Gratis verzending vanaf €35,-
Unieke producten
Milieuvriendelijk, hoogste kwaliteit
Professioneel advies: 085 - 743 03 12

Announcing Essentia Books, an imprint dedicated to idealism

Announcing Essentia Books, an imprint dedicated to idealism

Reading | Philosophy

The editors | 2023-07-16

Vintage,Books,On,Table,In,Library.

Today we are proud to announce the launch of Essentia Books, a new imprint. Through it, we will be publishing scholarly works relevant to metaphysical idealism, the notion that nature is essentially experiential. Among many other leading authors, we will publish the latest book by Federico Faggin, inventor of the microprocessor and MOS silicon gate technology, recipient of the National Medal of Technology and Innovation from U.S. President Barack Obama, and probably the most well-rounded idealist alive.

Essentia Books is a collaboration between Essentia Foundation and Collective Ink, a subsidiary of Watkins Publishing. Watkins handles production, distribution, marketing and sales, while Essentia Books does the editorial work and contributes to marketing. Essentia has full editorial control of the material to be published by the new imprint, and will use it to highlight important, though often overlooked, work in areas relevant to idealism.

Because Watkins is a for-profit company—and there is nothing wrong with that—the books will have commercial prices. The percentage of the proceeds that comes to Essentia Foundation will, however, be fully reinvested in the foundation’s activities, particularly in further marketing of the publications of the imprint.

If you are an academic or qualified scholar in areas relevant to our work, and you believe your material is sufficiently related to metaphysical idealism, consider submitting your manuscript to us via this link.

Our flagship book for the launch of the new imprint is Federico Faggin’s Irreducible. Federico is one of the greatest luminaries of high technology alive today. A physicist by education, he is the inventor of the microprocessor and the MOS silicon gate technology, both of which underlie the modern world’s entire information technology. With the knowledge and experience of a lifetime in cutting-edge fields, Federico now turns his attention to consciousness and the nature of reality, sharing with us his profound insights on the classical and quantum worlds, artificial intelligence, life and the human mind. In this book, he elaborates on an idealist model of reality, produced after years of careful thought and direct experience, according to which nature’s most fundamental level is that of consciousness as a quantum phenomenon, while the classical physical world consists merely of evocative symbols of a deeper reality. Irreducible will be officially launched on May 31st, 2024.

Irreducible

But already in a few days, on July 28, 2023, we will be launching our first book: Donna Maria Thomas’s Children’s Unexplained Experiences in a Post Materialist World: What children can teach us about the mystery of being humanHistorically, children’s inexplicable experiences—from telepathy and conversing with deceased relatives to out-of-body or near-death experiences—have been theorized through a traditional scientific lens, which may not have the explanatory power to account for such experiences. In the book, Thomas shares research that she and other scholars, past and present, have conducted with children and young people across the world. By placing children’s unexplained experiences and views about reality in the contexts of culture, consciousness and the nature of self, the book offers a middle-way for explaining these childhood experiences within post-materialist science and philosophy. Thomas suggests that children’s experiences could greatly contribute to a new paradigm for understanding the mystery of being human and the nature of reality.

ChildrensExperiences

Here are three more books currently in production.

 

How Life Arose from Mind: The mind-matter problem within the Neo-Darwinian materialist conception of nature

By Daniela Panighetti and Massimiliano Sorrentino

The purpose of this book is to show that the mind-body problem not only concerns the relationship between mind and brain, but rather our current conception of the very phenomenon of life. In fact, it concerns our conception of the whole of nature in materialist terms. The book addresses the Neo-Darwinian claim that, since a naturalistic explanation for the emergence and development of living organisms exists, or is in principle viable, living organisms are not what they appear to be: that is, they are not the product of a mind. We argue that such is not the case.

 

The Sapient Cosmos: What a modern-day synthesis of science and philosophy teaches us about the emergence of information, complexity, consciousness, and meaning

By James B. Glattfelder

Ever since the human mind awoke to its own existence, it wondered about its cosmic significance. By dispelling myths and religious convictions, science entered the arena of explanatory templates. A tectonic shift in understanding began when the mind started decoding the mathematical language of the universe. To this day, the technological prowess unleashed by scientific knowledge remains awe-inspiring. However, science excluded two crucial domains from its field of inquiry. Firstly, the formation of complex systems—especially metabolic structures—appears to defy the physics of cosmic disorder and decay. Then, the nature of consciousness itself was deemed unworthy of academic discourse until not too long ago. Furthermore, the adoption of physicalism as a metaphysical foundation for science was an enormously consequential choice, which today is erroneously seen as an integral part of the edifice of science. This book chronicles an ongoing paradigm shift affecting physics and philosophy. Consciousness is rediscovered at the core of existence, expressing an intrinsic yearning for cosmic complexity, while the fabric of reality is woven from threads of information.

 

A General Theory of Composition: Sound, synergy and meaning

By Aliona Yurtsevich

In exploring the nature of consciousness, one wonders whether it is an intrinsically vibratory phenomenon. For indeed, both ancient knowledge and recent discoveries in science indicate that all aspects of the world are vibratory, wave-like phenomena. There is thus a sense in which we are sound, waves born and inhabiting an inherently resonant field, a vibratory nursery that generously hosts and nurtures a broad spectrum of mingling frequencies. Borrowing an idea from ancient teachings—“Life is Music”—Yurtsevich argues that musicality is an essential quality and condition of life. This notion can be empirically developed through musical practice, and then expand its influence further into other fields of creative activity. Yurtsevich further argues that, to understand the nature of living consciousness, one must engage in a sensory-embodied practice, such as sound and music. From the insights gained through her personal practice and life-long work with sound, it is her strong belief that sound brings all aspects of reality together into one irreducible existence—the ‘Unity of One Sense’—which is our composed world.

