstrate that, when viewed in the light of monistic idealism, quantum paradoxes are not so shocking or paradoxical after all. Strict adherence to idealistic metaphysics, based on a transcendental, unifying consciousness that “collapses” the quantum wave, naturally resolves all the paradoxes of quantum physics. We will find that it is quite possible to do science within the conceptual framework of monistic idealism. The result is an idealistic science that unites spirit and matter.

The idea that consciousness collapses a quantum wave was originally proposed by the mathematician von Neumann in the 1930s. Why did it take us so long to take this idea seriously? Perhaps a brief discussion of how my own understanding of the problem has developed will help answer this question.

One of the difficulties that prevented me from accepting von Neumann’s hypothesis concerned the experimental data. Apparently, when we look, it always happens consciously. Then the question of consciousness “collapsing” quantum waves seems purely academic. Is it even possible to find a situation in which a person looks, but does it unconsciously? Notice how paradoxical this seems.

In 1983, I was invited to a ten-week seminar on consciousness in the psychology department at the University of Oregon. I was especially flattered that these psychological scientists patiently listened to six hours of lectures in which I talked about quantum concepts. However, I was truly rewarded when one of the graduate students in psychologist Michael Posner’s group reported some cognitive findings from a guy named Tony Marcel. Some of this data concerned “unconscious vision” – exactly what I was looking for.

I listened with bated breath to the report and only relaxed when I realized that these data were completely consistent with the idea that consciousness “collapses” the quantum state of the brain-mind during conscious vision (see Chapter 7). In unconscious vision there is no “collapse” and this really makes a huge experimental difference. I soon also realized how to resolve the small paradox that creates the difference between conscious and unconscious perception.

The trick is to differentiate between consciousness and awareness .

CHAPTER 5. OBJECTS LOCATED IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES

The fundamental principles of material realism simply do not hold up. Instead of causal determinism, locality, strict objectivity and epiphenomenalism, quantum mechanics offers probability and uncertainty, wave-particle complementarity, non-locality and confusion of subjects and objects.

Objecting to the probabilistic interpretation of quantum mechanics, which generates uncertainty and complementarity, Einstein used to declare that “God does not play dice.” To understand what he meant, imagine that you are conducting an experiment with a sample of a radioactive substance, the decay of which, of course, obeys probabilistic quantum laws. Your job is to measure the time during which ten radioactive decay events occur—ten clicks of your Geiger counter. Assume that ten decay events occur in about half an hour. Behind this average lies a probability. Some episodes are 32 minutes, others 25 minutes, etc. To complicate matters further, you need to catch a bus to meet your crush, who hates being kept waiting. And guess what happens? Your final episode takes forty minutes to complete because the random decay of a single atom does not occur as it would on average. So you miss the bus, your lover breaks up with you, and your life is ruined. This may be a somewhat silly and far-fetched example of what happens in a world where God plays dice, but it is quite indicative of the fact that probabilistic events can only be relied upon on average.

The randomness of atomic events—the capriciousness of fate, so to speak—is incompatible with determinism. The determinist thinks of probability as it is commonly thought of in classical physics and in everyday life: it is a property of large collections of objects—collections so large and complex that in practice we cannot predict them, although in principle such a prediction is possible. For a determinist, probability is simply a convenience of thinking: the physical laws governing the movements of individual objects are completely certain, and therefore completely predictable. Einstein believed this was the case in the quantum mechanical universe. Behind quantum uncertainties are hidden variables. The probabilities of quantum mechanics are merely for convenience. If this were so, quantum mechanics would be a theory of aggregates. Indeed, if we had not applied the probabilistic wave description to a single quantum object, we would not have encountered the paradoxes that concern us – particle-wave complementarity and the inseparability of a quantum object from the circumstances of its observation.

Unfortunately, it’s not that simple. Consideration of two quantum mechanical experiments will show how difficult it is to give a rational explanation for the paradoxes of quantum physics.

Double-slit experiment

We can never see the wave aspect of a single particle. Whenever we look, only a localized particle appears to our gaze. Must we, therefore, assume that the solution is a transcendental metaphysics? Or should we abandon the idea that there is a wave aspect to a single wave particle? Perhaps the waves that quantum physics deals with are a property that is inherent only in groups or collections of objects?

To determine whether this is true, we can analyze an experiment commonly used to study wave phenomena, the so-called double-slit experiment. In the setting of this experiment, a flow of electrons passes through a partition with two narrow slits (see Fig. 14). Since electrons are waves, a double-slit baffle divides the electron beam into two sets of waves. These waves then interfere with each other, and experimenters see the result of the interference on a fluorescent screen.

Img. 14.
Double-slit experiment for electrons

Quite simple? Let me consider the phenomenon of wave interference. If you are not familiar with the phenomenon of interference, it can be easily demonstrated by standing in a bathtub filled with water and rhythmically marching in place, causing two series of waves to appear on the water. The waves will create an interference pattern (Fig. 15, a). In some places they will reinforce each other (Fig. 15, b), and in others they will mutually destroy (Fig. 15, c). The result is an interference pattern.

Img. 15.
a – when waves on water interfere, an interesting picture of mutual amplification and cancellation is created; b – when the waves arrive in the same phase, they reinforce each other; c – waves arriving in antiphase cancel each other out

Similarly, there are places on a fluorescent screen where the electron waves from both slits arrive in the same phase; in such places their amplitudes add up, and the total wave intensifies. Between these bright spots there are places where the waves arrive in antiphase and cancel each other out. Thus, the result of this creative and destructive interference appears on the screen as a pattern of alternating light and dark stripes – an interference pattern (Fig. 16). It is important that the intervals between the stripes make it possible to measure the wavelength of electronic waves.

Img. 16.
Interference pattern of flashes on the screen

However, remember that electron waves are probability waves. Therefore, we must talk specifically about probability: electrons falling into bright regions have a high probability, and electrons falling into dark regions have a low probability. We should not get carried away and, based on the interference pattern, conclude that electron waves are classical waves, since electrons still hit the fluorescent screen in the same way as particles should: each electron gives one localized flash. It is the collection of spots formed by a large number of electrons that looks like a pattern of wave interference.

Suppose we take an intellectual risk and make the electron beam very weak—so weak that only one electron reaches the slits at any given moment. Do we still get an interference pattern? Quantum mechanics clearly answers yes. You might object – you can’t get interference without splitting the beam. Don’t you need two waves for interference? Can a single electron split, pass through both slits, and interfere with itself? Yes maybe. Quantum mechanics answers all these questions positively. In the words of one of the pioneers of the new physics, Paul Dirac: “Every photon (or in this case, electron) interferes only with itself.” Quantum mechanics offers a mathematical proof of this absurd claim, but this single claim is responsible for all the amazing magic that quantum systems are capable of and which has been proven by many experiments and technologies.

Try to imagine that an electron passes 50% through one slit and 50% through the other slit. It’s easy to get angry and disbelieve in this strange implication of quantum mathematics. Does an electron actually pass through both slits at the same time? Why should we take this for granted? We can find out through observation. We can shine a flashlight on the slits (metaphorically speaking) to see which slit the electron actually passes through.

So, we turn on the light and, seeing an electron passing through one or another slit, we look where the flash appears on the fluorescent screen (Fig. 17). We find that every time an electron passes through a slit, its flash appears exactly behind the slit through which it passes. The interference pattern has disappeared.

Img. 17.
When we try to determine through which slit an electron passes by shining a flashlight on the slits, the electron demonstrates its corpuscular nature. There are only two bands – exactly as you would expect if electrons were miniature balls

What happens in this experiment can be understood primarily as a consequence of the uncertainty principle. Once we detect an electron and determine which slit it passes through, we lose information about the electron’s momentum. Electrons are very sensitive; a collision with the photon we are using to observe the electron affects it in such a way that its momentum changes by an unpredictable amount. The momentum and wavelength of the electron are interrelated: quantum mechanics includes this great discovery of de Broglie. Therefore, the loss of information about the momentum of an electron is the same as the loss of information about its wavelength. If there were interference fringes, then we could measure the wavelength by the distances between them. The uncertainty principle states that once we determine which slit the electron is passing through, the process of observation destroys the interference pattern.

You must understand that measuring the position and momentum of an electron are actually complementary, mutually exclusive procedures. We can focus on the pulse and measure the wavelength—and thus the momentum—of the electron from the interference pattern, but then we cannot know which slit the electron is passing through. Or we may focus on the electron’s position and lose the interference pattern—information about its wavelength and momentum.

There is a second, even more clever way of understanding and reconciling all this – using the principle of complementarity. Depending on which device we use, we see the particle aspect (for example, with a flashlight) or the wave aspect (without a flashlight).

To a first approximation, the essence of the principle of complementarity comes down to the fact that quantum objects are both waves and particles, but we can see only one aspect using a particular experimental setting. This is undoubtedly a correct understanding, but experience teaches us some subtleties. For example, we must also say that an electron is neither a wave (since the wave aspect never appears for a single electron) nor a particle (since it appears on the screen in places where particles are prohibited). Then, being careful in our logic, we must say that a photon is neither a non-wave nor a non-particle, in order to avoid misunderstanding our use of the words “wave” and “particle”. This is very similar to the logic of someone who lived in the 1st century. n. e. idealist philosopher Nagarjuna – the most insightful logician of the Mahayana Buddhist tradition. Eastern philosophers convey their understanding of ultimate reality with the words neti, neti (neither this nor that). Nagarjuna formulated this teaching in the form of four negatives:

She doesn’t exist.
She is not non-existent.
It cannot be said about it
that it both exists and does not exist.
Or that it is neither
existing nor non-existent.

To better understand complementarity, suppose we return to the previous experiment, this time using weak batteries to make the flashlight with which we illuminate the electrons somewhat dimmer. Repeating the experiment shown in Fig. 17, with the lantern light becoming dimmer and dimmer, we find that the interference pattern begins to reappear, becoming increasingly clear as the lantern light becomes dimmer (Fig. 18). When the flashlight is completely turned off, the full interference pattern is returned.

Img. 18.
When using a dimmer flashlight, the interference pattern partially returns

As the flashlight dims, the number of photons scattering the electrons decreases, so that some electrons manage to avoid being “seen” by the light entirely. Those electrons that are visible appear behind slit 1 or slit 2, exactly where we would expect to find them. Each of the unseen electrons splits and interferes with itself, forming an interference pattern on the screen when enough electrons reach it. In the extreme case of bright light, only the corpuscular nature of the electrons is visible; in the limiting case of the absence of light, only the wave nature is visible. In intermediate cases of dim light, both aspects are visible to a similar intermediate degree: that is, here we see electrons (though never the same electron) as both waves and particles. Thus, the wave nature of a wave particle is not a property of the entire aggregate, but must remain valid for each individual wave particle when we are not looking at it. This must mean that the wave aspect of a single quantum object is transcendental since we never see it manifest.

A series of pictures help explain what is happening (Fig. 19). In the picture below on the left we only see the letter W; this corresponds to the use of a bright flashlight, which shows only the corpuscular nature of the electrons. Then, moving upward from picture to picture, we begin to see the eagle – just as when the brightness of light decreases, some electrons escape observation (and localization), and we begin to see their wave nature. Finally, in the last, top right picture, you can only see the eagle; The flashlight is turned off and all electrons are now waves.

Img. 19.
Sequence W—Heads

Niels Bohr once said: “Those who were not shocked when they first encountered quantum theory probably did not understand it.” As we begin to comprehend the workings of the principle of complementarity, this shock gives way to understanding. Then the official march of predictive science, valid for either the wave or the particle, is transformed into the creative dance of the transcendental wave particle. When we localize an electron by finding out which slit it passes through, we discover its corpuscular aspect. When we don’t localize an electron, regardless of which slit it passes through, we discover its wave aspect. In the latter case, the electron passes through both slits.

Delayed choice experiment

This unique property of the principle of complementarity should be clearly understood: what attribute a quantum waveparticle reveals depends on the way we choose to observe it. The importance of conscious choice in shaping manifest reality is best demonstrated by the delayed choice experiment proposed by physicist John Wheeler.

In Fig. 20 shows a device in which a ray of light is divided into two rays of equal intensity – reflected and transmitted – using a semi-transparent mirror M 1. Then both rays are reflected by two ordinary mirrors A and B and reach the intersection point P on the right.

To detect the wave aspect of a wave particle, we use the phenomenon of wave interference and place a second semi-transparent mirror M 2 at point P (Fig. 20, bottom left). Now the mirror M 2 causes both waves created by the beam that is split by the mirror M 1 to interfere creatively on one side of P (if you put a photon counter there, it will click) and destructively interfere on the other side (where the counter never clicks). Note that when detecting the wave mode of photons, we must recognize that each photon is separated in the mirror M 1 and travels in both paths A and B, otherwise how can there be interference?

Therefore, when mirror M 1 splits the beam, each photon is potentially ready to travel in both ways. If we now decide to detect the corpuscular mode of photon wave particles, we remove the mirror M 2 from point P (to prevent recombination and interference) and place counters behind the intersection point P, as shown in Fig. 20 bottom right. One or the other counter will click, identifying the localized path of the wave particle—reflected path A or transmitted path B —and demonstrating the particle aspect.

Img. 20.
Delayed choice experiment. BOTTOM LEFT: Experimental setup for seeing the wave nature of photons. One of the detectors never detects any photons, indicating extinction due to wave interference. The photon had to split and travel along two paths at the same time. BOTTOM RIGHT: Setting for seeing the corpuscular nature of photons. Both detectors click, but alternately, indicating which path the photon is taking.

The trickiest part of the experiment is this: in the delayed choice experiment, the experimenter decides whether or not to place a translucent mirror at point P – to measure the wave aspect or not – at the very last moment, at the very last picosecond (10 -12 s) (this was real carried out in the laboratory). This essentially means that the photons have already passed the separation point (if you think of them as classical objects). Even in this case, placing a mirror at point P always shows the wave aspect, and not placing a mirror shows the particle aspect. Did each photon travel along one path or two? Apparently, photons react instantly and retroactively even to our delayed choices. The photon travels along one path or both exactly according to our choice. How does he know about it? Does the effect of our choice precede its cause in time? In the words of John Wheeler: “Nature at the quantum level is not a machine going its inexorable path. Instead, the answer we get depends on what question we ask, what experiment we set up, what recording device we choose. We inevitably become involved in causing what happens.”

There is no manifest photon before we see it, and so how we see it determines its attributes. Before our observation, the photon is split into two wave packets (one packet for each path), but these packets are only packets of possibilities for the photon; in M 1 there is no reality in space-time, no decision-making. Does the effect precede its cause, thereby violating the law of causation? Undoubtedly, yes – if you think of the photon as a classical particle, always manifested in space-time. However, the photon is not a classical particle.

From a quantum physics perspective, by placing a second mirror at point P in our delayed choice experiment, both separated packets potentially connect and interfere; there is no problem here. If there were a mirror at point P and we removed it at the last possible picosecond, finding a photon on, say, path A, then the photon would appear to react retroactively to our delayed choice by moving along only one path. Therefore, in this case it would seem that the effect precedes the cause. This result does not violate the law of causality. How so?

It is necessary to understand a more subtle way of looking at the second experiment to detect the corpuscular aspect of photons; as Heisenberg explains: “If the experimental result now indicates that a photon is located in, say, the reflected part of the [wave] packet [path A], then the probability of finding a photon in another part of the beam immediately becomes zero. Then the experiment with the position of the reflected packet produces a kind of effect… at a distant point occupied by the passing packet, and the observer sees [that] this effect propagates at a speed exceeding the speed of light. However, it is also obvious that this type of action can never be used to transmit a signal, so it … does not contradict the postulates of the theory of relativity.”

This action at a distance constitutes an important aspect of the collapse of the wave packet. To denote such an action, a special term is used – non-locality – an action transmitted without signals that propagate in space. Signals that propagate in space in a finite time, due to the speed limit established by Einstein, are called local signals. Therefore, the collapse of a quantum wave is not local.

Note that Heisenberg’s statement is true both in the presence and absence of delayed choice. From a quantum point of view, the important thing is that we choose one or another outcome, which manifests itself; when in time we choose this outcome does not matter. The wave divides whenever there are two paths available, but the division occurs only in potency. When, later, we observe a photon on one path because we choose that outcome (removing the mirror from point P), the collapse we cause of the wave on one path has a non-local effect on the wave on the other path, which negates the possibility of seeing the photon on this other way. Such nonlocal influence may seem retroactive (i.e., transmitted back in time), but we influence only potentialities; There is no violation of the law of causality here, since, as Heisenberg says, we cannot transmit a signal using this kind of device.

In our search for the meaning and structure of reality, we are faced with the same riddle that Winnie the Pooh faced:
“  Hello, Pooh,” said Piglet, “what are you doing?”
“I’m hunting,” said Pooh.
Are you hunting? On whom?
“  I’m tracking someone,” Winnie the Pooh answered very mysteriously.
–  Who are you tracking? – asked Piglet, coming closer.
“  That’s exactly what I’m asking myself.” This is the whole question – who?
–  And how do you think you will answer this question?
“  I’ll have to wait until I catch up with him,” said Winnie the Pooh.
–  Look here. “He pointed to the ground directly in front of him. – What do you see here?
“  Traces,” said Piglet. – Paw prints! “He even squealed a little with excitement.
–  Oh, Pooh! Do you think this is… this is the scary Buka?
“  Maybe,” said Pooh. “Sometimes it’s like he is, and sometimes it’s like he’s not.” Can you guess by the footprints?..
“  …Just a minute,” said Winnie the Pooh, raising his paw. He sat down and thought as deeply as he could. Then he tried his paw on one of the Footprints… and then scratched his ear twice and stood up. “Yes,” said Winnie the Pooh. – Now I understand. “I was a stupid simpleton,” he said.
“  And I’m the most stupid bear cub in the world!”
–  What you! You are the best teddy bear in the world! – Christopher Robin consoled him.

Indeed, it is somewhat puzzling that, according to new physics, the “beech” traces that electrons and other submicroscopic particles leave in our condensation chambers are simply an extension of ourselves.

Classical science invariably saw only division in the world. Two centuries ago, the English Romantic poet William Blake wrote:

God save us from uniform vision and Newtonian sleep.

Quantum physics is the answer to Blake’s prayer. The modern scientist, having learned the lesson of the principle of complementarity, is not so stupid as to “obsess” with (apparent) separateness.

Quantum measurements bring our consciousness onto the stage of the so-called objective world. There is no paradox in the delayed choice experiment if we give up the idea that a constant and independent world exists even when we do not observe it. Ultimately it comes down to what you, the observer, want to see. This reminds me of a Zen story.

Two monks were arguing about the movement of a flag in the wind. One said: “The flag is moving.” Another objected: “No, it’s the wind that moves.” A third monk, passing by the debaters, made a remark that Wheeler would have approved: “The flag is not moving. The wind doesn’t move. Your mind moves.”

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 6. NINE LIVES OF SCHRÖDINGER’S CAT

Many of the founders of quantum physics found its strange implications difficult to accept. Schrödinger himself expressed his doubts about the interpretation of quantum mechanics in terms of probability waves in a paradox now known as Schrödinger’s Cat.

Suppose we put a cat in a cage with a radioactive atom and a Geiger counter. A radioactive atom will decay in accordance with the laws of probability. If the atom decays, the Geiger counter will go off and turn on the hammer, the hammer will break the bottle of poison, and the poison will kill the cat. Let’s say that the probability that this will happen within an hour is 50% (Fig. 21).

Img. 21.
Schrödinger’s cat paradox

Then how does quantum mechanics describe the state of a cat after an hour? Of course, if we look, we will find that the cat is either alive or dead. What if we don’t look? There is a 50% chance that the cat is dead. The probability that the cat is alive is also 50%.

If we think classically, as required by material realism, and are guided by the principles of determinism and causal continuity, then we could draw a mental analogy with a situation in which someone tossed a coin and then covered it with his palm. We don’t know whether it came up heads or tails, but of course it came up as one or the other. The cat is either alive or dead, with a 50% probability of each outcome. We simply do not know which outcome actually materialized. This is not the scenario suggested by the mathematics of quantum mechanics. Quantum mechanics approaches probabilities very differently. She describes the cat’s state at the end of the hour as half dead, half alive. Inside the box there is quite literally a “coherent superposition of a half-living and half-dead cat,” as it sounds in the technical jargon of quantum physics. The paradox of a cat being alive and dead at the same time is a consequence of how calculations are done in quantum mechanics. No matter how strange the consequences of this mathematics may be, we must take it seriously, since the same mathematics gives us the wonders of transistors and lasers.

This absurd situation is summed up by the following parody from T. S. Eliot’s Old Possum’s Book of Practical Cats:
Schrödinger’s cat is a mysterious cat,
it illustrates the laws;
the complicated things she does have no
apparent reason;
she confuses determinists
and drives them to despair,
because when they try to catch her,
there is no trace of the quantum cat!

The parody is, of course, true – no one has seen a quantum cat, or a coherent superposition, in fact, not even quantum physicists. In fact, if we look into the box, we see either a living or a dead cat. The inevitable question arises: what is so special about our act of observation that it can solve the cat’s devilish dilemma?

It’s one thing to plausibly talk about an electron passing through two slits at the same time, but when we talk about a cat being half alive and half dead, the absurdity of quantum coherent superposition becomes self-evident.

One way out is to insist that the mathematical prediction of coherent superposition should not be taken literally. Instead, following the interpretation in terms of ensemble statistics favored by some materialists, one can convince oneself that quantum mechanics makes predictions only about experiments with very large numbers of objects. If there were ten billion cats in exactly the same boxes, quantum mechanics would tell us that after an hour, half of them would be dead – and surely observation would confirm the truth of this statement. Perhaps for one cat the theory simply does not apply. A similar argument was made for electrons in the previous chapter. However, in fact, interpretation in terms of ensembles faces difficulties even when explaining the interference pattern in a simple double-slit experiment.

Moreover, interpretation in terms of ensembles is tantamount to abandoning quantum mechanics as a physical theory for describing a single object or single event. Since single events do occur (and even single electrons have been isolated), we should be able to talk about single quantum objects. In fact, quantum mechanics was formulated in relation to single objects, despite the paradoxes that this raises. We must be prepared for Schrödinger’s paradox and look for a way to resolve it. The alternative is to have no physics at all for single objects – and this is completely unacceptable.

Today, many physicists, when dealing with Schrödinger’s cat paradox, prefer to hide behind the anti-metaphysical philosophy of logical positivism. This philosophy grew out of the work of the Viennese philosopher Ludwig Wittgenstein, Tractatus Logico-Philosophicus, where he expressed his famous judgment: “What cannot be spoken about, one should remain silent about.” Following this rule, such physicists—we might call them the neo-Copenhagen School—claim that we should limit our discussion of reality to what is visible, rather than trying to assert the reality of something that cannot be observed. For them, the main argument is that we never see a coherent superposition. Is the unobserved cat half alive or half dead? They would say that such a question cannot be asked because it cannot be answered. Of course, this is sophistry. A question that cannot be answered directly can nevertheless be approached indirectly and the answer can be calculated based on consistency with what we can know directly. Moreover, complete avoidance of metaphysical questions is incompatible with the spirit of the original Copenhagen interpretation and with the views held by Bohr and Heisenberg.

According to Bohr, the Copenhagen Interpretation reduces the absurdity of the half-dead, half-alive cat through the principle of complementarity: coherent superposition is an abstraction; in the abstract, a cat can exist both alive and dead. This description is in addition to the description we give of a dead or living cat when we see it. According to Heisenberg, the coherent superposition—the half-living, half-dead cat—exists in transcendental potency. It is our observation that “collapses” the dual state of the cat into a single one.

How should we understand this concept of a half-living, half-dead cat existing in potency? The answer, which sounds like science fiction, was proposed by physicists Hugh Everett and John Wheeler. According to Everett and Wheeler, both possibilities are realized—a living cat and a dead cat—but they occur in different realities, or parallel universes. For every living cat we find in a box, our counterpart in a parallel universe opens our box’s counterpart to discover our cat’s dead counterpart. Observing the dual state of a cat causes the universe itself to split into parallel branches. It’s an intriguing idea, and one that some science fiction writers (notably Philip K. Dick) have capitalized on. Unfortunately, this is also an expensive idea. According to it, the amount of matter and energy would double every time an observation causes the universe to split apart. This offends our penchant for parsimony, which may be a prejudice but nevertheless serves as the cornerstone of scientific reasoning. Moreover, since parallel universes do not interact, this interpretation is difficult to test experimentally and is therefore scientifically useless. (In fiction, things are simpler. In Philip K. Dick’s novel The Man in the High Castle, parallel universes interact. Otherwise, how could there be a plot?)

Fortunately, there is a possible idealistic solution. Since our observation magically resolves the cat’s dichotomy, it must be us—our consciousness—that causes the collapse of the cat’s wave function. Material realists don’t like this idea because it makes consciousness an independent causal entity; to admit this would be to hammer nails into the coffin of material realism. Contrary to materialism, such famous scientists as John von Neumann, Fritz London, Edmond Bauer and Eugene Wigner agreed with this solution to the paradox.

Idealistic solution

According to the idealist solution, it is the observation of the conscious mind that puts an end to the live-or-dead cat dichotomy. Like Plato’s archetypes, coherent superpositions exist in a fairyland of transcendental order until we collapse them and bring them into the world of manifestation by our act of observation. In this case, we select one aspect from two, or many, resolved by the Schrödinger equation; Of course, it is a limited choice, subject to the general probability constraint of quantum mathematics, but it is a choice nonetheless.

Even if material realism is wrong, should we be quick to abandon scientific objectivity and invite consciousness into our science? One of the pioneers of quantum physics, Paul Dirac, once said that great breakthroughs in physics always involve the abandonment of some great prejudice. Perhaps it is time to abandon the bias of strict objectivity. Bernard D’España considers the objectivity allowed by quantum mechanics to be weak objectivity. Instead of the independence of events from the observer that strict objectivity requires, quantum mechanics allows for some intervention by the observer—but in such a way that the interpretation of events does not depend on any individual observer. This weak objectivity represents the invariance of events with respect to the observer: whoever the observer is, the event remains the same. Because individual measurements involve subjective choice, this principle is clearly a statistical one, and observer invariance only holds for large numbers of observations—which is nothing new. Having long ago accepted the probabilistic interpretation of quantum mechanics, we are now obliged to accept the statistical nature of some of our scientific principles – for example, the principle of causality. As cognitive psychology regularly demonstrates, we can certainly do science with weak objectivity defined in this way. We don’t really need strict objectivity.

The solution to Schrödinger’s paradox by the intervention of consciousness is the simplest – so simple that it is sometimes called the naive solution. However, many questions have been asked about this decision, and only by answering these questions can we overcome the charge of naivety.

Questions about the idealistic solution

One question you may still be asking is how can a cat be half alive and half dead? It can’t – if you think in terms of material realism. Material realism suggests that the state of a cat at any given time must be causally continuous in being either one or the other, or alive or dead. However, materialistic thinking is a consequence of assumptions of causal continuity and either/or descriptions of events. These assumptions are not necessarily true, especially when tested by quantum mechanics experiments.

The idealistic philosopher is not particularly concerned about the paradox of a cat being both alive and dead. In one story, a Zen master was shown a so-called dead man who was about to be buried. When asked whether a person was alive or dead, the Zen master replied, “I can’t tell.” How could he? According to idealism, the essence of man – consciousness – never dies. Therefore, it would be wrong to directly say that a person is dead. However, when a person’s body is being prepared for burial, it would be absurd to say that he is alive.

Is the cat alive or dead? When the Zen master Zeshu was asked, “Does a dog have Buddha nature?” he answered “mu.” Here again, to answer “no” would be wrong, since, according to the teachings of the Buddha, all beings have Buddha nature. Answering “yes” would also be risky, since Buddha nature must be realized and lived, not intellectually understood. Therefore, the master answered “mu” – neither yes nor no.

Apparently, when quantum mechanics states that at the end of an hour, Schrödinger’s cat is half alive and half dead, it assumes an idealistic philosophy similar to that of the Zen masters. How can it be? How can consciousness play a decisive role in shaping the reality of the physical world? Doesn’t this presuppose the supremacy of consciousness over matter?

If before we look inside the box Schrödinger’s cat is both alive and dead, but after we look inside it is in a single state (alive or dead), then simply by looking we must do something . How can a fleeting glance affect a cat’s physical condition? Realists ask these questions in an attempt to refute the idea that consciousness causes the collapse of a coherent superposition.

Yes, the idealistic solution does imply the action of consciousness on matter. However, this impact only poses a problem for material realism. In this philosophy, consciousness is considered an epiphenomenon of matter, and it seems impossible that an epiphenomenon of matter could affect the very tissue from which it is formed—essentially causing itself. This causal paradox is avoided by monistic idealism, in which consciousness is primary. In consciousness, coherent superpositions represent transcendental objects. They become immanent only when consciousness, through a process of observation, selects one of the many aspects of a coherent superposition, although its choice is limited by the probabilities that the quantum calculus allows. (Consciousness is law-abiding. The creativity of the cosmos comes from the creativity of its quantum laws, and not from arbitrary lawlessness.)

According to monistic idealism, objects are already in consciousness as incipient, transcendent, archetypal forms of possibility. Collapse does not consist of doing something to objects by measuring, but of choosing and recognizing the result of that choice. Take another look at the earlier illustration of the gestalt “My wife and mother-in-law” (Fig. 12). This illustration contains an overlay of images. When we see a wife (or mother-in-law), we do nothing with the picture. We simply choose and acknowledge our choice. The process of consciousness collapsing a wave function is something like this.

However, there are dualists who try to explain the action of consciousness in Schrödinger’s paradox, finding evidence of psychokinesis – the ability to move matter through the action of the mind. Eugene Wigner argues that if a quantum object can influence our consciousness, then consciousness must be able to influence a quantum object. However, the evidence for psychokinesis is insufficient and questionable. In addition, consideration of another paradox—the “paradox of Wigner’s friend”—essentially excludes a dualistic interpretation.

Wigner’s friend paradox

Suppose two people open the box containing Schrödinger’s cat at the same time. If the outcome of collapse is chosen by the observer, as the idealistic solution implies, then what if the two observers make different choices—wouldn’t that create a problem? If we say no, then only one of the observers can make the choice, and supporters of realism rightly consider this decision unsatisfactory.

In Wigner’s friend paradox, formulated by physicist Eugene Wigner, what happens is that instead of observing the cat himself, Wigner asks his friend to do so. His friend opens the box, sees the cat and then reports the results of his observation to Wigner. At this stage we can say that Wigner has just actualized a reality that includes his friend and the cat. There is a paradox here: was the cat alive or dead when Wigner’s friend observed it, but before he reported the result of the observation? To say that when Wigner’s friend observed the cat, its state did not collapse is to say that his friend was unconscious until Wigner asked him – that his friend’s consciousness could not decide whether the cat was alive or dead without prompting from Wigner. This is quite similar to solipsism, a philosophy that believes you are the only conscious being and everything else is a figment of your imagination. Why should Wigner be the privileged agent who is allowed to cause the cat’s state function to collapse?

Suppose we instead say that the collapse of the superposition causes the consciousness of Wigner’s friend. Doesn’t this open Pandora’s box? If Wigner and his friend look at a cat at the same time, whose choice will matter? What if two observers make different choices? If individual people determined the behavior of the objective world by their choices, then life would turn into absolute hell, since, as we know, subjective impressions are often contradictory. In such a case, the situation would be similar to that in which motorists moving from different directions would choose the color of the traffic light (red or green) at will. This argument is often considered the death blow to the solution of Schrödinger’s paradox by conscious intervention. However, it is only fatal to the dualistic interpretation. To understand why this is so, let’s look at Wigner’s paradox in more detail.

Wigner compared his paradoxical situation to one in which an inanimate device is used to make observations. When a mechanism is used, there is no paradox. There is nothing paradoxical or frustrating about a machine being in a state of uncertainty for any length of time, but experience tells us that the observation of a conscious being is crucial. Once a conscious being makes an observation, material reality becomes manifest in a single state. According to Wigner:

Therefore, a conscious being must have a different role in quantum mechanics than an inanimate measuring device… This argument assumes that “my friend” has the same kinds of impressions and sensations as I do – in particular, that after interacting with an object he is not in an unconscious state… It is not necessary to see a contradiction here from the point of view of orthodox quantum mechanics, and there is none if we believe that the alternative is meaningless – does my friend’s consciousness contain… the impression of what he saw [either a dead or a living cat]. However, to deny the existence of a friend’s consciousness to such an extent is undoubtedly an unnatural position, approaching solipsism, and few people in their hearts will agree with it.