Simulated selves in a simulated world

Simulated selves in a simulated world

Reading | Philosophy

Felix Haas | 2023-07-09

beautiful woman drawing self portrait with pencil

The world we think we live in is but a projected tiling of our own concepts onto the scaffolding of external reality, and our personal selves mere illusions, argues Felix Haas. This is a thoroughgoing physicalist essay, which may raise the question: why is a foundation dedicated to idealism publishing it? We think it is an interesting piece for several reasons. In providing an informed ‘grand summary’ of the latest physicalist thought, this essay touches on important commonalities between modern physicalism and idealism. Amongst these are the notion that the personal self is an illusion, and that the world we perceive is as much a projection of our own conceptual categories as it is a standalone external reality. The essay also allows us to extrapolate the direction in which physicalism thought is evolving, as well as the ethos of that progression. Key to it is a kind of deconstructivist approach to self and world, in which our normal intuitions about what we are, and what the world is, are called into question. Importantly, this is the same ethos that underpins idealism. Our invitation to you, thus, is to question whether this essay goes far enough in addressing the very doubts it raises. Does it consistently and consequently pursue its own premises and conclusions to their ultimate implications? In our view, a truly self-consistent, critical investigation of how much we project our own concepts onto reality must call into question the notion of matter as a purely quantitative, empirically inaccessible ontological category. Might we be inadvertently projecting our own abstractions—our very concept of matter—onto reality as well?

That voice in my head that does the talking, that narrator of my thoughts and impressions, that is me. I exist. Every day I decide, I act, I experience and feel. This sensation of ‘me,’ my self, is real. Focusing only on what it is saying, rather than what our self is, we usually experience it as something like Descartes’ ghost in the machine. It is not surprising, then, that philosophers and religions over the millennia have invented daemons, souls, or the atman to inhabit our bodies and to be us. Even if you count yourself an atheist, who claims to disregard any kind of metaphysics, you might still believe your self to be some free-thinking, free-willed entity, distinct from your body, inhabiting it, rather than being part and product of it.

My aim here is to shed doubt on this notion.

But, before we get to the core of the argument, let me start by first trying to poke some holes into your everyday experience of your self. My hope being, that this might leave you more susceptible to what is to follow.

 

The self is not immutable and ever-present

It is difficult to give a satisfactory and exhaustive definition of what our self is. And still, we all have a seemingly intuitive understanding of it. Our sense of self is what we most identify with—that sense of being a small homunculus sitting behind our eyes, being the experiencer of our experiences, the thinker of our thoughts, the feeler of our feelings, the decider of our actions. Attempts at more analytical definitions of the self often include dissecting it into its different qualities. The neuroscientist Anil Seth, in his book “Being You” [1], for example, distinguishes the embodied, perspectival, volitional, narrative, and social self. Many scientists and philosophers, alongside Seth, have stressed that each of these facets are not constant and immutable, but can be altered. This is true even in everyday life.

We are all familiar with situations where the self becomes significantly less prominent. We speak of ‘losing our-self’ in work, of being in a flow state of high focus, fully identifying with a problem or activity, where our self becomes translucent. Likewise, our embodied self does not always inhabit the same boundaries. I am not referring to us growing or gaining weight, but the way we feel the boundaries of a car whilst driving or how we feel the limits of a racket when playing tennis. We learn to extend our sense of embodied self for tool use, helping us navigate complex situations.

Beyond everyday life, the perceived qualities of our self can be severely manipulated in experiment or through neuronal damage. Allow me pick some of Seth’s strata of the self and exemplify how readily hijacked they can be.

The Embodied self: Next to tool use, the ‘rubber-hand illusion’ is arguably the most canonical example. To create this illusion, a healthy subject’s left hand is hidden from his visual field and, instead, a rubber hand is placed in front of him. When an experimenter then starts brushing the rubber hand, interestingly, after a short while, the subject will start feeling the strokes he sees the rubber hand receiving.

The Perspectival self: Autoscopic hallucinations are situations in which the subject’s sense of the location of her self is not aligned with the location of her body. Out-of-body experiences (OBEs) are the most extreme examples in this category, where the subject experiences her self leaving her body and seeing it from the outside. Many cases of OBEs are documented, and neuroscientists have even been able to induce OBEs by electrically stimulating a patient’s brain [2]. Furthermore, OBE-like sensations have also been created virtually using cameras and virtual reality goggles.

The Narrative self: There are several documented cases of extreme amnesia. One of the most striking is the case of Clive Wearing whose hippocampi were left severely damaged after a brain infection. His condition has left him unable to experience his self as extended over time. Internal and external events that lie more than a few seconds in the past remain forever lost to him.

More astonishingly still, not only can the various facets of the self be significantly reduced, but total dissolution of any sense of self is also possible. One of the central objectives of Buddhism—particularly pronounced in Dzogchen—is trying to help its practitioners come to their own subjective experience of the absence of self, seeing their own consciousness devoid of self.

Outside of meditation, ritual dance, and hallucinogenic drugs, which can all induce a temporary absence of self, patients suffering from a psychiatric disorder called ‘Cotard’s syndrome’ seem to permanently lack any significant notion of self. As a result, they may stop using first person pronouns or even deny their own existence.

However, its reduction and dissolution are not the only arguments that call the fundamental nature and immutability of the self into question. It also seems possible to divide the self. This is the case in so called split-brain patients. These are people who had their corpus callosum, the neuronal bundle connecting the brain’s two hemispheres, severed either by accident or surgery (e.g. to treat forms of severe epilepsy). Both patients’ reports and experiments show that the two hemispheres of split-brain patients begin to function as independent minds.