It’s an insidious paradox, but Wigner is right. We need not say that as long as Wigner does not manifest his friend, the friend is in an unconscious state. Equally, we need not resort to solipsism. There is another alternative.

Wigner’s paradox arises only when he makes the unfounded dualistic assumption that his consciousness exists separately from the consciousness of his friend. The paradox disappears if there is only one subject, rather than separate subjects as we usually understand them. The alternative to solipsism is a single subject-consciousness.

When I observe, I see the whole world of manifestation, but this is not solipsism, since there is no individual seeing self as opposed to other selves. Erwin Schrödinger was right when he said: “Consciousness is the only thing for which there is no plural.” Etymology and spelling have preserved the uniqueness of consciousness. However, the existence of terms such as “I” and “mine” in language leads us into a dualistic trap. We consider ourselves separate because we talk about ourselves in this way.

In the same way, people get used to thinking about having consciousness, as in the question: “Does a cat have consciousness?” Only in material realism does consciousness represent something that can simply be possessed. Such a consciousness would be deterministic, not free, and would not be worth having.

The watched pot still boils

Let’s look at another complication in Schrödinger’s paradox. Let us assume that Schrödinger’s cat is itself a conscious being. The situation becomes even more critical if we assume that there is a person in the box with a radioactive atom, a bottle of poison and everything else. Then suppose that after an hour we open the box, and if he is still alive, we ask him whether he experienced the half-dead, half-alive state? He will answer – of course not! Think a little. What if we asked him if he felt alive the whole time? If this person is thoughtful enough, after some thought he will probably say no. You see, we are not aware of our body all the time. In fact, under ordinary circumstances, a person is very little aware of his body. Therefore, from the point of view of an idealistic interpretation, what happens can be described as follows. Over the course of an hour, the man became aware from time to time that he was alive. In other words, he observed himself. At these moments, his wave function collapsed, and, fortunately, each time the choice was the living state. Between these moments of wave collapse, its wave function expanded and became a coherent superposition of the dead and living states in a transcendental realm beyond experience.

You know how we see movies. Our brain-mind is not capable of distinguishing between individual still pictures running before our eyes at a speed of twenty-four frames per second. Likewise, what appears to be continuity to one observing oneself is in fact a mirage consisting of many discrete collapses.

This last argument also means that we could not save Schrödinger’s cat from the cruel outcome of radioactive decay by constantly looking at it, and thus continuously collapsing its wave function and keeping it alive. This is a noble impulse, but it is doomed to fail – for the same reason that a pot that is watched boils, although the proverb suggests otherwise. It’s good that the pot is being watched, because if we could prevent change by simply looking at an object, the world would be full of narcissists trying to avoid old age and death by meditating on themselves.

Erwin Schrödinger’s reminder should be taken into account: “Observations should be considered individual discrete events. There are gaps between them that we cannot fill.”

The solution to Schrödinger’s cat paradox tells us a lot about the nature of consciousness. By manifesting material reality, it makes a choice between alternatives; it is transcendental and one; and his actions elude our normal everyday perception. Of course, from a common sense perspective, none of these aspects of consciousness seem self-evident. Try to curb your disbelief and remember what Robert Oppenheimer once said: “Science is an extraordinary sense.”

Quantum collapse is a process of selection and recognition by a conscious observer; ultimately there is only one observer. This means we need to resolve another classic paradox.

When does the measurement end?

According to some realists, a measurement is complete when a classical measuring device, like the Geiger counter in Schrödinger’s cat cage, measures a quantum object; it completes when the counter clicks. Note that if we make this decision, then the paradox of the dual state of the cat does not arise.

This reminds me of a story. Two elderly gentlemen were talking, and one of them complained of chronic gout. Another said with some pride: “I don’t have to worry about gout; I take a cold shower every morning.” The gentleman with gout looked at him mockingly and replied, “So you get chronic cold showers instead!”

These realists try to replace the dichotomy of Schrödinger’s cat with the dichotomy of the quantum and classical levels. They divide the world into quantum objects and classical measuring instruments. However, such a dichotomy is untenable and completely unnecessary. We can state that all objects are subject to quantum physics (the unity of physics!), and at the same time satisfactorily answer the question: “When does measurement end?”

What is the definition of measurement? To put it slightly differently, when can we say that quantum measurement is finished? You can approach the answer historically.

Werner Heisenberg, who proposed the uncertainty principle, formulated a thought experiment that Bohr further refined. David Bohm recently described an experiment that I will use here. Let us assume that the particle is at rest in the plane of the microscope target and we analyze its observation from the position of classical physics. To observe a target particle, we point another particle at it (using a microscope), which is deflected by the target particle onto a photographic plate, leaving a mark on it. Based on the study of the trace and our knowledge of how the microscope works, we can, in accordance with classical physics, determine both the position of the target particle and the momentum imparted to it at the moment of deflection. Specific experimental conditions do not affect the final result.

In quantum mechanics, all this changes. If the target particle is an atom and if we look at it with an electron microscope in which an electron is deflected by the atom onto a photographic plate (Fig. 22), the following four considerations arise:

1. The deflected electron should be described both as a wave (while it moves from the object O to the image P) and as a particle (when it reaches P and leaves a trace T).

2. Due to this wave aspect of the electron, the image P gives us only the probability distribution of the position of the object O. In other words, the position is determined only within the limits of some uncertainty ∆x.

3. In the same way, Heisenberg argued, the direction of the trace T gives us only the probability distribution of the impulse O and, thus, determines the impulse only within the limits of uncertainty ∆p. Using simple mathematics, Heisenberg was able to show that the product of two uncertainties is greater than or equal to Planck’s constant. This is the Heisenberg uncertainty principle.

4. In a more detailed mathematical description, Bohr showed that the wave function of an observed atom cannot be determined separately from the wave function of the electron used to observe it. In reality, Bohr said, the wave function of the electron cannot be separated from the wave function of the photographic emulsion. And so on. It is impossible to draw an unambiguous dividing line in this chain.

Img. 22.
Bohr-Heisenberg microscope</strong”>

Despite the uncertainty in drawing the dividing line, Bohr felt that we must draw it due to the “necessary use of classical concepts in the interpretation of all correct measurements.” Bohr was reluctant to admit that the experimental setting should be described in purely classical language. It must be assumed that the dichotomy of quantum waves ends in the measuring device. However, as the philosopher John Schumacher has convincingly shown, all actual experiments contain a second built-in Heisenberg microscope: the process of seeing a trace in an emulsion involves the same kind of consideration as what led Heisenberg to the uncertainty principle (Fig. 23). Photons from the emulsion are amplified by the experimenter’s own visual apparatus. Can we ignore the quantum mechanics of our own vision? If not, then aren’t our brain-mind-consciousness inextricably linked to the process of measurement?

Img. 23.
Mechanics of vision. Another Heisenberg microscope in action?

Does a cat belong to the quantum or classical world?

When we think about it, it becomes clear that Bohr was replacing one dichotomy with another – the dichotomy of the cat with the dichotomy of a world divided into quantum and classical systems. According to Bohr, we cannot separate the wave function of an atom from everything else in the cell (various measuring instruments for determining the decay of an atom, such as a Geiger counter, a bottle of poison, and even a cat), and therefore the line we draw between the microworld and the macrocosm turns out to be completely arbitrary. Unfortunately, Bohr also talked about the need to recognize that measurement with a mechanism—a measuring device—resolves the dichotomy of the quantum wave function.

However, any macroscopic body is ultimately a quantum object; there is no such thing as a classical object unless we are willing to accept the vicious quantum/classical dichotomy in physics. It is true that in most situations the behavior of a macroscopic body can be predicted based on the rules of classical mechanics. (In such cases, quantum mechanics makes the same mathematical predictions as classical mechanics—this is the correspondence principle that Bohr himself discovered.) For this reason, we often roughly consider macroscopic bodies to be classical. However, the measurement process is not such a case, and the correspondence principle does not apply to it. Of course, Bohr knew this. In his famous debates with Einstein, Bohr often invoked quantum mechanics to describe macroscopic bodies in measurement to counter Einstein’s pointed objections to probability waves and the uncertainty principle.

As an example of the dispute between Bohr and Einstein, consider the situation of the double-slit experiment, but with one additional aspect. Let’s assume that before hitting the double slit, the electrons pass through a single slit in the diaphragm – its purpose is to accurately determine the initial position of the electrons. Einstein proposed installing this first slot on extremely sensitive springs (Fig. 24). He argued that if the first slit deflects an electron to the upper of the two slits, then, due to the principle of conservation of momentum, the first diaphragm will move down, and if the electron is deflected down to the lower of the slits, then the opposite will happen. Thus, measuring the recoil of the diaphragm will tell us which slit the electron actually passes through—information that is impossible from the point of view of quantum mechanics. If the first diaphragm had truly been classical, then Einstein would have been right. In defending quantum mechanics, Bohr pointed out that, ultimately, this diaphragm is also subject to quantum uncertainty. Therefore, when measuring its momentum, its position becomes uncertain. Bohr was able to demonstrate that this widening of the first slit effectively eliminated the interference pattern.

Img. 24.
Einstein’s idea: an initial slit on springs for a double-slit experiment. If, before passing through a partition with two slits (not shown), electrons pass through a slit in a diaphragm mounted on springs, is it possible to determine which slit the electron passes through without destroying the interference pattern?

However, let us further assume that the principle of complementarity operates and that sometimes a macroscopic device does acquire a quantum dichotomy (as the Bohr-Einstein controversy shows), but that at other times this does not happen – as in the case of a measuring device. This original idea, called macrorealism, comes from the brilliant physicist Tony Leggett, whose work led to the creation of a magnificent experimental device called SQUID (Superconducting Quantum Interference Detector).

Ordinary conductors conduct electricity, but always offer some resistance to the passage of electric current, which results in the loss of electrical energy in the form of heat. In contrast, superconductors allow current to flow without resistance. If you create an electric current in a superconducting circuit, then this current will flow almost forever – even without a source of energy. Superconductivity is due to a special correlation between electrons that spreads throughout the superconductor. Electrons require energy to escape from this correlated state, making the state relatively immune to the random thermal motion present in a normal conductor.

A SQUID is a piece of superconductor with two holes that almost touch at a point called the “weak link” (Figure 25). Let’s say we create a current in a loop around one of the holes. Current creates a magnetic field, just like any electromagnet, and magnetic field lines passing through a hole are also a common occurrence. In the case of a superconductor, what is unusual is that the magnetic flux—the number of field lines per unit area—is quantized; the magnetic flux passing through the hole is discrete. This gave Leggett his key idea.

Img. 25.
Will the flux line split between the two holes, showing quantum interference at the macroscopic level?

Let us assume that we create such a small current that there is only one flux quantum. Then we created a double-slit interference situation. If there is only one hole, then it is obvious that the quantum can be anywhere in it. If the link between two holes is too thick, the flow will be confined to only one hole. Is it possible, with a suitable size of the weak link, to create quantum interference so that the flux quantum is non-localized, being in both holes at the same time? If so, then quantum coherent superpositions clearly persist even at the level of macroscopic bodies. If no such delocalization is observed, then we can conclude that macroscopic bodies are indeed classical and do not admit coherent superpositions as their allowed states.

There is still no evidence of a violation of quantum mechanics in the case of SQUID, but Leggett stubbornly expects the collapse of quantum theory. At a recent conference, he said: “But at times, when the full moon shines brightly, I do what in the physics community may be the intellectual equivalent of becoming a werewolf: I wonder whether quantum mechanics is the complete and final truth about the physical universe… I am inclined to believe that somewhere between the atom and the human brain it [quantum mechanics] not only can, but must fail.”

He spoke like a true material realist!

Many physicists feel inclined to ask the same questions that inspire Leggett, so SQUID research continues. I suspect that one day they will provide evidence for quantum mechanics and show that quantum coherent superpositions are clearly present even in macroscopic bodies.

If we do not deny that ultimately all objects acquire a quantum dichotomy, then, as von Neumann first argued, if a chain of material mechanisms measures a quantum object in a state of coherent superposition, they all acquire the object dichotomy in turn, ad infinitum (Fig. 26). How to get out of the deadlock created by the von Neumann chain? The answer is amazing: jumping out of the system, out of the material order of reality.

Img. 26.
Von Neumann chain. According to von Neumann’s proof, even our brain-mind becomes infected with the cat dichotomy, so how does the chain end?

We know that observation by a conscious observer ends the dichotomy. It is therefore quite obvious that consciousness must act from outside the material world; in other words, consciousness must be transcendental—nonlocal.

Ramachandran’s paradox

If you are still concerned about the transcendence of consciousness, then you may enjoy considering the paradox that neuroscientist Ramachandran came up with.

Suppose that thanks to some super technology it is possible to record, using electrodes or something like that, everything that happens in the brain when external stimuli act on it. You can imagine that from these data and with the help of some super-mathematics you can obtain a complete and detailed description of the state of the brain in the situation of the action of a given stimulus.

Suppose the stimulus is a red flower; you show it to several people, collect the data, analyze it and get a set of brain states corresponding to the perception of a red flower. You would expect that, barring minor statistical variations, you would get essentially the same description of the state each time (something like there was a reaction in certain cells in a certain area of ​​the brain involved in color perception).

You might even imagine yourself using super technology to record and analyze your own brain data (while seeing a red flower). The brain state you find in yourself should not be any noticeably different from all others.

Consider this interesting twist to the experiment: You have no reason to suspect that the description of everyone else’s brain states is incomplete (especially if you fully believe in your superscience). And at the same time, in relation to your own brain state, you know that something is missing – namely, your role as an observer – your awareness of the experience corresponding to your brain state, the actual conscious perception of red. Your subjective experience cannot be part of the objective state of the brain, because in such a situation, who would observe the brain? A famous Canadian neurosurgeon was similarly puzzled when contemplating the prospect of operating on his own brain: “Where is the subject and where is the object if you operate on your own brain?”

There must be a difference between your brain as an observer and the brains of those you observe. The only alternative conclusion is that the brain states you construct, even with superscience, are incomplete. Since your brain state is incomplete, and other people’s brain states are identical to yours, then they must also be incomplete, because they do not take into account consciousness.

For material realists this is a paradox, since from their point of view none of the above solutions are desirable. The material realist will not be willing to give special privileges to the individual observer (which would be tantamount to solipsism), but will also be reluctant to admit that any achievable description of the state of the brain using materialist science would be ipso facto incomplete.

An important clue is provided by the neurosurgeon’s question – where is the subject and where is the object if you operate on your own brain? The essence of the problem is conveyed by the expression: “What we are looking for is what is looking for.” Consciousness presupposes paradoxical self-reference—the taken-for-granted ability to relate to ourselves separately from our surroundings.

Erwin Schrödinger said: “Unconsciously, and without being strictly consistent in this matter, we exclude the Subject of Knowledge from the sphere of nature that we are trying to understand.” A theory of quantum measurement that dares to invoke consciousness in matters of quantum objects must deal with the paradox of self-reference. Let’s clarify this concept.

When does the measurement end? (Summary)

A subtle criticism can be made from the statement that transcendental consciousness causes the collapse of the wave function of a quantum object: the consciousness causing the collapse of the wave function could be the consciousness of the eternal, omnipresent God, as in the following humorous passage:
There was once a man who said: “To God
It must seem exceedingly strange
If He discovers that this tree
continues to exist
when no one is around.”
Dear sir, your surprise is strange,
I am always nearby
and that is why the tree will continue to be,
as it is observed by I,
yours truly, God.

However, an omnipresent God causing the collapse of the wave function does not resolve the measurement paradox, since we can ask, “At what point is measurement finished if God is always looking?” The answer is crucial: measurement is not complete without the inclusion of immanent awareness. The most familiar example of immanent awareness is, of course, the awareness of the mind-brain of a human being.

When is the measurement completed? When transcendental consciousness causes the collapse of the wave function through the immanent mind-brain looking with awareness. This formulation is consistent with our ordinary observation that there is never an experience of a material object without an accompanying mental object, that is, the thought “I see this object,” or at least without awareness.

Note that a distinction must be made between consciousness with awareness and without awareness. Wave function collapse occurs in the first case, but not in the last. In psychological literature, consciousness without awareness is called unconscious.

Of course, in the idea that immanent awareness is required to complete a dimension, there is a certain causal circle, since without the completion of a dimension there can be no immanent awareness. Which comes first, awareness or measurement? What is the root cause? Are we facing an unanswerable “chicken or the egg” question?

One Sufi story has a similar connotation. One night, Mullah Nasreddin was walking along a deserted road when he noticed a group of horsemen approaching. Mulla got nervous and ran. The horsemen saw him running and galloped after him. Now the mullah was really scared. Having reached the walls of the cemetery and driven by fear, he jumped over the wall, found an empty grave and lay down in it. The horsemen saw him jump over the wall and followed him into the cemetery. After a little searching, they found the mullah, fearfully looking up at them.
“Something happened? – the horsemen asked the mullah. – Can we help you somehow? Why are you here?”
“Well, it’s a long story,” answered the mullah. “In short, I am here because of you, and I can see that you are here because of me.”

If only one order of reality is imposed on us – the physical order of things, then this is a genuine paradox for which there is no solution within the framework of material realism. John Wheeler called the circular nature of quantum measurement the “circle of meaning.” This is a very insightful description, but the real question is who is reading the meaning. There is no paradox here only for idealism, since consciousness acts from outside the system and completes the cycle of meaning.

This solution is similar to the solution to the so-called prisoner’s problem, an elementary game theory problem. You plan to escape from your prison cell through a tunnel dug with the help of your friend (Fig. 27). Obviously, your escape will be much easier if you and your friend are digging from opposite sides of the same camera corner; however, you cannot communicate and the cell has six angles from which to choose. The chances of escaping don’t look too good, do they? But think a little about the shape of your camera and you will realize that you will most likely decide to dig in corner number 3. Why? Because this is the only corner that looks different (concave) from the outside. So you would expect your friend to start digging here. Likewise, only corner number 3 is concave from the inside, so your friend will probably expect you to start digging there too.

Img. 27.
Prisoner’s dilemma: which angle to choose?

But what is your friend’s motivation for digging in this corner? It is you! He imagines you choosing this angle for the same reason you imagine him choosing it. Note that in this case we cannot establish any causal sequence and therefore no simple hierarchy of levels. Instead of a linear causal hierarchy, we have a circular causal hierarchy. Nobody chose the plan. Instead, the plan was a joint creation driven by a higher goal—the prisoner’s escape.

Douglas Hofstaedter called this type of situation a complex hierarchy—a hierarchy that is so intricate that it is impossible to distinguish higher and lower levels on the hierarchical totem pole. Hofstadter suggests that self-reference may arise from such a complex hierarchy. I suspect that in the brain-mind situation, in which consciousness causes the wave function to collapse, but only when awareness is present, our immanent self-reference comes from a complex hierarchy. The von Neumann chain ends precisely with the observation of a self-correlating system.

Irreversibility and the arrow of time

When does the measurement end? Idealism states that it ends only when self-referential observation has occurred. In contrast, some physicists argue that the measurement ends when the detector detects a quantum event. How does the detector differ from the previous measuring device? These physicists claim that detection by the detector is irreversible.

What is irreversibility? There are some processes in nature that can be called reversible, since observing these processes in reverse order, it is impossible to determine the direction of time. An example is the movement of a pendulum (at least for a short period of time): if you film its movement and then run it in the opposite direction, you will not find any visible difference. In contrast, filming an irreversible process cannot be replayed without revealing its secret. For example, suppose that while you are filming the movement of a pendulum on a table, you are also filming a cup that falls to the floor and breaks. When you rewind the movie, the pieces that fly up from the floor and become a whole cup again will reveal your secret – that you are rewinding the movie.

To understand the difference between a reversible meter and a detector, consider the following example. Photons have a characteristic called polarization, which can take two meanings: it is an axis directed (or polarized) in only one of two mutually perpendicular directions. Polarized sunglasses polarize normal non-polarized light. They transmit only those photons whose polarization axis is parallel to the polarization axis of the glasses. You can check this by placing two polarized glasses perpendicular to each other and looking through them. You will only see darkness. Why? Because one polarized glass polarizes photons, say, vertically, while another allows only horizontally polarized photons to pass through. In other words, both glasses together act as a double filter that blocks out all the light.

A photon polarized at 45° is a coherent superposition of half vertically and half horizontally polarized states. If such a photon passes through a polarization box with vertically and horizontally polarized channels, then it randomly appears in either the vertically polarized or horizontally polarized channel. This can be judged by the readings of detectors placed behind each of the channels (Fig. 28, a).

Now suppose that in the setup shown in Fig. 28, a, we will place a polarizer with a polarization angle of 45° between the polarizing box and the detectors (Fig. 28, b). It turns out that the photon restores its original state of polarization at an angle of 45° – a state of coherent superposition; he is reborn. Thus, a polaroid alone is not enough to measure photons – since the photons still retain their potential to become a coherent superposition. The measurement requires a detector in which irreversible processes occur, such as a fluorescent screen or photographic film.

Img. 28.
Experiments with photons polarized at an angle of 45°

If you think in terms of time reversal, then the motion of 45° polarized photons that pass through a polarization box and then again through a 45° polarizer is time reversible. However, if the photons are detected by some detector with an irreversible process, then by imagining this process in reverse, you are able to distinguish between forward and backward motion.

Remember the story about the scene filmed for silent films. The heroine was supposed to be tied to the tracks in front of an approaching train. According to the plot of the film, the heroine had to be saved – the train stopped at the last moment. Since the actress (for obvious reasons) did not want to risk her life, the director filmed the entire scene backwards – starting with the moment when the actress is tied to the rails, and the train stands motionless next to her. Then the train began to move backwards. But what do you think audiences saw when the film was played backwards? In those days, trains were driven by coal-fired steam locomotives. In the film, which was run in reverse, the smoke entered the locomotive’s chimney instead of coming out of it, thereby revealing the film’s secret. Smoke formation is an irreversible process.

Does this mean that the solution to the problem of quantum measurement is close – and without the assumption of the participation of consciousness? We only need to recognize the irreversibility of certain measuring instruments called detectors, and then perhaps we can break free of the von Neumann chain. Once the detectors have been triggered, the coherent superposition can no longer be reconstructed and can therefore be said to have truly ended.

But is this really so? Is the detector sufficient to complete the von Neumann chain? Von Neumann himself answers no. The detector must become a coherent superposition of the needle readings, for the simple reason that it also obeys quantum mechanics. The same is true for any subsequent measuring device – reversible or “irreversible”. The von Neumann chain continues.

The point is that the quantum Schrödinger equation is time reversible: it does not change when the sign of time changes. As the mathematician Jules Henri Poincaré showed, the behavior of any macroscopic body subject to a time-reversible equation cannot be truly irreversible. Therefore, a generally accepted point of view is emerging that absolute irreversibility is impossible; The apparent irreversibility that we observe in nature is due to the low probability of reversing the evolutionary path of a macroscopic body to the initial configuration, which has greater relative order.

Accounting for irreversibility provides an important lesson. Although ultimately all objects are quantum objects, the apparent irreversibility of some macro objects allows us to make a rough distinction between the classical and the quantum. We can say that a quantum object is restored, while the restoration time of a classical object is extremely long. In other words, we can say that while quantum objects do not have a noticeable retention of their history – they do not have memory, classical objects – for example, detectors – have memory, in the sense that it takes a long time for the memory to be erased.

Another important question arises: if there is no absolute irreversibility in the movement of matter, then how does the idealistic interpretation cope with the idea of ​​unidirectional flow of time, the arrow of time? According to the idealist interpretation, in the transcendental realm time is a two-way street, showing signs of only approximate irreversibility for the movement of increasingly complex objects. When consciousness collapses the brain-mind wave function, it exhibits the unidirectional time we observe. Irreversibility and the arrow of time enter nature through the process of collapse itself—the quantum dimension—as physicist Leo Szilard suspected many years ago.

Apparently, the irreversibility of detectors does not solve the measurement problem. Such a solution can only be approached if we are willing to accept irreversibility in the form of disorder even more fundamental than quantum mechanics. There is a proposal to do just that.

Suppose that matter is fundamentally disordered and that the disordered behavior of the substrate of particles, through random fluctuations, gives rise to approximately ordered behavior, which we can call quantum. If this were true, then quantum mechanics itself would be an epiphenomenon—like all other ordered behavior. There is no experimental evidence to support this kind of theory, although if it could be proven it would be an ingenious solution to the measurement problem. However, some physicists still admit that there is a hidden underlying environment that causes randomness; They draw an analogy with the random movement of molecules that causes the random movement of pollen particles in water visible through a microscope (called Brownian motion). However, the assumption of an underlying environment is inconsistent with the Aspect experiment unless it involves nonlocality. And within the framework of material realism, it is difficult to accept non-local Brownian motion.

Nine Lives

Stephen Hawking says, “Whenever I hear about Schrödinger’s cat, I want to grab a gun.” Almost every physicist experiences a similar impulse. Everyone wants to kill the cat—that is, the cat paradox—but it apparently has nine lives.

In its first life, the cat is treated statistically, as part of an ensemble. The cat is offended (since this interpretation robs it of its distinctiveness), but is unharmed.

In its second life, philosophers of macrorealism saw the cat as an example of the quantum/classical dichotomy. The cat refuses to exchange its life/death dichotomy for yet another dichotomy.

In the third life, the cat is presented with irreversibility and randomness, but the cat says – prove it.

In the fourth life, the cat encounters hidden variables (the idea that its state never becomes dualistic, but is, in fact, entirely determined by hidden variables), and what happens remains hidden.

In the fifth life, representatives of the neo-Copenhagen school try to get rid of the cat using the philosophy of logical positivism. By most accounts, the cat remains unharmed.

In the sixth life, the cat encounters multiple worlds. Who knows, maybe she died in some universe, but as far as we can tell, not in this one.

In the seventh life, the cat encounters Bohr and his principle of complementarity, but is saved by the question: what constitutes a dimension?

In the eighth life, the cat comes face to face with consciousness (of the dualistic variety), but is saved by Wigner’s friend.

Finally, in the ninth life, the cat finds salvation in an idealistic interpretation. This ends the story of the nine lives of Schrödinger’s cat.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 7. I CHOOSE, THEREFORE I AM

We have not yet faced the important question: what is consciousness? And how to distinguish consciousness from awareness?

Alas, consciousness is not easy to define. The English word consciousness comes from two Latin words: the verb scire , meaning “to know,” and the preposition sit, meaning “with.” Thus, consciousness etymologically means “to know with.”

Moreover, the Oxford English Dictionary gives not one but six definitions of the word “consciousness”: shared or common knowledge; internal knowledge or belief, especially regarding one’s own ignorance, guilt, shortcomings, etc.; the fact or state of consciousness or awareness of something; the state or power of being conscious as a condition or concomitant of thinking, feeling and will; the sum of all impressions, thoughts and feelings that make up a person’s conscious existence; a state of awareness considered as the normal state of healthy waking life.

None of these definitions are completely satisfactory, but taken together they give a rough idea of ​​what consciousness is. Imagine a situation in which each of these definitions applies (we will denote each of them with a subscript number from 1 to 6). A bouquet of roses is delivered to you. Both you, the messenger, and the sender share a consciousness regarding this gift. It is in your consciousness that you know the history, associations and meaning of roses, and what they mean as a gift to you (and in that consciousness you may or may not value the gift). Your sensory experience of roses is in your consciousness, through which you are able to smell their scent, see their color and feel their thorns. However, it is your consciousness that gives you the ability to assign meanings, consider relationships, and make decisions related to a gift (such as accepting or rejecting roses). Your consciousness is what makes you unique and different from your lover and from every other person who reacts in one way or another to the gift of roses. It is only because of your consciousness that you are able to accept roses at all or experience and demonstrate any of the above states of consciousness.

Even this analysis of the word “consciousness” does not exhaust its meaning. Consciousness has four different aspects. First, there is the field of consciousness, sometimes called the field of mind or the global workspace. This is what I called awareness. Secondly, there are objects of consciousness, such as thoughts and feelings, that arise and disappear in this field. Thirdly, there is a subject of consciousness – the experiencer and/or witness. (In reality, dictionary definitions refer to the subject of consciousness, or the conscious self with which we identify.) Fourth, in idealist philosophy consciousness is understood as the basis of all existence.

The everyday definition of consciousness equates it to conscious experience. Talking about the subject of consciousness without talking about experience is like talking about a ballet scene without ballet. Note that the concept of conscious experience is not limited to waking consciousness. Dreaming is a conscious experience, although different from the experience of the waking state. The states that we experience in meditation, under the influence of psychoactive substances, in hypnotic trances – all such altered states of consciousness include experience.

Common sense also tells us that conscious experience has many attendant circumstances, both internal and external. For example, as I type this page, I observe my mind while my fingers tap the keys of the typewriter. I’m thinking: How good is this page? Should I change this phrase? Am I explaining too little or too much? And then I hear a knock on the door. I scream – who’s there? No answer. I have to make a choice – either shout again, louder, or get up and open the door.

It’s simple with external accompanying circumstances. I don’t identify with my fingers, even when they do things that are useful to me, like typing this page. Few of us would identify consciousness with sensations, sensory impressions or motor actions. Can you imagine saying, “I am my walk to the door?” Of course not. Common sense tells us that the external concomitant circumstances of conscious experience are not fundamental elements of consciousness.

When it comes to the inner stuff of the mind—thoughts, feelings, choices, etc.—things become much less clear. For example, many people – following Descartes – identify themselves with their thoughts: I think, therefore I exist. For others, consciousness is synonymous with feelings: I feel, therefore I am. Some of us can even identify with the ability to make choices. For example, Nietzsche equates being with will.

Science is different from common sense: we turn to science when common sense fails. However, turning to psychology does not help. As the eminent cognitive scientist Ulrik Neisser said: “Psychology is not ready to take on the problem of consciousness.” Fortunately, the physics is ready. This means a return to quantum theory and the problem of measurement, which gave rise to the very discussion of the problem of consciousness.

An idealistic solution to Schrödinger’s cat paradox requires the consciousness of the observing subject to select one aspect from the multi-aspect coherent superposition of the living and dead cat, thus deciding its fate. The subject is the one who chooses. This is not cogito, ergo sum, as Descartes believed, but opto, ergo sum: I choose, therefore I am.

The mind and the laws of the mind are hidden in the night.
God said, “Let there be Descartes,” and there was light.
This didn’t last long. The devil shouted: “Hey!
Schrödinger’s cat is here! Restore the status quo!”

(We, of course, apologize to A. Pope.)

I know that classical physicists will shake their heads in disapproval, since they believe that in our deterministic world there is no freedom of choice, or free will. Because of their assumption of causal determinism, they tried to train us to believe that we are material machines. Suppose we forget for a while what we have been taught. In the end, our hypothesis allowed us to resolve the Schrödinger’s cat paradox.

In the same spirit of inquiry we ask: what then? In response, the door opens. No matter how fascinated we are by thoughts and feelings, they come from old, unchanging, learned contexts. Is the same true for free will? Our choices set the context for our action, and therefore when we choose, the possibility of a new context arises. It is this possibility of leaving the old context and moving to a new one at a higher level that makes our choice free.

There is a special language specifically for describing a situation of this kind—the hierarchical structure of context levels. This language, known as logical type theory, was originally developed by Bertrand Russell to solve problems arising in set theory. Russell’s basic idea was that a set made up of members of a set is of a higher logical type than the members themselves, since the set defines the context for thinking about the members. Likewise, the name of a thing, which defines the context of the thing it describes, is of a higher logical type than the thing itself. Thus, of the three internal concomitant circumstances of conscious experience, choice really stands apart – it belongs to a higher logical type than thoughts and feelings.

Does this mean that it is the ability to choose that makes conscious for us the experiences we choose? At every moment we are literally faced with a myriad of alternative possibilities. By choosing from them, we realize the direction of our formation. Thus, our choices and recognition of choices define our selfhood. The main question of self-awareness is to choose or not to choose.

The idea that choice is the determining factor of self-awareness has some experimental support. Data from cognitive psychology experiments show that in response to unconscious perception of stimuli, thoughts and feelings arise, but not choices. According to the data described in the next section, we seem to make choices only if we act consciously—with awareness of ourselves as subjects.

This raises the question of what it means to act without awareness—the question of the unconscious. What is our unconscious? The unconscious is that for which there is consciousness, but there is no awareness. Note that there is no paradox here, since in the philosophy of idealism consciousness is the basis of existence. It is omnipresent, even if we are unconscious.