 

What does it mean when we say the self is an illusion?

Few would doubt that our sense of self, and our experience of its qualities, is real. So, what do people mean when they claim that our self is an illusion? As Sam Harris puts it in his book “Waking Up” [3], “like many illusions, the sense of self disappears when closely examined.” It is the idea of our self as the originator and audience of our thoughts, actions, feelings, which Harris and others have called into question. Meditation often starts with the realization that thoughts enter and leave consciousness without there being anything like a thinker. In this way, you can see your self to be but one concept—albeit a seemingly pervasive and pivotal one—amongst many populating your consciousness at any point in time.

Some people, Anil Seth amongst them, have taken issue with labelling the self or its qualities ‘illusions,’ rightfully pointing out that it is just as real as ‘redness’ or ‘blueness’ [1]. Yet, both Harris and Seth would likely agree that the self is constructed (that is why it can be deconstructed) by our brains, and that it does not exist independently of them. Importantly, in being assembled by our brain, our self is not set apart, but in line with nearly everything else we experience.

The world we experience, our phenomenological world, is not the way the world really is. There is a map between the two, which must preserve certain primary aspects of reality, such as the volume and elasticity of objects, their relative positions, or velocities. However, qualities like ‘redness,’ ‘chairness,’ ‘sweetness,’ and many other dimensions in which we experience our world, live only in our internal models of the world, not in the world itself. We experience the world in a manner that is useful to us, not in the way it really is.

Things in the world have no smell or taste, yet the molecules they emit give us clues about how edible, dangerous, or ready for mating they might be. Humans, for instance, have five senses—six when adding balance. Other animals have different sets of senses, some of which are completely alien to us, such as echolocation (e.g. bats) or electroception (e.g. sharks). And even when considering the same category of sense across species, different instantiations of this sense are bound to create different internal worlds. It is how a perceived object scores against an individual’s four F’s—fighting, fleeing, feeding, and sex—that determines how attractive or repulsive it is constructed by that individual’s brain.

Historically, we think of perception as happening outside-in: The world impacts our senses, which send signals to different brain networks that ultimately decide what it is we are seeing. However, this view is being challenged (see e.g. [4]) and gradually replaced by an opposing, inside-out view: Our brain predicts what it is we might see in the next moment, which then gets validated against incoming sensory data. We can only see a car or a laptop if our brain already knows the concepts ‘car’ and ‘laptop.’ We would always be able to see the object that is a car. But, if we had never seen, heard of, or experienced any other mode of transport but our own two legs, we would only be able to recognize a car’s shape, size, color, etc.—we would not see any of the essence of what makes a car a car. In short, we would not see a car, but only a hallow block of painted metal.

Cognitive neuroscientist Lisa Feldman Barrett extends the idea of knowing a concept as a prerequisite to perceiving it far beyond the outside world. She understands interoception, perceiving our bodily states and signals, working in precisely the same way. Our brain reads bodily data like our breathing or heartbeat and uses socially shared emotional concepts (anxiety, love, hate, Schadenfreude, etc.) to predict what emotional concept best fits our state in each context. Deciding for a matching emotional concept, according to Feldman Barrett, is feeling that emotion. The concept comes before the emotion—we can only feel betrayal, if we already know the concept ‘betrayal.’

Concept building and concept matching, to Feldman Barrett, is also what gives rise to our experience of the self. “In my view,” she writes in her book “How Emotions Are Made” [5], “the self is a plain, ordinary concept, just like ‘Tree’. It is a goal-based concept in which the goal shifts based on context.”

Both Seth and Feldman Barrett paint our minds as predicting machines that read out signals from extero- and interoception, as well as from our brain itself, and match them against evolving concepts to decide what is in the world and what is ‘in our hearts,’ i.e. our feelings and desires. However, this decision lies not with our self—rather, this process runs almost entirely subconsciously. Our self is a product of this prediction-construction process, not its origin.

We are brains in bony cages who receive electric signals from our sensory organs. We then build models of the origins of these signals and call it the world. And we treat ourselves no different, because building our world model with us in it is the best that we can do. Our body, our feelings, our sense of self are model-elements that represent best guesses at what there is—with the word ‘best’ not being meant ontologically, but pragmatically. That is, our world-model is not constructed to be closest to the way the world really is, but such that we maximize our ability to navigate it and to achieve our goals of survival and reproduction.

Cognitive psychologist Donald Hoffman has long argued (see e.g. [6]) that a core idea of classical evolutionary epistemology is mistaken—namely, that evolutionary fitness would naturally drive our internal models progressively closer to the world as it really is. Hoffman, who has devised numerous experiments and simulations to support his claim, chooses the desktop metaphor to illustrate his point. We know that the computer desktop is not literally how our computer is organized or functions. Humans have designed and built all its hardware and software, none of which contains a literal desktop. The desktop is merely an abstract representation layer that enables the human user to navigate it more easily. Hoffman sees our internal model’s relationship to the outside world in much similarity to the desktop and the computer: a relationship that does not optimize for faithful representation, but for ease of navigation.

 

Why don’t we experience the world as a model?

If the world we experience is merely a model that our brains create, why don’t we perceive it as such, as a model? Why, instead, is our world-model what philosophers call ‘transparent’? That is, why does everything seem so real, with no scaffolding showing? Why don’t we, for example, perceive the process of how our brains decide whether something has ‘redness’ or not?