The confusion associated with the term “unconscious perception” is partly due to the historical features of its etymology. It is our conscious self that is unaware of certain things most of the time and is unaware of anything in deep dreamless sleep. In contrast, the unconscious appears to be aware of everything, all the time. It never sleeps. That is, it is our conscious self that does not have the consciousness of our unconscious, but our unconscious is conscious – these two terms are confused. You can read more about this in Daniel Goleman’s book, Vital Lies , Simple Truths.

So, when we talk about unconscious perception, we are talking about events that we perceive, but we are not aware that we are perceiving.

Experiments with unconscious perception

I know this sounds strange. How can a phenomenon called unconscious perception exist? Isn’t perception synonymous with awareness? The authors of the Oxford English Dictionary clearly believe this to be the case. However, new evidence from cognitive psychology points to a distinction between the concepts of perception and awareness.

The first experiments involved two monkeys. Researchers Nick Humphrey and Lewis Weiskrantz removed areas of the monkeys’ cortex associated with vision. Because the cortical tissue does not regenerate, it was assumed that these monkeys would remain permanently blind. However, it turned out that their vision was gradually restored – enough for the researchers to be convinced that the monkeys could see.

One of the monkeys, Helen, was often taken for walks on a leash. She gradually learned to do some things unusual for a creature that should have been blind. For example, she could climb trees. She also took her favorite food when it was close enough to grab it, but ignored it when it was out of reach. Helen obviously saw, but with what?

It turned out that there is a secondary pathway from the retina to a structure in the hindbrain called the superior colliculus . This visual pathway allowed Helen to see through what the experimenters called “blindsight.” Nick Humphrey accidentally came across a man who had “blindsight”. A disorder in the cerebral cortex left him blind in the left visual field of both eyes. Now experimenters had the opportunity to ask the subject what was happening in the mind when performing certain tasks using blind vision. The answers were strange.

For example, if this person was shown a light on the left side, to which he was blind, he could accurately point to its source. In addition, he could distinguish a cross from a circle and vertical lines from horizontal ones when they were presented in the left, blind visual field. But when asked how he saw these things, he insisted that he did not see them. He claimed that he was just guessing – even though the accuracy of his answers far exceeded what could have been the result of random guesses.

What does all this mean? There is now general agreement among cognitive psychologists that blindsight is an example of unconscious perception—perception without awareness of it. Thus, perception and awareness do not necessarily seem to be inextricably linked.

Further physiological and psychological evidence of unconscious perception came from studies conducted in the United States and Russia. The researchers measured the subjects’ brain electrical responses to various subthreshold signals. The responses were typically stronger when a meaningful picture, such as a bee, flashed on the screen for one millisecond than when a more neutral picture, such as an abstract geometric shape, was presented. (None of the subjects were obviously mathematicians.) In addition, when subjects were asked to report what words came to mind after these subthreshold signals, meaningful pictures produced words that were clearly related to the picture presented. For example, a picture of a bee evoked words such as sting and honey. In contrast, the abstract geometric figure evoked virtually nothing related to the object. There was clearly a perception of the image of a bee, but there was no conscious awareness of that perception.

The popular press seized on these experiments, declaring them experimental proof of Sigmund Freud’s idea of ​​the unconscious, which had captivated the scientific world in the late 19th and early 20th centuries. However, what is our unconscious? The unconscious is something for which there is consciousness (as the basis of being), but there is no awareness and no subject. Thus, in the case of unconscious perception, we are talking about events that we perceive (that is, events that are perceived as stimuli and processed by the brain as such) but are not aware of this perception. In contrast, conscious perception involves the perception of stimuli, their processing, and awareness of the perception.

The phenomenon of unconscious perception raises a crucial question. In unconscious perception, is any of the three usual concomitants of conscious experience (thoughts, feelings, and choices) absent? The experiment with subthreshold signals suggests that thinking is present, since as a result of the unconscious perception of the image of a bee, the subjects thought about the words “sting” and “honey”. We obviously continue to think in our unconscious, and unconscious thoughts influence our conscious thoughts.

With regard to feelings, important evidence has been provided by experiments with split-brain patients. In these patients, the left and right hemispheres of the brain were surgically separated, except for cross-connections in the hindbrain centers associated with emotions and feelings. When one patient’s right hemisphere was presented with a photograph of a naked man among a series of geometric patterns, she blushed, thereby demonstrating her embarrassment. However, when asked why she was blushing, she denied being embarrassed. She had no conscious awareness of these inner feelings and therefore could not explain why she blushed. Thus, feelings are also present in unconscious perception, and unconscious feelings can cause inexplicable conscious feelings.

Finally, we ask whether choice can also take place in unconscious perception? To find out, we must send an ambiguous signal to the mind-brain, suggesting the possibility of alternative reactions. In a related cognitive experiment, psychologist Tony Marcel used polysemic words that have more than one meaning. His subjects watched a series of three words flash on a screen in succession at intervals of 600 ms, or 1.5 s . 6 Subjects were asked to press a button when they consciously recognized the last word in the series. The original purpose of the experiment was to use the subject’s reaction time as a measure of the relationship between the congruence (or lack thereof) between words and the meanings assigned to words in series such as hand-palm-wrist (congruent), watch-palm-wrist ( not creating a preset), tree-palm-wrist (incongruous), and watch-ball-wrist (not connected). For example, one might expect that presenting the word “hand” before the word “palm” would create a preconditioning for the perception of the word’s hand-related meaning (i.e., “palm”), which would then improve the subject’s reaction time during recognition third word, “wrist” (conformity). If the tuning word had been preceded by “tree,” then the word palm would have been assigned the lexical meaning of palm (tree), and recognition of the meaning of the third word, “wrist,” would have taken longer. This is the result that was actually obtained.

However, when the middle word was masked with a pattern so that the subject saw it unconsciously but not consciously, there was no longer any noticeable difference in reaction time between the consistent and incongruent series. This should be surprising since both meanings of the ambiguous word were likely available to the subject, regardless of the priming context, but neither was favored over the other. Choice appears to be a matter of conscious experience, but not of unconscious perception. Our subjective consciousness arises when a choice is made: we choose, which means we exist.

It is suitable. When we do not choose, we are not aware of our perceptions. Therefore, a person with “blindsight” denies that he sees anything when he avoids an obstacle. A woman with a split cortex blushes but denies feeling embarrassed.

Cognitive psychology may eventually be able to help explain consciousness—especially if it is used to test ideas based on the quantum theory of the subject/self. Both quantum theory and these experiments show that there is a scientific basis for the Western tradition’s emphasis on freedom of choice as a central premise of human experience.

Note that if the quantum explanation of Marcel’s experiment is correct, then this experiment indirectly demonstrates the existence of coherent superpositions in our mind-brains. Before choice, the state of the mind-brain is uncertain – like the state of Schrödinger’s cat. In response to an ambiguous word, the mind-brain state becomes a coherent superposition of two states, each corresponding to a different meaning of the noun palm: a tree or part of a hand (palm). Collapse (reduction) consists of choosing between these states. (Due to conditioning, there may be some bias toward one meaning. For example, a California resident might have a slight preference for the meaning of palm as a tree. In this case, the weighted probability of the two possibilities would not be equal, but would favor the biased meaning However, there would still be a non-zero probability for the other value and it would still be a matter of choice.)

I choose, therefore I am. Remember also that in quantum theory the subject who chooses is the singular universal subject, not our personal ego or self. Moreover, as the experiment discussed in the next chapter shows, this choosing consciousness is nonlocal.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 8. EINSTEIN-PODOLSKY-ROSEN PARADOX

The idealistic interpretation of the collapse of the quantum wave function rests on the nonlocality of consciousness. So we need to ask whether there is any experimental evidence for nonlocality. We’re lucky. In 1982, Alain Aspect and his collaborators at the University of Paris-Sud conducted an experiment that convincingly demonstrated quantum nonlocality.

In the 1930s Einstein helped create the paradox, now commonly known as the EPR paradox, to prove the incompleteness of quantum mechanics and support realism. Given Einstein’s philosophical beliefs, EPR might well stand for “Einstein in Support of Realism.” Ironically, the paradox turned out to be a blow to realism – at least to material realism – and Aspect’s experiment played no small role in this.

Recall Heisenberg’s uncertainty principle: at any given time, only one of two complementary variables—position or momentum—can be measured with absolute certainty. This means that we can never predict the trajectory of a quantum object. Together with two of his collaborators, Boris Podolsky and Nathan Rosen, Einstein came up with a scenario that seemed to contradict this uncertainty.

Imagine that two electrons—let’s call them Joe and Mo—interact with each other for some time and then stop interacting. These electrons are, of course, identical twins, since electrons are indistinguishable. Suppose that when Jo and Mo interact, their distances from some source along a certain axis are equal to x J and x M , respectively (Fig. 29). The electrons move and, therefore, have angular momentum (momentum). We can designate these pulses (along the same axis) as J and M . From quantum mechanics it follows that we cannot simultaneously measure both p J and x J , or both p M and x M due to the uncertainty principle. However, quantum mechanics allows us to simultaneously measure their distance from each other (X = x J – x M ) and their total momentum (P = p J + p M )

Img. 29.
Correlation of Joe and Mo in EPR. The distance between them, x J – x M , always remains the same, and their total momentum is always equal to p J + p m

Einstein, Podolsky and Rosen argued that when Joe and Mo interact, they become correlated because, even if they later stop interacting, measuring Joe’s position (x J ) allows us to calculate exactly where Mo is – the x value of M – (since x M = x J – X , where X is the known distance between them). If we measure p J (Jo’s momentum), we can determine p M (Mo’s momentum), since M = P – J , and P is known. Thus, by performing a proper measurement of Joe, we can determine either the position or the momentum of Mo. However, if we take measurements of Joe when Joe and Mo are no longer interacting, then those measurements probably have no effect on Mo. Thus, the position and momentum of Mo must be available simultaneously.

The EPR conclusion stated that a correlated quantum object (Mo) must simultaneously have certain values ​​of both position and momentum. This conclusion supported realism, since we could now, in principle, determine Mo’s trajectory. On the contrary, he seemed to seriously compromise quantum mechanics, since it agrees with the idealism that the trajectory of a quantum object cannot be calculated, since the trajectory does not exist – only possibilities and observable events exist!

Einstein argued that if the trajectory of a correlated quantum object is predictable in principle, but quantum mechanics is unable to predict it, then there is something wrong with quantum mechanics. Einstein’s favorite conclusion from this dilemma was that quantum mechanics is an incomplete theory. Her description of the states of two correlated electrons is incomplete. Thus, he indirectly supported the idea of ​​​​the existence of hidden variables – unknown parameters that control electrons and determine their trajectories.

Physicist Heinz Pagels described the concept of hidden variables this way: “If we imagine that reality is a deck of cards, then quantum theory can only predict the probability of cards being dealt to different players. If there were hidden variables, it would be like looking at a deck and predicting the individual cards for each player.”

Einstein supported the idea of ​​deterministic hidden variables in order to demystify quantum mechanics. Remember, he was a realist. For Einstein, probabilistic quantum mechanics implied a playing God, and he believed that God does not play dice. He believed it was necessary to replace quantum mechanics with some kind of hidden variable theory in order to restore deterministic order to the world. Unfortunately for Einstein, the difficulty posed to quantum mechanics by EPR analysis could be resolved without resorting to hidden variables, as Bohr was the first to show. Bohr is reported to have told Einstein, “Don’t tell God what to do.”

In order to revive trajectories and therefore material realism, Einstein, Podolsky and Rosen started from the doctrine of locality. Recall that locality is the principle that all interactions are mediated by signaling across spacetime. Einstein and his colleagues tacitly assumed that measuring the position (or momentum) of the first electron (which we called Joe) could be done without affecting the second electron (Mo), since the two electrons were separated in space and did not interact through local signals during the measurement. This lack of interaction is what we usually expect for material objects, since the theory of relativity, which limits the speed of propagation of any signals to the speed of light, prohibits instantaneous interaction at a distance, or nonlocality.

The main question is one of separability: are correlated quantum objects separable when there is no local interaction between them, as is undoubtedly the case with objects obeying classical physics?

Why is the EPR result considered a paradox? Einstein’s principle of separability forms an integral part of the philosophy of material realism, which Einstein defended until the end of his life. This philosophy considers physical objects to be real and independent of each other and of their measurement or observation (the doctrine of strict objectivity). However, in quantum mechanics it is difficult to support the idea of ​​the reality of physical objects independent of the measurements we make on them. Thus, Einstein was motivated by a desire to discredit quantum mechanics and restore material realism as the basic philosophy of physics. The EPR paradox states that we must choose between locality (or separability) and the completeness of quantum mechanics, which means there is no choice at all, since separability is required.

But is it? The answer is a resounding “no!”, because in fact, the resolution of the EPR paradox lies in the recognition of the complete inseparability of quantum objects. A measurement of one of two correlated objects affects the second. This was essentially Bohr’s answer to Einstein, Podolsky and Rosen. When one object (Joe) of a correlated pair collapses into the momentum state p J , the wave function of the other (Mo) also collapses (to the momentum state P – p J ), and we can say nothing more about Mo’s position. And when Joe collapses as a result of measuring position at x J , Mo’s wave function also collapses immediately, corresponding to position x – X, and we can say nothing more about its momentum. Collapse is non-local, just as correlation is non-local. In EPR, correlated objects have a non-local ontological connection, or inseparability, and have an instantaneous influence on each other, not mediated by signals – no matter how difficult it is to believe from the point of view of material realism. Separability is the result of collapse. Only after the collapse are there independent objects. Thus, the EPR paradox forces us to recognize that quantum reality must be a non-local reality. In other words, quantum objects should be thought of as objects in potency that define a non-local realm of reality that transcends local space-time and is therefore beyond the jurisdiction of Einstein’s speed limits.

Although Bohr understood inseparability, he was reluctant to talk about quantum metaphysics. For example, he was not very precise about what he meant by measurement. From a completely idealistic point of view, we say that measurement always means observation by a conscious observer in the presence of awareness. Thus, the lesson of the EPR paradox seems to be that a correlated quantum system has the property of a certain inextricable integrity that includes an observing consciousness. Such a system has an innate integrity that is non-local and transcends space.

Before following this line of thought, we must recognize that from a purely experimental point of view it is difficult to justify the correlation of two electrons in the way required to resolve the EPR paradox. Does Mo’s wave function actually collapse when we observe Joe from a distance when they are not interacting? David Bohm, the initiator of deciphering the message of new physics, was thinking of a very practical way of correlating electrons, one that we can use to experimentally confirm the nonlocality of collapse.

An electron has a property called spin, which can have two discrete values. Think of the back as an arrow on an electron that points down or up. Bohm proposed that, under certain circumstances, we could cause two electrons to collide with each other in such a way that after the collision they would be correlated in the sense that their spin arrows would point in opposite directions. In this case, they say that both electrons are in a “singlet” state, or correlated in their polarization.

Proof of non-locality: Aspect’s experiment

Alain Aspect used the singlet type of correlation between two photons to prove the existence of a non-signal-mediated influence acting between two correlated quantum objects. He confirmed that the measurement of one photon affects another photon, polarization-correlated with it, without any exchange of local signals between them.

Imagine the following experimental setting: an atomic source emits pairs of photons, and two photons of each pair move in opposite directions. Each pair of photons is correlated in polarization—their polarization axes lie on the same line. Thus, if you see one photon through polarizing glasses with a vertical polarization axis (as they are usually worn), then your friend, located at a distance on the other side of the atomic source, will see a second correlated photon only if he also wears polarizing glasses with a vertical polarization axis. vertical axis. If he tilts his head so that the polarization axis of his glasses becomes horizontal, he will not be able to see his photon. If he tilts his head in such a way that it allows him to see his photon, then you will not be able to see the second photon of the correlated pair, since the polarization axis of your glasses does not match the polarization axis of your friend’s glasses.

Of course, the photon beams themselves are not polarized. They do not have a specific polarization unless you observe them with polarizing glasses; all directions of rays have the same probability of occurrence. Each photon is a coherent superposition of polarizations “along” and “across” each direction; it is our observation that collapses a photon with a certain polarization – longitudinal or transverse. In a long series of collapses there will be as many collapses with so-called longitudinal polarization as with transverse polarization.

Suppose that at first the polarization axes of both of your glasses are vertical, so that each of you can see one of the correlated photons (Fig. 30); but then you suddenly tilt your head so that the polarization axis of your glasses becomes horizontal instead of vertical. By your action (since you only see a photon if it is horizontally polarized) you have caused the photon you see to become horizontally polarized. However, oddly enough, your friend no longer sees the second photon of the pair unless he simultaneously turns his glasses around, since this correlated photon has also become horizontally polarized as a result of your action. This is a non-local collapse, isn’t it?

Img. 30.
Observations of polarization-correlated photons

If you really believe in material realism, you see something strange in this quantum theoretical construction of events, since what you do to one photon simultaneously affects its distant partner. No matter which direction you turn your polarizing glasses to see a photon, that photon’s correlated partner always takes on the direction of polarization along the same axis, no matter where or how far away from you it is. How does a photon know which way to turn unless it, in some sense, knows it from its partner? How can it recognize instantly, ignoring the speed limit of any signals to the speed of light?

Erwin Schrödinger wrote in 1935: “It is very inconvenient that [quantum] theory should allow the experimenter to introduce or direct a system into one state or another at his whim, despite the fact that he has no access to it.”

Material realists have been concerned for the last fifty years with the implications for their philosophy of such strong correlations between quantum objects. Until recently, they were still able to argue that the influence is mediated by an unknown local signal between photons and that it therefore strictly obeys the principle of realism. However, Alain Aspect and his collaborators, in their revolutionary experiment, proved that influence is transmitted instantly, and without any intermediate local signals.

As an example, suppose you take turns drawing cards from a deck. Your friend, who is sitting with his back to you, tells people what card you are drawing – and he is right every time. This correlation between you might be confusing to viewers at first. However, over time, people would figure out that you were somehow giving your friend a local signal. This is exactly how many so-called magic tricks work. Now suppose that due to circumstances there is simply no time for the exchange of a local signal between you and your friend. However, the magic of correlation continues to work – you draw a card and your friend names it correctly. This is the strange and extremely important result of Alain Aspect’s experiment.

Aspect used polarization-correlated photons emitted in opposite directions by calcium atoms. A detector was installed along the path of each photon beam. The crucial feature of the experiment – which made its conclusion irrefutable – was the use of a switch that changed the polarization setting of one of the detectors every one ten-billionth of a second (this time is shorter than it takes light or other local signal to travel the distance between the two detectors). But even so, changing the polarization setting of the detector with a switch changed the measurement result elsewhere—as it should, according to quantum mechanics.

How did information about changes in detector settings travel from one photon to its correlated partner? Certainly not using local signals. There wasn’t enough time for this.

How can this be explained? Let’s take Pagels’ comparison of reality to a deck of cards. The results of Aspect’s experiment are similar to the cards drawn in New York being identical to the cards drawn in Tokyo. The question remains: is the secret of nonlocality contained in the maps themselves, or is the consciousness of the observer also at play?

Material realists are reluctant to accept that quantum objects have nonlocal correlations and that if the collapse scenario is to be taken seriously, quantum collapse must be nonlocal. However, they refuse to see the significance of this and therefore miss the most important thing in the new physics.

One way to resolve the EPR paradox is to postulate that behind the scene of space-time there is an ether in which the transmission of signals faster than the speed of light is allowed. This solution would also mean a rejection of locality and materialism, and would therefore be unacceptable to most physicists. Additionally, FTL signals would make time travel into the past possible; This prospect worries people, and with good reason.

I prefer the obvious interpretation of Aspect’s experiment. According to the idealist interpretation, in this experiment it is your observation that collapses the wave function of one of the two correlated photons, causing it to assume a specific polarization. The wave function of its correlated partner also immediately collapses. A consciousness capable of instantly collapsing the wave function of a photon at a distance must itself be nonlocal, or transcendental. Thus, rather than regarding nonlocality as a property mediated by superluminal signals, the idealist argues that nonlocality is an integral aspect of the collapse of the wave function of a correlated system and is therefore an attribute of consciousness.

Thus, Einstein’s suspicion of the incompleteness of quantum mechanics, which was the working hypothesis of the EPR paradox, led to astonishing results. The intuition of a genius often turns out to be fruitful in unexpected ways, unrelated to the details of his theory.

This reminds me of a Sufi story. Mulla Nasrudin once encountered a gang of swindlers who wanted to take possession of his shoes. Trying to deceive the mullah, one of the scammers said, pointing to the tree: “Mullah, it’s impossible to climb this tree.”

“Of course available. “I’ll show you,” said the mullah, succumbing to the provocation. At first he was going to leave his shoes on the ground while he climbed the tree, but then he changed his mind, tied them and attached them to his belt. Then he began to rise.

The guys were discouraged. “Why are you taking your shoes with you?” – one of them exclaimed.

“Oh, I don’t know, maybe there’s a road up there and I might need them!” – responded the mullah.

The mullah’s intuition told him that scammers might try to steal his shoes. Einstein’s intuition told him that quantum theory must be incomplete because it could not explain correlated electrons. After all, what if the mullah discovered that there was a road at the top of the tree! Essentially, this is what Aspect’s experimental study of the EPR paradox showed.

Bell’s Theorem: The Death Knell for Material Realism

The paradox of the Aspect experiment is non-local collapse. Is it possible to avoid nonlocal collapse by assuming that pairs of photons in the experiment are emitted with a certain direction of their polarization axes? This is impossible in probabilistic quantum mechanics, but is it possible to correct the situation using hidden variables? If this eliminates nonlocality, then can invoking hidden variables save material realism? No, he can not. The proof of this is provided by Bell’s theorem (named after the physicist John Bell who discovered it), which shows that even hidden variables cannot save material realism.

Of course, the hidden variables that Einstein hoped would explain the EPR paradox and restore material realism were intended to be consistent with locality. They had to act on quantum objects in a local manner, as causal agents whose influence spreads through space-time with a finite speed and in a finite time. The locality of hidden variables is consistent with both the theory of relativity and the deterministic belief in local cause and effect, but is inconsistent with experimental evidence.

John Bell was the first to propose a set of mathematical relationships to test the locality of hidden variables; although these were not equations, they were no less rigorous. They described a type of relationship called inequalities. Aspect’s experiment, which proved that the connection between correlated photons is not mediated by any local signals, also showed that the inequalities formulated by Bell do not hold for real physical systems. Thus, Aspect’s experiment disproved the locality of hidden variables. It is no coincidence that quantum mechanics also predicts that Bell’s inequalities do not hold for quantum systems. Bell’s theorem states that in order to be compatible with quantum mechanics (and, as it turns out, with experimental data), hidden variables must be nonlocal.

The far-reaching consequences of the work of EPR and Bell are noteworthy. First, the study of the paradox pointed out by Einstein, Podolsky and Rosen revealed the nonlocality of quantum correlations and quantum collapse. Bell then showed that we cannot avoid nonlocality by invoking hidden variables, since they too exhibit nonlocality; therefore they cannot save material realism.

Consider physicist Nick Herbert’s simple, concise, and elegant treatment of Bell’s inequality.

Two beams of polarization-correlated photons move from the source in opposite directions. Let’s call the photons of the correlated pair Jo and Mo (J and M). Two experimenters observe the J-group and M-group photons using detectors made from calcite crystals, which serve as polarizing glasses. Let’s call these calcite crystals J-detector and M-detector (Fig. 31, a). As in the similar experiment shown in Figure 30, when the J-detector and M-detector are installed in parallel (that is, with parallel polarization axes) at any angle to the vertical, each observer sees one of the correlated photons. When the detectors are set at 90 degrees to each other, if one observer sees the photon, the other does not see its correlated partner. By definition, if an observer sees a photon, then the photon is polarized along the polarization axis of the calcite crystal of his detector (this polarization is indicated by the letter A), but if the observer does not see the photon, then the photon is considered to be polarized perpendicular to the polarization axis of his calcite crystal (this polarization is indicated by the letter R). Notice that now, thanks to hidden variables, we allow photons to have specific (correlated) polarization axes independent of our observations. This is the most important point – thanks to hidden variables, photons have predefined attributes.

So, a typical synchronized sequence of photon detection by two remote observers with parallel detector installations will show a picture of complete correspondence, for example:
Joe: APAAPPAPAPAAAPRRRR
Mo: ARAARRARARAAAARRRRR
And with perpendicular detector installations we will see a complete mismatch, for example:
Joe: RARAARARRAAAARRRRA
Mo: ARARRARAARRRRAAAAAR

None of these results are surprising anymore. Since the polarizations of the photons are predetermined, no collapse occurs (Note that the individual beams are not polarized, since in a long sequence each observer sees a mixture of 50-50 polarizations A and P).

We can quantify the polarization correlation, PC, which depends on the angle between the detectors. Obviously, if the detectors are located at exactly the same angle (PC = 1), we have complete correlation, and if they are perpendicular to each other (PC = 0), we have complete anti-correlation.

Here Bell asked – what is the value of PC for the intermediate angle? Obviously, it must be between zero and one. Suppose that for angle A the value of PC is 3/4. This means that with such installation of detectors (Fig. 31, b) for every four pairs of photons the number of matches (on average) is 3, and the number of mismatches is 1, as in the following sequence:
Joe: APRRRRARRRAAAAAA
Mo: APARRAARARPAPAPA

If we think of polarizations as messages in binary code, then the messages are no longer the same for both observers: Moe’s message (compared to Joe’s message) has one error for every four observations.

The case of the inequality relation described by Bell now becomes clear. Let’s start with a parallel arrangement of detectors; the observed sequences are now identical. Let us change the installation of the Mo detector to angle A (Fig. 31, b), and the sequences are no longer identical; they now contain errors—an average of one error for every four observations. In a similar way, let’s return to the parallel installation of detectors, and this time we will change the installation of Joe’s detector to the same angle A (Fig. 31, c); again there will be an average of one error for every four observations. This result does not depend on how far apart the detectors and their observers are. One could be in New York, another in Los Angeles, and the source of the photons is somewhere in between.

Img. 31.
How Bell’s inequality arises. If the latent variables were local, then the error rate (deviation from perfect correlation) in the experimental setting (d) would be equal, at most, to the sum of the error rates in the two settings (b) and (c).

If locality is true, if the postulated hidden variables that cause photons to take the particular directions of polarization required by the situation are local, then we can say with complete confidence: whatever you do with Joe’s detector, it cannot change Mo’s message – at least not not instantly. And vice versa. Thus, if, starting with parallel installations, observer Joe rotates the Joe detector by an angle A and if observer Mo simultaneously rotates the Mo detector by the same angle in the opposite direction (so that the detectors are now located at an angle of 2A to each other, Fig. 32, d), what should be the error rate? If the assumption of locality of hidden variables is true, then the action of each observer leads, on average, to one error per four observations, so that the total error rate will be 2 per four observations. However, it may happen that Joe’s mistake cancels out Mo’s mistake from time to time. Thus, the error rate will be less than or equal to 2/4 – this is Bell’s inequality. However, quantum mechanics predicts an error rate of 3/4. (Proving this is beyond the scope of this book.) So Bell’s theorem states that the theory of local hidden variables is incompatible with quantum mechanics.

Bell’s inequalities have been experimentally studied. In 1972, Berkeley physicists John Clauser and Stuart Friedman showed that Bell’s inequalities were indeed violated, and quantum mechanics was rehabilitated. Aspect then proved with his experiment that there can be no local signals at all between the two detectors.

Note that Bell’s work (and also Bohm’s work, since it led to the idea of ​​measuring polarization correlation) set the stage for Aspect’s experiment, which established nonlocality in quantum mechanics. Now you can appreciate why, at a physics conference in 1985, a group of physicists sang the following words to the tune of “Jingle Bells”:
Singlet Bohm, singlet Bell
Singlet all the way.
Oh, what fun it is to count
Correlations every day.
(Singlet Bom, singlet Bell, singlet all the way.
Oh, what fun it is to count correlations every day.)

According to Bell’s theorem and the Aspect experiment, if hidden variables exist, they should be able to instantly affect correlated quantum objects, even if they are located on different ends of the galaxy. In an Aspect experiment, when one experimenter changes the setting of his detector, hidden variables control not only the photon reaching that detector, but also its distant partner. Hidden variables can act non-locally. Bell’s theorem destroys the dogma of local cause and effect accepted in classical physics. Even if you introduce hidden variables to find a causal interpretation of quantum mechanics, as David Bohm does, these hidden variables must be nonlocal.

David Bohm compares Aspect’s experiment to seeing a fish on two different screens on two televisions. Whatever one fish does, the other does too. If we consider the images of a fish to be the primary reality, this seems strange, but from the point of view of a “real” fish, everything is very simple.

Bohm’s analogy is reminiscent of Plato’s allegory of shadows in a cave, but there is one difference. In Bohm’s theory, the light that the image of a real fish projects is not the light of creative consciousness, but the light of cold, causal hidden variables. According to Bohm, everything that happens in spacetime is determined by what happens in the nonlocal reality outside spacetime. If this were true, then our free will and creativity would ultimately be illusions, and human drama would have no meaning. The idealistic interpretation promises just the opposite: life is filled with meaning.

It’s a bit like the difference between film and stage improv. In film, action and dialogue are defined and fixed, but in live improvisation, variations are possible.

According to the idealistic interpretation, violation of the inequalities described by Bell means non-local correlation between photons. No hidden variables are needed for explanation. Of course, to collapse the wave function of nonlocally correlated photons, consciousness must act nonlocally.

If we return to Bohm’s analogy with the fish and its images on two televisions, then the idealist interpretation agrees with Bohm that the fish exists in a different order of reality; however, this order is the transcendental order in consciousness. The “real” fish is a form of possibility already present in consciousness. In the act of observation, the images of the fish simultaneously appear in the world of manifestation as the subjective experience of observation.

Let’s take another facet of Aspect’s experiment. This experiment and the concept of quantum nonlocality have led some people to hope that it is somehow related to a violation of causality—the idea that cause always precedes effect. Not necessary. Since each observer in the Aspect experiment always sees a disordered mixture of 50-50 polarizations A and P, it is impossible to send a message with their help. The correlation we see between both observers comes after we compare the two data sets. Only then does its meaning arise in our minds. Therefore, Bell’s theorem and Aspect’s experiment do not imply a violation of causality, but that simultaneously occurring events in our space-time can be meaningfully attributed to a common cause located in a non-local sphere outside of space and time. This common cause is the act of non-local collapse of the wave function by consciousness. (The fact that meaning is discovered after the fact is extremely important and will come up again in this book.)

The Aspect experiment shows not the transmission of a message, but communication in consciousness, a community inspired by a common cause. Psychologist Carl Jung coined the term synchronicity to describe meaningful coincidences that people sometimes experience—coincidences that happen without cause, except perhaps for a common cause in the transcendental realm. The nonlocality of Aspect’s experiment exactly corresponds to Jung’s description of synchronicity: “Synchronic phenomena prove the simultaneous occurrence of significant equivalences in heterogeneous, causally unrelated processes; in other words, they prove that the content perceived by the observer can at the same time be represented by an external event, without any causal connection between them. It follows from this that either the psyche cannot be localized in time, or that space is secondary in relation to the psyche.” Further, Jung expresses, in our opinion, a striking conjecture: “Since psyche and matter are contained in the same world and, moreover, are constantly in contact with each other, and are ultimately based on inconceivable transcendental factors, it is not only possible but it is even quite probable that mind and matter are two different aspects of the same thing.” This characterization will be useful in our consideration of the brain-mind problem.

If synchronicity still seems like a confusing concept to you, perhaps the following story will help. The rabbi was walking through the city square when a man suddenly fell on him from a balcony. Since the man’s fall was broken by the rabbi, nothing happened to him, but the poor rabbi’s neck was broken. Since the Rabbi was a respected wise man who always learned himself and taught others through his own life experiences, his followers asked, “Rabbi, what is the lesson in having your neck broken?” The Rabbi replied: “Well, they usually say, what goes around comes around. Look what happened to me. A man falls from a balcony and I break my neck. Some sow and some reap.” This is synchronicity.

The same is the case with two correlated photons or electrons, or with any other quantum system. You observe one of them and it instantly affects the other as non-local consciousness synchronously collapses their wave functions.

Jung had a term for the transcendental realm of consciousness, where the common cause of synchronous events is located – the collective unconscious. It is called “unconscious” because we are normally not aware of the non-local nature of these events. Jung empirically discovered that, in addition to Freud’s discovery of the personal unconscious, there is a transpersonal collective aspect of our unconscious that must operate outside of space-time, that is, be non-local, since it appears to be independent of geographical origin, culture or time.