Thomas Metzinger—in his book “Ego Tunnel” [7]—answers this question by pointing towards evolutionary fitness. What would be the consequences, he asks, if our model of the world was opaque, and we could see the workings of the model itself? First, to not be significantly disadvantaged compared to a brain producing a transparent world-model, a brain producing an opaque model would still need to construct the same world content, but face additional metabolic expenditure to create whatever additional impressions represent the workings of the model itself. This is contrary to the basic strategy of how our brains evolved, trying to automate as many processes as possible, running them subconsciously at lower energy requirements.

But, even if energy were not an issue, our attention and ability to act would likely be impaired by an opaque world-model. If we no longer perceive only the content of our model, but also its workings, we might often focus on the latter, without gaining any evolutionary advantage. Our attention might be captured by observing how our model creates our phenomenological world, rather than focusing on model content, which alone influences our survival and procreation abilities.

 

What about free will?

The fact that an individual’s actions can be meaningfully influenced is what social media and online marketing build their business models around. More strikingly still, people—after having been subconsciously coerced into a specific action—will often concoct explanations for why they believe they chose what they did not [8]. What is not in itself a watertight argument against free will, does at least prove that we are able to convince ourselves that we are completely free to choose even in situations where we are not. It shows the propensity of our mind to invent volition.

Our biochemistry, genes, culture, and social surroundings, even our microbiome, are part of a long list of factors that influence our actions. However, some might want to brush this off as shaping our ‘character’ and ‘mood,’ rather than limiting or even negating our free will.

There are neuroscientific arguments, most famously the Libet experiment, in which the build-up of a so-called readiness potential in a subject’s brain is used to read out the subject’s imminent choice before they themselves become aware of it. However, variations of this classical experiment by other scientists [9] have raised some doubts about the original interpretation. Other defenses of free will call Libet’s classical interpretation into question by distinguishing between choosing an action and becoming aware of the choice. For some, philosopher Daniel Dennett [10] amongst them, such an implicit redefinition of ‘free will,’ which includes our subconscious, circumvents the Libet experiment being the last nail in the free will coffin. Others, counting Sam Harris [11], question whether this flavor of compatibilism is still talking about the thing that the free will debate set out to discuss.

The argument against free will, however, which holds the most punch, is that of determinism. As far as we know today, the future is determined by the past plus intrinsically random quantum fluctuations. In such a world, in our world, how do you insist on free will without violating the laws of nature?

Compatibilists of free will and determinism have more than struggled to find a convincing loophole to get to have their cake and eat it too. Typically, their arguments involve redefining the terms ‘free will’ or ‘self,’ along the lines of what we just discussed, or trying to find agency in quantum randomness. The former arguably fails to address the original problem. The latter lies outside of any accepted science or interpretation thereof.

Even without evoking our prior discussions on the predictive nature of our minds and the nature of the self, the volitional self and the concept of free will are near impossible to vindicate. Adding our prior conclusions to the discussion, does not make the compatibilist’s challenge any easier (what is free will without a self?). “We project causal power into our experience of volition in just the same way that we project redness into our perceptions of surfaces,” Seth writes in “Being You.”

 

Why does it all matter?

We typically meet attacks on ideas that give meaning or identity to our existence with hostility. The idea of our self as our brain’s simulation, lacking coherence and free will, not only goes against most of our everyday experiences but, put very simply, hearing it doesn’t feel very good. However, neither our emotional resistance, nor our intuition are fair arguments for keeping with these notions. My liking something or not has no bearing on its truth. And there is an abundant list of what we now accept as fundamental aspects of reality where our intuitions have led us astray for millennia—from the idea of a spherical earth and the heliocentric worldview to the ontological insights of quantum mechanics and general relativity.

But why does it all matter? Unlike quantum mechanics and general relativity, the realization of the true nature of the self is unlikely to allow us to build next-generation machines. Rather, you may worry that the absence of free will negates moral responsibility. Much has been written (see e.g. [11] for a short introduction, or [12] for an extensive treatment) on why this is not the case, which I do not intend to reproduce here. However, if our will indeed is not free, then it has always been this way, and so, it might suffice to ask, why our ethics and justice system should disintegrate with the realization of this truth?

On the other hand, realizing the illusionary nature of free will does further enable us to build criminal justice systems around the principles of deterrence, security detention, and rehabilitation, rather than revenge.

Little collective or individual happiness has come from the agrandizement of one’s nation, race, or self. Instead, psychology and self-help literature is ripe with arguments for deflating our ego and self-focus and for finding meaning in something larger than ourselves. After realizing that you are no longer—and never were—the all-controlling CEO of your body and actions, you may choose to be liberated rather than impoverished. Rather than seeing the self-simulation as a bleak truth, you can choose to take it as a call for further introspection and first-person exploration of the structure and possibilities of your own consciousness. You may realize that the fact that all your perceptions, including your own feelings, are constructed by your brain, means that it is not the world that angers you or makes you happy, but your own mind. You may also come to find a greater deal of intellectual humility and empathy. You may be less certain of the finality of the conclusions you draw and thus more receptive to criticism. Likewise, you may be more careful and forgiving when believing that you have spotted errors in others’ conclusions or actions.

Finally, you may find beauty in the idea of how, in an expanding, slowly cooling universe, complexity built over time to give rise not only to life, but eventually also to conscious minds. Through us and other minds before and after us, the universe started modelling and understanding itself. We are not divine souls, beings from a different realm that inhabit our physical bodies, but literally all we are is the stuff that the universe is made from.

 

References

[1] Seth, Anil (2021). Being You: A New Science of Consciousness. Faber & Faber.

[2] Blanke, O. et al (2002). Stimulating illusory own-body perceptions. Nature 419, 269-270.