The nonlocal correlations of Bell’s theorem and the Aspect experiment are acausal coincidences, and their meaning – as in the case of synchronous events – always arises after the fact when observers compare their data. If these correlations are examples of the synchronicity described by Jung, then the associated aspect of nonlocal consciousness must be akin to Jung’s collective unconscious. When we observe a quantum object, our nonlocal consciousness collapses its wave function and chooses the outcome of the collapse, but we are usually not aware of the nonlocality of the collapse and choice. We will discuss this issue further in Chapter 14.

Physics becomes a link with psychology

My interpretation of quantum mechanics opens the way for the application of physics to psychology. However, further discussion of this interpretation may be useful, as understanding emerges in the heat of debate.

If we are not aware of the actions of nonlocal consciousness, then isn’t nonlocal consciousness another unnecessary assumption, like the assumption of hidden variables? While it is certainly possible to think of nonlocal consciousness as analogous to hidden variables, one might just as well assume that the idealist interpretation offers a new way of understanding hidden variables. Nonlocal consciousness does not constitute causal parameters, as Bohm imagined them, but acts through us; or, more correctly, it is us – only in a thinly disguised form (and, as mystics of all times testify, man is able to penetrate this disguise to a greater or lesser extent). Moreover, non-local consciousness does not operate with causal continuity, but with creative discreteness – from moment to moment, from event to event, as when the quantum wave function of the mind-brain collapses. Discreteness, a quantum leap, is an integral part of creativity; it is precisely the abrupt exit from the system that consciousness needs in order to see itself, as in self-reference.

At one time, probabilistic quantum mechanics encouraged philosophers to take a fresh look at the problem of free will. However, if you still believe in materialism, probability provides only a pale semblance of free will. When you are at a T-intersection, where should you go? Are your free choices determined by quantum mechanical probabilities or are they the result of some classical determinism operating in your mind? The difference is not that important. There are other situations where true freedom of choice comes into play.

Let’s take creative work. In creativity, we constantly make leaps that take us out of the context of our past experiences. In these cases, we must use the freedom to be open to new contexts.

Or take a case where you have to make a moral decision. Religious creeds may suggest that moral values ​​must be dictated by an authoritative source, but a closer look at the process by which human beings make moral decisions reveals that a truly moral decision based on faith and values ​​requires real freedom of choice—the freedom to change the context of a situation.

As an example, consider the struggle for independence from so-called benevolent imperial rule. Ordinary violent uprisings quickly become unethical – don’t they? Yet Gandhi succeeded in overthrowing the rule of the British Empire because he was able to change the context of India’s struggle for independence, time after time using his only weapon: creative choice. His methods were non-violent protest against the imperialists and non-cooperation with the government – these methods were effective and, at the same time, ethical.

Most importantly, let us take the perception of meaning, which is a common feature of many interesting phenomena in the subjective sphere. There is a book on the table in front of you. The person takes it and makes a meaningless sound, purposefully drawing your attention to it. Suddenly you understand the meaning of his behavior. He pronounces the word “book” in his language. How did the meaning of his action emerge in your mind? This is due to non-locality – a jump from your local space-time system.

The surprising nature of this communication may not be obvious to you because it is so familiar. However, imagine that you are young Helen Keller, deaf-blind from birth. When Annie Sullivan alternately dipped Helen’s hand into the water and wrote the word “water” with her finger on her palm, she was using the same communication context as in the example above with the word “book.” Helen must have thought her teacher was crazy until the meaning of Annie’s actions dawned on her—until she made the leap from her existing context to a new context.

“The more the universe seems intelligible, the more it seems devoid of meaning,” writes Nobel Prize winner Steven Weinberg at the conclusion of his popular book on cosmology. We agree with this. Concepts such as non-local and unifying consciousness and the idea of ​​non-local collapse make the universe less understandable to the materialist scientist. These concepts also make the universe much more meaningful to everyone else.

Farsighting as a nonlocal quantum effect

According to the idealistic interpretation, the observation of quantum nonlocal correlations is also an obvious manifestation of the nonlocality of consciousness. Can we therefore find confirmation of quantum nonlocality in subjective experience? Does such evidence exist? Yes. Such evidence is controversial, but interesting.

Imagine that an image of a statue that you have never seen appears before your mind’s eye, so clearly that you can draw it. Next, imagine that your friend is actually looking at the statue at the very moment when its image appears in your head. This would be telepathy, or far-sighting, and could well be an example of communication through non-local consciousness.

A skeptical scientist might suspect that you already know what your friend will be looking at. So suppose two researchers, using a computer, designed an experiment so that neither you nor your friend (or, for that matter, the researchers themselves) knew in advance which object would be viewed, but only the time at which telepathic transmission occurs.

A skeptic might still argue that the drawing is open to different interpretations. Can you objectively decide whether your drawing actually matches what your friend saw? So researchers use impartial judges—or, even better, a computer—to compare dozens of your drawings with dozens of places your friend sees. Would you hope that a skeptical scientist would change his mind about telepathy?

Such experiments were carried out in many different laboratories, and positive results were obtained with subjects both with and without psychic abilities. The correlations still persisted. Then why is telepathy still not recognized as a scientifically proven discovery? One reason, from a scientific perspective, is that extrasensory perception (ESP) data are not strictly reproducible—they are only statistically reproducible. In this regard, it is believed that if ESP were possible, we could somehow convey meaningful messages through it, which would create chaos in the orderly world of causality. However, the most important reason for skepticism about ESP may be that it does not seem to be related to any local signals perceived by our senses, and is therefore prohibited by material realism.

We can try to explain far-sighted data as experiences of non-local correlation that arise in our experience because our mind is of a quantum nature. (Suspend your disbelief for a moment if necessary.) From the perspective of quantum nonlocality, as demonstrated by Aspect’s experiment, the ESP problem appears to be one of choice. Only two correlated telepaths, like the two photons in the Aspect experiment, non-locally share information. In that experiment, that the photons are correlated is indicated by the choice of experimental setting, the source of the photons, and the meaning assigned to the data. Likewise, the correlation of psychics in a far-sighting experiment must be related to the preparation of the experiment, the setting, and the meaning assigned to the data.

Both the uncausality and significance in visioning (and perhaps ESP in general) argue strongly for understanding these phenomena as synchronicity events caused by nonlocal quantum collapse. Recall that the reason nonlocal quantum collapse does not contradict the principle of causality is that it does not allow message passing.

It could be the same with far-sightedness. Perhaps non-local communication between psychics is not associated with the transfer of useful information. The correlation between the distant vision of one psychic and the drawing of another psychic correlated with it is statistical in nature, and the significance of the communication becomes apparent only after comparing the drawing with the location in question. Similarly, in Aspect’s experiment, the significance of communication between correlated photons becomes apparent only after comparing two sets of distant observations.

A recent experiment by Mexican neuroscientist Jacobo Greenberg-Silberbaum and colleagues directly supports the idea of ​​nonlocality in the human mind-brain—the experiment is the brain equivalent of Aspect’s experiment with photons. Two subjects were asked to communicate for thirty or forty minutes until they began to feel “direct communication.” They then entered individual Faraday cages (boxes made of metal mesh that block all electromagnetic signals). Now, one subject, unbeknownst to his partner, was presented with a flashing light signal that caused an “evoked potential” (an electrophysiological response to a sensory stimulus, recorded using an EEG) to appear in his brain. But, surprisingly, while the experimental partners maintained their “direct communication,” the second subject’s brain also exhibited electrophysiological activity, called “carryover potential,” very similar in shape and strength to the evoked potential in the first subject’s stimulated brain. (In contrast, control subjects had no transfer potential.) A simple explanation for these results is quantum nonlocality: by virtue of their quantum nature, the two mind-brains act as a nonlocally correlated system in which the correlation is established and maintained through nonlocal consciousness .

It is important to note that none of the subjects in the experiment had any conscious experience of the transference potential. Thus, there was no transfer of information at the subjective level, and the principle of causality was not violated in any way. The nonlocal collapse and subsequent similarity of evoked and transfer potentials should be considered a synchronicity event; the significance of the correlation becomes clear only after comparing the potentials. This is similar to the situation in the Aspect experiment.

Can we also find evidence of nonlocality in time? Is there any truth to the so-called cases of foresight that sometimes become known to the public? For example, they claim that someone foresaw the assassination of Robert Kennedy. An experiment with foresight is difficult to plan in advance. So I don’t see much point in arguing about whether a certain psychic actually had genuine precognition or not. However, there is a clever analysis of Schrödinger’s cat paradox, which, at least from a naive point of view, entails the idea of ​​non-locality in time. According to what we said earlier about the necessity of consciousness for the collapse of the living/dead cat dichotomy, until we observe the cat, it is in an indeterminate intermediate state. Suppose we sprinkle soot on the floor around the cage and arrange for an automatic device to open the box after an hour. Suppose we come back another hour later and find that the cat is alive. Question: Will cat tracks be visible on the soot? If so, how did the cat leave these marks? After all, an hour ago the cat was still in an uncertain state. The idea of ​​nonlocality in time provides the easiest way to explain the paradox – as suggested by the delayed choice experiment.

Out-of-body experiences (out-of-body experiences)

Are there other parapsychological phenomena other than far-sighting that can be explained by the quantum/idealistic model of consciousness? While it’s too early to definitively say that this is the case, there are indications that suggest we’d better keep an open mind on the matter.

Many people claim that they have actually experienced leaving their body. During these outings, they may visit friends, observe surgery performed on their own body, or even travel to distant places. This phenomenon is called “out-of-body experience” (OBE). The similarity of the OBE with the translocation of the “I” of the mind outside the body is undeniable, but how can this be? This is very similar to mind-body dualism.

The reality of out-of-body experience as a genuine phenomenon of consciousness is increasingly being questioned. Consider, for example, Michael Sabom’s book Memories of Death, which reports significant and systematic research into OBEs in relation to near-death experiences. As a cardiologist with access to medical records, Sabom was in a unique position to verify many of the technical details in patients’ reports of resuscitation efforts performed on their nearly dead bodies. His patients described very accurately procedures that were clearly beyond the sight of their physical bodies.

Since these patients had a long history of repeated hospitalizations and were very familiar with medical procedures, it would not be too surprising if they made successful guesses based on this knowledge. To rule out this possibility, Sabom used a control group of patients with the same medical histories, including near-death experiences, who had not experienced an OBE. When these patients were asked what they thought happened in the intensive care unit while they were on the verge of death, they gave very imprecise answers that, even in general, corresponded very little to the facts. Initially a skeptic, Sabom conducted his research with extreme care and evaluated its results according to the strict standards of the methodology of modern experimental psychology.

Can the mind really leave the body? In such parapsychological phenomena as OBEs, this is certainly the case. This legitimate question cannot be dismissed unceremoniously by citing hallucinations, as local materialist scientists sometimes try to do. Sabom, who has explored the question of whether OBEs are hallucinatory in nature very carefully, stated the following: “Unlike NDEs [near-death experiences], autoscopic hallucinations [seeing oneself] involve: 1) perception by the physical body (“original”) your projected image (“double”); 2) direct interaction between the “original” and the “double”; 3) are perceived as unreal; and 4) tend to evoke negative emotions. For these reasons, autoscopic hallucinations do not appear to be a plausible explanation for NDE.”

To be honest, when I first learned about OBEs in the early 1980s, I was quite impressed by this and other studies and began to look for some alternative way of understanding this phenomenon that would allow me to explain it from a scientific point of view – without citing neither hallucinations nor transmigration of the mind. Anyway, the talk of disembodied minds, or astral bodies as they are called in certain circles, watching their physical bodies undergo surgical operations seemed to me an unconvincing and simplistic explanation of what I could only accept as a subjective perception of the optical illusions.

To clarify this difference, let’s take the example of a well-known optical illusion. I have always been fascinated by the moon illusion: the fact that the moon on the horizon looks much larger in nature than in a photograph. Detailed experiments conducted by scientists, as well as my own loose tinkering with this phenomenon, have convinced me that it is associated with the illusion of size. When the moon is above the horizon, the brain mistakenly perceives it as being further away than when it is at its zenith and makes adjustments to make the image appear larger.

I continued to be haunted by the idea that the OBE must be some kind of illusion, but of what? In the meantime, I also studied the literature on visioning. It suddenly occurred to me that an OBE must be an illusory construct of far-sightedness, which is a non-local vision outside a person’s physical field of vision. From an objective point of view, Saibom’s patients, who were on the verge of death, did just that. But why the illusion of being outside the body?

When very young children see or hear something outside their field of sense perception, they experience the opposite difficulty to that experienced by the adult visionary. This childhood difficulty – the difficulty in exteriorizing the universe – arises from the fact that all our awareness of the external world actually occurs in our head, since visual and auditory images are formed in our brain. Gradually, using primarily their senses of touch and taste, children learn to exteriorize the world. They develop selective perception, allowing them to recognize distant visible or audible objects.

In an adult, the unfamiliar experience of far-sighting an object outside the visual field should cause significantly more cognitive chaos than a child’s experience. The adult’s conditioned and deeply ingrained perceptual system tells us that the object is somewhere else; therefore, in order to “see” it, you need to be “there”. As with the moon illusion, the brain mistakenly interprets non-local far-sightedness as an out-of-body experience. So if a person watches himself being operated on under general anesthesia, which is normally impossible, his soul, or astral body, should be near the ceiling or at the other end of the room – since that is where he seems to perceive what is happening.

Once I realized that an OBE could be a visionary phenomenon, the veil lifted. Finally I had an explanation for the OBE that could satisfy the scientist’s skepticism. The key to resolving the paradox is the nonlocality of our consciousness.

By the way, if you are skeptical about the nonlocality of far vision and believe that it may be mediated by some as yet undetected local signals, then you should know that researchers, especially in Russia, have been searching for such signals for many years and have found nothing. In some of their experiments, psychics had to demonstrate their ESP abilities while sitting in a Faraday cage, but these shielding cages do not appear to have any noticeable effect on ESP.

In addition, local signals propagate from their source into the surrounding space, so their intensity should decrease with distance from the source. In contrast, with non-local communication no such attenuation is observed. Since the available evidence suggests that there is no decline in distance vision, distance vision must be nonlocal. It is therefore logical to conclude that psychic phenomena, such as far vision and out-of-body experiences, are examples of non-local action of consciousness.

Any attempt to dismiss a misunderstood phenomenon by simply explaining it as a hallucination becomes irrelevant if a consistent scientific theory can be applied. Quantum mechanics supports such a theory, providing decisive evidence for the nonlocality of consciousness; it poses an empirical challenge to the dogma of locality as a universal limiting principle.

Perhaps even more surprising is that the idea of ​​nonlocality of consciousness resolves not only the paradoxes of extrasensory perception, but, as we will see in the next chapter, also the paradoxes of ordinary perception.

In all likelihood, as it becomes clear that Bell’s theorem and the Aspect experiment have indeed heralded the demise of material realism, scientific resistance to accepting the validity of far-sighting experiments and other psychic phenomena will begin to wane. At a recent Physical Society conference, someone overheard one physicist say to another, “Only someone with a brain of stone would care about Bell’s theorem.” Even more encouragingly, a survey of physicists attending the conference found that Bell’s theorem concerned 39% of them. Given such a high percentage, it is quite possible to expect that the idealistic paradigm of physics will receive an impartial assessment.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM

Material realism cannot be saved. Then there are two important questions to answer: first, why does the macroscopic universe look so real? Second, how can we do science without some brand of realism? The solution is to incorporate material realism into monistic idealism. Before we talk about how this can be done, let’s think about why an interpretation of quantum mechanics is required in the first place. Why is philosophy needed to understand it? Why can’t she speak for herself? Here is a short list of reasons:

1. The state of a quantum system is determined by the Schrödinger equation, but the solution to this equation—the wave function—is not directly related to anything we observe. So the first question of interpretation concerns what the wave function represents: a single object? a group of similar events? ensemble of objects? The square of the wavefunction determines the probabilities, but how should we understand the probabilities? This requires interpretation. We prefer the interpretation in terms of a single object, but this is still a matter of philosophy.

2. Quantum objects are subject to the Heisenberg uncertainty principle: it is impossible to simultaneously and accurately measure pairs of conjugate variables, such as position and momentum. Is it just a matter of measurement (that quantum probes transfer an uncontrollable amount of energy to the object they measure), or does the uncertainty principle come from the nature of things? The uncertainty principle arises from the nature of the wave packets that we have to construct in order to extract localized particles from the waves. Again, this question is a matter of interpretation and philosophy.

3. The paradox of wave-particle duality—that quantum objects have both wave and particle aspects—needs resolution, which means interpretation and philosophy.

4. What physical reality, if any, could a coherent superposition have? Can Schrödinger’s cat paradox really be resolved without seriously considering such a question? And its consideration is inevitably connected with interpretation and metaphysics.

5. Are discreteness and quantum jumps truly fundamental aspects of the behavior of quantum systems? In particular, we depicted the collapse of the wave function or coherent superposition in the measurement situation as a discrete event. But is collapse necessary? Is it possible to find an interpretation that avoids collapse and therefore avoids discreteness? Note that the motivation for seeking such an interpretation is a desire to strengthen a philosophical position—that of realism.

6. Bohr’s correspondence principle states that under certain conditions (for example, for very close energy levels in atoms) quantum mechanical predictions reduce to those of classical mechanics. This guarantees that classical mechanics can be used to make predictions in most situations, but does it guarantee that measuring instruments behave classically when needed? Some physicists (all of them realists) believe that this is a matter of philosophy.

7. Bell’s theorem and the Aspect experiment force us to ask: how should we interpret the meaning of quantum nonlocality? This has extremely serious consequences for our philosophy.

Material realism, stymied by quantum mechanics, is helpless whenever the nature of quantum reality is questioned, whether in connection with the uncertainty principle, wave-particle duality, or coherent superpositions. Whenever we ask whether there is some other kind of reality beyond material reality, we put material realism in an uncomfortable position. Likewise, true discreteness points to a transcendental order of reality and thus the collapse of material realism.

The paradoxes of quantum measurement (for example, the paradox of Schrödinger’s cat) create insurmountable difficulties for material realism. A materially real cat, having no other order of reality for its existence, must face the problem of coherent superposition. Can a cat really be alive and dead at the same time?

Finally, the decisive challenge to material realism comes from Bell-Aspect nonlocality. There are only two alternatives, and neither of them is compatible with a strictly materialistic philosophy. Obviously, the abandonment of locality in favor of superluminal signals in a sphere beyond space-time, as well as the assumption of non-local hidden variables, is a leap beyond the limits of the material order. The rejection of strict objectivity, or the recognition of any kind of role of conscious observation, relegates material realism to the category of outdated theories, which include the flat earth, the ether and phlogiston (a never-discovered substance assumed to be the active source of heat and light in combustion).

Is it possible to reconcile the theory of many worlds with idealism?

All the various models that have been proposed to resolve Schrödinger’s cat paradox are untenable, with the exception of three – the theory of many worlds, the theory of nonlocal hidden variables, and the theory proposed here, based on monistic idealism. From the discussions in the previous chapter, you can see good reason to question latent variable interpretation. Here idealism has a clear advantage. Can idealism also claim an advantage over the theory of multiple worlds?

Many worlds theory attempts to resolve the difficulty posed by Schrödinger’s cat paradox by postulating that the universe splits into two branches: one with a dead cat and a sad observer, and another where the cat is alive and the observer is happy. However, try using this theory to resolve the paradox of quantum nonlocality. Measuring one electron still splits the world of the second electron correlated with it, and just as instantly. Thus, this interpretation appears to compromise locality and thus ultimately does not support material realism.

Even though the many-worlds theory does not help support material realism, it should certainly be considered a viable alternative to the idealist interpretation. But this theory (like the theory of nonlocal hidden variables) abandons many of the revolutionary aspects of the Copenhagen interpretation. On the contrary, monistic idealism begins where the Copenhagen interpretation becomes vague; it explicitly declares that quantum waves or coherent superpositions are real, but exist in a transcendental realm beyond material reality.

In fact, the idea of ​​multiple worlds is easy to incorporate into an idealist interpretation. If we look closely at the many worlds theory, we find that it uses conscious observation. For example, how is it determined when the universe branches? If this happens during measurement, then the theory, according to the definition of measurement, includes the role of the observer.

According to the idealist interpretation, coherent superpositions exist in the transcendental realm as formless archetypes of matter. Let us assume that the parallel universes of the multiple worlds theory have archetypal rather than material content. Let us assume that they represent the universes of the mind. Then, instead of saying that each observation separates a branch of the material universe, we can say that each observation creates a causal path in the fabric of possibilities in the transcendental realm of reality. As soon as the choice is made, all paths except one are excluded from the world of manifestation.

See how this way of reinterpreting the many-worlds formalism gets rid of the costly proliferation of material universes.

An attractive feature of many worlds theory is that the existence of many worlds makes it a little more pleasant to apply quantum mechanics to the entire cosmos. Because quantum mechanics is a probabilistic theory, physicists are uncomfortable thinking about the wave function of the entire universe, such as Hawking talks about. They doubt whether it is possible to attribute meaning to such a wave function if nothing else exists besides it. Many worlds theory—even in the transcendental realm—helps deal with this problem.

The truly cosmological question can now be answered: How did the cosmos exist for the last fifteen billion years if for most of that time there were no conscious observers to collapse the wave functions? Very simple. The cosmos never came into being in a concrete form and never remains the same. Past universes, one after another, cannot be seen as paintings on canvases from which the present universe unfolds in time, although, if you think about it, this unfolding universe is exactly how material realism depicts the situation.

I propose that the universe exists as formless potency in many possible branches in the transcendental realm and only becomes manifest when observed by conscious beings. Of course, there is the same circular character here that gives rise to the self-referentiality discussed in Chapter 6. It is these self-referential observations that chart the causal history of the universe, rejecting countless parallel alternatives that never reach material reality.

This interpretation of our cosmological history may help explain a puzzling aspect of the evolution of life and intelligence, namely that there is only a very small probability of life evolving from prebiological matter through favorable mutations leading to the emergence of humans. Once we accept that biological mutation (including the mutation of prebiological molecules) is a quantum event, we understand that the universe splits at each such event in the transcendental realm, becoming many branches until a conscious being appears in one of the branches, capable of looking with awareness and make a quantum measurement. At this point, the causal path leading to this conscious being collapses into space-time reality. John Wheeler calls this scenario of closing the circuit of meaning “observer complicity.” Meaning arises in the universe as it is observed by conscious beings, choosing causal paths from an infinite variety of transcendental possibilities.

If this sounds like we are restoring an anthropocentric view of the universe, so be it. The time has come and the context has emerged for a strong formulation of the anthropic principle—the idea that “observers are needed to create the universe.” It is time to recognize the archetypal nature of the creation myths (found in the Book of Genesis of the Judeo-Christian tradition, in the Hindu Vedas, and in the sacred texts of many other religious traditions). Space was created for our sake. Such myths are compatible with quantum physics and do not contradict it.

Much of the misunderstanding arises because we tend to forget what Einstein said to Heisenberg: “What we see depends on the theories we use to interpret our observations.” (Of course, Immanuel Kant and William Blake had already talked about this, but they were ahead of their time.) How we reconstruct the past always depends on the theories we use. For example, think about how people viewed sunrise and sunset before and after the Copernican Revolution. Copernicus’s heliocentric model shifted the focus—we were no longer the center of the universe. But now the tide is returning. Of course, we are not the geographical center of the universe, but that’s not what we’re talking about. We are the center of the universe because we represent its meaning. Idealist interpretation fully recognizes this dynamic aspect of the past – the fact that the interpretation of what we see, like myth, changes as our concepts change. And we should not be chauvinists: it can just as easily be assumed that in a universe that has collapsed into a space-time reality, there is the possibility of the evolution of enormous numbers of intelligent, self-aware beings on billions and billions of planets in all corners of the expanding universe.

How can an idealistic cosmos create the appearance of realism?

If reality consists of ideas ultimately manifested by consciousness, then how can we explain such great unanimity? If idealism wins in a philosophical debate and if the philosophy of realism is untenable, then how can one practice science? David Bohm said that science is impossible without realism.

There is a certain truth in Bohm’s statement. But I will present convincing logical evidence that the essence of scientific realism fits well under the broader umbrella of idealism.

To fully consider this issue, let us turn to the origin of the dichotomy of realism and idealism in the paradox of perception. The artist Rene Magritte depicted a smoking pipe in the painting, but the caption under the painting read: “This is not a pipe.” Then what is it? Suppose you say, “This is a picture of a pipe.” This is a good answer, but if you are really good at solving riddles, you will say: “I see an image evoked in my head (brain) by the sensory impressions of the image of a pipe.” Exactly. Nobody ever sees a painting in an art gallery. You always see a picture in your head.

Of course, a painting is not the object depicted in it. The map is not the territory. Is there a picture at all? We can only say with certainty that there is some kind of image in our head – a purely theoretical image. In any perceptual event we actually see this theoretical, very personal image. We assume that the objects we see around us are empirical objects of a common reality: completely objective and public, fully accessible to empirical study. However, in fact, our knowledge of them is always acquired by subjective and personal means.

Thus arises the old philosophical puzzle as to what is real: a theoretical image that we actually see – but only in a private way, or an empirical object that we do not directly see, but about which we reach a common opinion?

The internal and private character of the theoretical image would not be a problem, and no noticeable dichotomy would arise, if there were always a one-to-one correspondence between this image and the empirical object, which could be directly confirmed by other people. But that’s not true; There are optical illusions. There are creative and mystical experiences of subjective images that do not necessarily correspond to anything in the immediate reality of consensus. Therefore, the authenticity of theoretical images is doubtful, and this, in turn, calls into question the authenticity of empirical objects, since we never experience them without the mediation of a theoretical image. This is the paradox of perception: apparently, we cannot trust either the authenticity of our theoretical image or the authenticity of the public, empirical object in the reality of consensus. It is from such paradoxes that philosophical “isms” are born,

Throughout history, two schools of philosophy have constantly argued about what is truly real. The idealistic school believes that the theoretical image is more real and that the so-called empirical reality is only ideas of consciousness. In contrast, realists argue that there must be real objects independent of us about which we form common opinions.

Each of these points of view has its practical applications. Without some form of realism—the assumption that there are empirical objects independent of the observer—natural science is impossible. Agree. However, science is equally impossible without the conceptualization and testing of theoretical ideas.

Therefore, we need to overcome the paradox. This was done by the philosopher Gottfried Leibniz, and then by another philosopher, Bertrand Russell, with a seemingly absurd idea: both views could be true if we had two heads, so that the empirical object was inside one of them, but outside is different. The empirical object would be outside what might be called our little head, and this would confirm realism; at the same time the object would be inside our big Head and thus would be a theoretical idea in this big Head, which would satisfy the idealists. By a clever philosophical maneuver, the object became both an empirical object outside the empirical heads and a theoretical image within the all-encompassing theoretical Head.

You may ask, is this theoretical big Head just theoretical or does it have some kind of empirical reality? The matter becomes more complicated when we realize that this big Head contains all the empirical small heads and can thus itself be an object of empirical study. Suppose we take the idea of ​​this big Head seriously.

On closer examination we begin to think that the big Head does not have to be separate, but can be installed in all empirical heads (that is, there is no reason to postulate more than one such Head, since it contains all empirical reality; we can all have one common Head). Suppose that the head, the brain, forms part of consciousness, which has two aspects, two different ways of organizing reality: a local aspect, completely limited to the empirical brain, and a global consciousness, which contains the experience of all empirical objects, including empirical brains.

It is easy to recognize the non-locality in the last statement. The idea of ​​nonlocality lends respectability to the seemingly absurd speculations of Leibniz and Russell. If, in addition to local ways of collecting data, there is a nonlocal organizing principle associated with the brain-mind—nonlocal consciousness—what then? This is equivalent to saying that we have two heads, and the paradox of perception is resolved.

How close our ideas about reality now seem to be to what the compilers of the Upanishads guessed thousands of years ago:
It is inside all of this,
It is outside of all of this.

Moreover, both idealism and realism can now be justified. They’re both right. For if the brain-mind itself is an object of non-local consciousness that contains all reality, then what we call objective empirical reality resides in that consciousness. It represents the theoretical idea of ​​this consciousness – and therefore idealism is justified. However, when this consciousness becomes immanent as a subjective experience in a part of its creation (in the mind-brain located in our head) and looks through sensory perceptions at other locally separate parts as objects, then the doctrine of realism is useful for studying the patterns of behavior of these parts .

Now let’s ask an important question: why is there such great consensus about reality? The phenomenal world appears undeniably objective for two reasons. First of all, classical bodies have enormous masses, which means that their quantum waves expand very slowly. The small expansion makes the trajectories of the centers of mass of macroscopic objects very predictable (whenever we look, we always find the moon where we expect to see it), which gives the appearance of continuity. Additional continuity is brought by the perceptual apparatus of our mind-brain.

Secondly, and more importantly, the complexity of macroscopic bodies translates into a very long time for complete renewal. This allows them to create a memory or record, however temporary it may ultimately be. Because of these records, we tend to view the world from a causal point of view, using the concept of unidirectional time, which is independent of consciousness.

Conglomerates of quantum objects, which we call classical, are needed as measuring instruments to the extent that we can determine their approximate trajectories and talk about their memory. Without these classical objects, measuring quantum events in spacetime would be impossible.

In nonlocal consciousness, all phenomena, even so-called empirical, classical objects, are objects of consciousness. It is in this sense that idealists say that the world consists of consciousness. It is clear that the idealistic and quantum views converge if we accept a nonlocal solution to the paradox of perception.

I trust my intuition, which tells me that the idealistic interpretation of quantum mechanics is correct. Of all the interpretations, only this one promises to take physics into a new arena – the arena of the brain-mind-consciousness. If history is to be believed, all new breakthroughs in physics expand its field of application. Could quantum mechanics and idealistic philosophy together form the basis of an idealistic science that could resolve the tangled paradoxes that have perplexed us for millennia? Yes, I suppose they can. In the next part of this book I try to lay the foundation for this solution.

Abraham Maslow wrote: “If there is any cardinal rule of science, it is, in my opinion, the acceptance of the obligation to recognize and describe all reality, all that exists, all that happens… At its best it [science] is completely open and doesn’t exclude anything. She doesn’t have any ‘entrance exams’.”

With idealistic science we arrive at a science that requires no entrance exams, that excludes neither the subjective nor the objective, neither spirit nor matter, and is therefore capable of unifying the deep dichotomies of our thinking.

The book “The Self-Aware Universe. How consciousness creates the material world.” Amit Goswami

Contents

PREFACE
PART I. The Union of Science and Spirituality
CHAPTER 1. THE CHAPTER AND THE BRIDGE
CHAPTER 2. OLD PHYSICS AND ITS PHILOSOPHICAL HERITAGE
CHAPTER 3. QUANTUM PHYSICS AND THE DEATH OF MATERIAL REALISM
CHAPTER 4. THE PHILOSOPHY OF MONISTIC IDEALISM
PART II. IDEALISM AND THE RESOLUTION OF QUANTUM PARADOXES
CHAPTER 5. OBJECTS IN TWO PLACES AT THE SAME TIME AND EFFECTS THAT PRECEDE THEIR CAUSES
CHAPTER 6. THE NINE LIVES OF SCHRODINGER’S CAT
CHAPTER 7. I CHOOSE WITH THEREFORE, I AM
CHAPTER 8. THE EINSTEIN-PODOLSKY-ROSEN PARADOX
CHAPTER 9. RECONCILIATION OF REALISM AND IDEALISM
PART III. SELF-REFERENCE: HOW ONE BECOMES MANY
CHAPTER 10. EXPLORING THE MIND-BODY PROBLEM
CHAPTER 11. IN SEARCH OF THE QUANTUM MIND
CHAPTER 12. PARADOXES AND COMPLEX HIERARCHIES
CHAPTER 13. “I” OF CONSCIOUSNESS
CHAPTER 14. UNIFICATION OF PSYCHOLOGIES
PART IV . RETURN OF CHARM
CHAPTER 15. WAR AND PEACE
CHAPTER 16. EXTERNAL AND INTERNAL CREATIVITY
CHAPTER 17. THE AWAKENING OF BUDDHA
CHAPTER 18. IDEALISMAL THEORY OF ETHICS
CHAPTER 19. SPIRITUAL JOY
GLOBAR OF TERMS

20/31
Мы используем cookie-файлы для наилучшего представления нашего сайта. Продолжая использовать этот сайт, вы соглашаетесь с использованием cookie-файлов.
Принять