[3] Harris, Sam (2014). Waking Up: A Guide to Spirituality Without Religion. Simon & Schuster.

[4] de Lange, F. P., Heilbron, M., and Kok, P. (2018). How do expectations shape perception? Trends Cogn. Sci. 22, 764–779.

[5] Feldman Barrett, Lisa (2017). How Emotions are Made: The Secret Life of the Brain. Houghton Mifflin Harcourt.

[6] Hoffman, Donald D. (2019). The Case Against Reality: Why Evolution Hid the Truth From Our Eyes. W.W.Norton & Company.

[7] Metzinger, Thomas (2009). The Ego Tunnel – The Science of the Mind and the Myth of the Self. Basic Books.

[8] Robson, David (2015). The hidden tricks of powerful persuasion [online]. BBC Future. Available from: www.bbc.com/future/article/20150324-the-hidden-tricks-of-persuasion

[9] Schurger A, Sitt J, Dehaene S (2012). An accumulator model for spontaneous neural activity prior to self-initiated movement. Proc Natl Acad Sci USA 109: E2904–E2913.

[10] Dennett, Daniel (2003). Freedom evolves. Penguin.

[11] Harris, Sam (2012). Free Will. Free Press.

[12] Parfit, Derek (2011). On What Matters, vols. 1 and 2. Oxford University Press. and Parfit, Derek (2017). On What Matters, vol. 3. Oxford University Press.

How a neuroscientist came to embrace the reality of acausal synchronicities

How a neuroscientist came to embrace the reality of acausal synchronicities

Reading | Metaphysics

Laleh K. Quinn, PhD | 2023-07-02

Mystical,Background,With,A,Woman,For,Astrology,,The,Girl,Stands

In this courageous and very personal essay, UC San Diego neuroscientist Dr. Laleh Quinn shares with us her journey from materialism to opening up to the possibility of acausal, transpersonal, mind-like organizing principles in nature. The events that led to this opening-up will amaze you. We salute Dr. Quinn for her sincerity, integrity, and commitment to truth.

In Jung’s seminal work, “Synchronicity,” he argues for a fourth organizing principle—alongside space, time, and causality—at work in the universe. As Jung experienced himself, there are some events experienced by humanity that cannot be explained on the basis of the first three principles, and thus there must be a fourth. This fourth is what Jung referred to as “synchronicity.” He defined it as an “acausal” principle, in that it did not seem to be governed by causation as normally defined within the standard scientific model.

Jung describes several instances of synchronistic experiences that could not be explained through physical causation. One of the most famous of these examples is that of the Scarab: while tending to what he refers to as an overly rational patient in his office, who had yet to break through to the deeper reality of her psyche, she related a dream she had the previous evening about a piece of jewelry in the shape of a golden scarab. Just at that moment, a scarab with green and golden hues flew into the office. Jung famously stated: “here is your scarab.” The patient was immediately lifted out of her narrow vision of the world.

This example is noteworthy, as it seems to violate what may be acceptable as normal coincidence. Coincidences happen all the time, but they vary in their probability of occurrence. What lies at the heart of Jung’s insistence that at least some coincidences cannot be explained by mere chance is a seeming violation of such probabilities. It may be a coincidence that my friend and I both decide to wear red shirts and dark jeans on the same day; however, discussing a dream of a golden scarab and immediately witnessing the appearance of a golden scarab—especially if those appearances are extremely rare, as they were in Jung’s environment—seems to point to another level of ‘coincidence.’ For Jung, these are the true synchronicities. There is something in this conjoining of events that goes far beyond our normal understanding of the way the universe works; far beyond normal chance occurrences governed by the laws of physical causation. After all, the discussion of a dream certainly did not cause the scarab to arrive at the window. Moreover, this synchronicity was a highly meaningful one for the patient.

Many of us have had these types of extraordinary experiences in our lives, but we tend to either pass them off as normal coincidences or, at best, delegate them as anomalies that must have a proper scientific explanation that we don’t yet comprehend. Experiences such as thinking of someone not thought of in years and having them call the next day have probably happened to most of us, but if you’re like me—at least the former me—I would have a twinge of excitement about coincidences like that and then promptly forget about them. For Jung, however, attending to such synchronicities goes far beyond feelings of excitement. To him, doing so provides us access to one of the most important forces of the universe, and to an understanding that, at the heart of everything, is meaning. As he states in “The interpretation of Nature and the Psyche,”

Synchronicity … means the simultaneous occurrence of a certain psychic state with one or more external events which appear as meaningful parallels to the momentary subjective state. ( p. 36)

To get to this place of understanding the meaning that underlies reality, we need to somehow overcome the natural bias we have when it comes to belief in an acausal principle. For a lot of us, including myself for most of my adult life, I would not have taken the scarab incident to prove anything mystical or meaningful. I, like many others, would maybe find it initially fascinating and interesting, but then would return to normal life; a kind of mental sweeping it under the rug. Somehow that’s easier than having to accept a new worldview. But, for Jung, sweeping it under the rug and returning to a comfortable normalcy entails missing the entire point of existence.

So, how can we better have access to the level of meaning that true synchronicities are capable of providing us? How can we get past the skeptical blocks that are our natural tendency? For me, as a neuroscientist steeped in the scientific method and rational materialism, it has been a difficult road. It took a deep personal tragedy for me to begin to experience and pay attention to the immense significance of synchronicities, and to come to have faith in their legitimacy. If I hadn’t had the personal experiences I had, I would most likely still be lacking in that faith.

Soon after my best friend and colleague of twenty years died, I began having very unusual experiences that made me suspect that he was somehow still around me. As a skeptic with lingering distrust of people who claim to be able to talk to the dead, this was—to put it mildly—out of my comfort zone. And yet, I was extremely curious. Could he really still be here in some way? I began to perform my own experiments to prove to myself he was. This was accomplished in a few ways. One was to ask him, upon waking up in the morning, for images in my mind that I would encounter later in the day. The first time I did this I saw an image of an old Wright Brothers plane. When I went running later that morning, I looked up at the rooftop of a house I had run past many times and there, attached to the roof, was a small replica of exactly that type of plane. While this was impressive, I was still able to question whether the information came from him. Maybe I’d subconsciously seen that plane but forgot, and I put that image in my mind myself. I had other experiences like this one but I was still not convinced. To be more sure, I tried a different method: asking for specific signs from him to let me know he was still here.

I was overwhelmed by the response. There were so many shocking moments where I would ask for something particular to be placed in front of me and I would receive it. Still I was hesitant to shake my materialist way of understanding the world. But one chain of signs I received finally resulted in the relaxation of my logical mind and allowed myself to jettison the last vestiges of doubt: I was driving north from San Diego to the Bay Area to visit my father in hospice. While on the highway, I asked my dead colleague to send me a sign that he was around me. I told him it didn’t matter what the sign was, just something I would understand.

Just after I asked, I heard some clicking sounds, then the song “Hello it’s Me” came on the radio. That jarred me. Was he saying hello? I wasn’t sure. Needing more evidence, I asked for further signs. Just then a car passed me with a license plate that said Camelot on it, where instead of the ‘e’ there was a heart. This was actually very significant to me for two reasons. First, I knew that Camelot was associated with King Arthur. This reminded me of the nickname (Wort) that my father—who loved the Arthurian Legends—gave me, which was the nickname Merlin gave King Arthur when he was a child. And second, it also reminded me of the last time I had visited my colleague and he took me to his favorite baking shop, “King Arthur’s flour.” This was all deeply meaningful to me, but I wanted yet more proof. So to clinch it, I asked him out loud, “if that was you I want the next song on the radio to have the word ‘knight’ in it.” Amusingly, in my mind I was thinking of the song “ Knights in white Satin,” (amusing because later I came to know it isn’t “Knights” but “Nights” in white satin). So there I was, waiting for “Nights in white satin” to miraculously play on the radio, but it didn’t happen. Instead, the very next song that came on was a song by Neil Young that I hadn’t heard for many years, called “After the Gold Rush.” As soon as it started playing, I had a chill and a knowing: the first line of the song is “Well, I dreamed I saw the knights in armor coming sayin’ something about a queen.”

I can’t describe the feeling I had at that moment; that absolute thrill of discovery and understanding; the sense of expansion and excitement. This was not something I could ignore or deny or sweep under the rug. More came, but it felt like icing on the cake. I passed two trucks on the way that were “Knight Transportation” trucks. Then, when I arrived at my family home, two books were placed facing out on the bookcase of the room I was sleeping in: both had “Knight” in the title. Next to those two books, also facing out, was Richard Feynman’s book “The Meaning of it All.” At that point I was floating, somewhere between our world and theirs, feeling the magic and wonder I didn’t allow myself to previously feel. No amount of rationalization could penetrate this new knowledge. The probability of these chains of signs—and the specificity of the signs after having asked for them—being due to mere coincidence was astronomically low. As a scientist utilizing statistical methods to show the likelihood of hypotheses, I knew the likelihood that what occurred was due to normal causation was at least one in a trillion. Most likely much, much more improbable than that. And as a good scientist should, I accepted that my working hypothesis that my colleague was sending me a message, indicating he was still somehow ‘here,’ was validated.

I consider these types of experiences to be the epitome of synchronicity, as Jung would have us understand it. Not only the improbability of this chain of events, but the deep meaning behind them, both with respect to my deceased friend and my soon to be deceased father, was profound. But I also understand that anecdotal evidence is very hard to accept. I wouldn’t have accepted it if it hadn’t happened to me. The implications are too great, if wrong. It opens up the possibility of there being a universe that is not governed by rational laws as we know them. It opens up the possibility of parapsychological phenomena being true.I deeply understand the implications of allowing oneself to open to a vast, seemingly nonsensical reality. It’s as if the floor has been taken out from under us, everything we used to ‘know’ thrown aside. But I came to understand that this can also be a great opportunity; an opportunity to willingly suspend our disbelief for a moment and consider the other possibilities, and whether those possibilities are worth the risk involved. There appears to be a choice here that we should be considering prior to dismissing Jung’s principle out of hand.

Is there a way to help with making this choice? Blaise Pascal, a brilliant French mathematician and catholic writer living in the 17th century, grappled with how to make belief in God a rational choice. When I first started studying philosophy, I was taught about Pascal’s famous wager. This intrigued me, as I was brought up in a religious household but rejected organized religion from a very early age, feeling even then that ‘faith’ wasn’t good enough a reason to believe in something. Pascal provided a kind of calculus, or cost benefit analysis, in addition to faith to help those who require more. He provided a matrix of possibilities having to do with belief in God, and asked us to consider these four options: (1) God either exists or (2) doesn’t. And we either (3) believe in God or (4) don’t believe. The ramifications of the options, according to Pascal, are shown in the table below.

Here it seems clear that the ‘rational’ choice is to believe in God, whether or not God exists. Weighing the pros and cons, the decision became obvious to Pascal. Some wasted time and effort is not worth eternal damnation. While Pascal’s wager had a definite fearsome catholic spin, and may not seem applicable to many of us, we can still use it as a way to delineate the possible scenarios under our purpose of belief in Jungian acausal synchronicities. Assume our decision matrix consists of the following: (1) synchronicities exist, (2) synchronicities don’t exist, (3) we believe in them, and (4) we don’t believe in them. How might we fill in our matrix?

First, there is the situation where we believe in sychronicities but they don’t exist. What is the cost of believing that ‘coincidences’ are meaningful when in fact they are not? This may lead to a life of succumbing to some form of magical thinking when there is no magic. We may suffer ridicule from our peers, or worse, from ourselves. This is the situation that, as an academic, I used to be very fearful of. Belief in something ‘nonsensical,’ or ‘irrational,’ may indicate that I’m less intelligent than my peers, or at best mentally unstable, which for many academics is to be preferred over lack of intelligence. Many great thinkers, even Jung himself, struggled with the fear of irrationality.

Now, what happens if synchronicities don’t exist and we don’t believe in them? We may have a sense of being vindicated, knowing we haven’t been gullible and have correctly adhered to a rational state of mind.

But let’s assume synchronicities exist and we do believe in them. This could lead to a deep understanding of reality where we have tapped into something beautiful and meaningful, providing our lives with a depth of knowledge and understanding we couldn’t have otherwise.

Finally, if synchronicities exist and we don’t believe in them, this may result in our having lost access to the true meaning of our lives.

It’s one thing to be told that the rational choice is to believe in something; it’s another thing to actually believe it even if you want to. When asked how one is to make oneself believe if they don’t already, Pascal’s advice was to fake it at first. Spend time with religious people, pray to God, act as if it’s true and ultimately it will become second nature. Perhaps we can take his advice and ‘fake’ it at first, and see what happens. One way to get started is to begin to properly research and do experiments of your own, like I did. Understand that dismissal of a hypothesis out of hand is anti-scientific. If you really want to know the truth, some dedication to rigorous research on the topic is required.

It’s ironic how many materialist scientists have not actually researched the ideas that they are lambasting. Hypotheses concerning acausal phenomena such as synchronicity, ESP, telepathy, mediumship, psychic abilities of any sort, for this type of person, are not worthy of proper analysis. They are not worth analysis at all. To them, they are wrong, a priori. As a scientist, I find this unacceptable if we are truly searching for truth. So jump in and at least make an attempt at discovery, even if it’s difficult at first. Addressing the possibility of synchronicities in a logical, rational way might just lead you to discovering they exist and, thus, provide you with the beginnings of a much more meaningful existence. Open your eyes and mind to a new understanding and see what Jung understood, that “synchronicity is an ever present reality for those who have eyes to see.”

The missing subject: a critique of Philip Goff’s panpsychism

The missing subject: a critique of Philip Goff’s panpsychism

Reading | Theology

Brain science, intelligence and genius education concept, smart head profile with space inside, the future and ai,Element of the image provided by NASA

Rev. Joshua Farris makes the case that, once we accept that experiential qualities are irreducible to physicality, we become logically committed to subjectivity itself as the fundamental substance at the foundation of reality, distinct from physicality, and in terms of which we can account for physicality. Notice that Rev. Farris uses the term ‘naturalism’ in a manner distinct from our own usage. For us, ‘naturalism’ means simply that nature behaves spontaneously, according to its own intrinsic regularities, as opposed to deliberate planning or external intervention. In this sense, naturalism and idealism are entirely compatible. In Rev. Farris’ usage, however, ‘naturalism’ refers to the physical, or material, phase of reality as determining its intrinsic structure. Neither definition is incorrect; they are simply different. But it is important to understand what is meant in each case.

To accept experiential qualities as fundamental aspects of an otherwise physical reality, as proposed by panpsychist philosopher Philip Goff, isn’t enough. We need more. We need subjects of qualities, and it’s dubious that panpsychism can fit the bill.

In one place, Goff makes the rather odd claim that “The main objection made to panpsychism is that it is ‘crazy’ and ‘just obviously wrong’.”[1] He proceeds to claim that we know something of the intrinsic nature of matter: “In fact, the only thing we know about the intrinsic nature of matter is that some of it—the stuff in brains—involves experience.”[2] Goff is convinced that his version of panpsychism steers a via media between physicalism and substance dualism. It takes seriously the nature of conscious experience like substance dualism, but without the sort of bifurcation of mind and body of substance dualism. It posits qualities as the intrinsic nature of matter and the latter, in turn, as the explanation for everything else. Yet, Goff also claims that his view is distinct from any version of idealism (whether it be theistic or involving a cosmic mind undergirding Nature).[3]

It is to the subject of qualities that I believe we must turn to as we reflect on the nature of the world; not just the qualities of life, but the subject of those qualities—something I and others contend is a challenge for panpsychists.

 

The historical progression of panpsychism

Defenders of panpsychism often trace their view back to Bertrand Russell, who affirms a form of monism concerning the physical and the phenomenal. Prior to Russell, there are traces of panpsychism in Leibniz as well—although Leibniz’s panpsychism is a form of idealism, entailing that minds furnish a fundamental structural role to the physical. Indeed, Leibniz departs in important ways from contemporary panpsychism.

The sort of view intimated by Russell differs significantly from Leibniz’s idealism in that it does not posit something as undergirding matter, but that matter itself is imbued with qualities, mind-lets, or mind-like properties at the fundamental level. This position on matter and consciousness, however, has only come to prominence in the last 20 years or so. It wasn’t until David Chalmers and Thomas Nagel that it made its way back into contemporary academic discussions.

 

The historical overlap between panpsychism and naturalism

Arguably, it is not coincidental that the historical rise of panpsychism overlaps with the conceptual desire to maintain naturalism as the framework for understanding the world. Those naturalists that take the mind seriously, do so because they realize that a physicalist ontology is unable to account for qualities, or ‘qualia,’ as philosophers put it. A quality is the felt experience of what it is like to, e.g., taste strawberry ice cream, smell the ocean air, or experience the beauty of a pink sky as evening turns to night. Felt experience is lost in a naturalist’s world; one that, as Goff argues, is the result of Galileo’s worldview, which attempted to understand the world in terms of quantities without qualities. It is here that panpsychism has an advantage over physicalist naturalism: at least we are able to re-introduce qualities where they are otherwise excised and unaccounted for. For the naturalist, it might be their last hope.

Even naturalist Annaka Harris (wife of the famous new atheist Sam Harris) recognizes the conceptual overlap between panpsychism and its historical origins with physicalists. Toward the end of her book, Consciousness, she offers little more than the common promissory note that there may still be hope for physicalism to make sense of consciousness. But she also realizes that it may turn out that something like Nagel’s or Goff’s panpsychism may rescue the naturalist by saving consciousness. She says: “while theoretical physicists can happily propose ideas such as the predictions of string theory—from ten (or more) dimensions of space to the vast landscape of possible universes—and still have their work get a fair hearing, it is considered a risk to one’s reputation to suggest that consciousness might exist outside the brain.”[4] This is a fascinating concession from Harris, and suggestively points the way forward for the naturalist. Nonetheless, is it sufficient?

 

Naturalist panpsychism

Just what is naturalist panpsychism? In short, it is naturalism with qualities; naturalism with what Nagel calls ‘subjective appearances.’[5] Nagel reflects on this move when he considers the insufficiency of physicalism, whereby physics is the only—or best—way of knowing the world. He states:

The existence of consciousness seems to imply that the physical description of the universe, in spite of its richness and explanatory power, is only part of the truth, and that the natural order is far less austere than it would be if physics and chemistry accounted for everything.

But in order to get qualities, there is the question of whether qualities can exist like physical particles, as part of the bedrock of a naturalist world. In other words, qualities of this sort would exist as a foundational feature or property of the natural world, without any additional reality undergirding them.[6] This seems problematic: qualities need what philosophers call a ‘substance’ (i.e., a substrate with standalone existence), such as a mind or subject of experience. Indeed, it seems incoherent to speak of qualitative experiences without a subject that experiences those qualities as modes of its mind.

 

We need a wholly distinct substance

A substance has been traditionally defined as a thing that is countable or able to exist independently of other things. In the language of philosopher Ralph Weir, “they exist by themselves.” Since qualities don’t exist by themselves, they need a substance of which they can be the qualities. This is where defenders of substance dualism have an advantage over the naturalist panpsychism of Goff. Substance dualism is the view that there are two distinct sides of being, as put in the language of the dualist Uwe Meixner. And these two sides are properties—with property-bearers—including the phenomenal and the physical. According to the advocate of substance dualism, there is a distinct type of particular necessary to make sense, and instantiate the fact, of phenomenal qualities. While Goff, in his post-Galilean way, is keen to preserve qualities, it is not clear how he can do so. Qualities themselves aren’t enough. They must exist in something that owns them, bears them, and conceivably knows them.

This doesn’t necessarily require a Cartesian substance dualism, although that is one option. As E. J. Lowe has so famously put it, the thing having the phenomenal qualities presumes a fact about the substance in those qualitative states that is quite distinct from physical parts. Lowe highlights the particularity necessary to make sense of personal identity:

[P]art of what makes an experience of mine numerically distinct from a qualitatively indistinguishable experience of yours is the very fact that it is mine as opposed to yours.[7]

The fact of your existence as an experiencing thing is required, it seems, for qualities. In other words, you can’t just add qualities to your system and leave it at that, precisely because doing so entails a distinct type of substance for those qualities. But if this is the case, Goff isn’t really offering a via media between physicalism and substance dualism; either his view leads to substance dualism or some form of idealism.

 

You can’t just add qualities to your system

Goff argues that his version of panpsychism is distinct from, e.g., analytic idealism: “The main difference is that whilst panpsychists think that the physical world is fundamental, idealists think that there is a more fundamental reality underlying the physical world.” But at least in the analytic idealist framework the individual consciousnesses of people—which are obviously contingent, as individual consciousnesses come into existence at particular times—are explained by a distinct type of substance and are grounded in some necessary and ultimate reality. As Ralph Weir has clearly shown, this understanding of nature doesn’t entail the muddy middle of panpsychism.

One of the crucial lines of argument Goff uses to defend his panpsychism leverages the conceivability of so-called ‘zombie twins.’ A ‘zombie-twin’ is taken to be a complete physical duplicate of a person, but without being a phenomenal duplicate; i.e., your zombie twin is unconscious from within, despite being physically indistinguishable from you from without. Commonly used as an objection to the view that humans are solely physical, the conceivability of zombie twins—and, therefore, their metaphysical possibility—seems to defeat physicalism: if your purely physical zombie twin is metaphysically possible, then there is something extra, distinct from physicality, that renders you conscious. For Goff, this ‘extra’ is the qualities that constitute the intrinsic nature of matter.

Yet, an entirely analogous argument can be used in defense of substance dualism, for we can also coherently conceive of a ‘ghost twin’: an entity identical to you as far as your conscious inner life is concerned, but devoid of any physical property. Insofar as the conceivability of a ghost twin entails its metaphysical possibility, there may be a substance in nature distinct from physicality, and therefore substance dualism may be true. Nonetheless, Goff rejects dualism.

Goff’s rejection here, despite the equivalence between the ‘ghost twin’ and the ‘zombie twin’ arguments (the latter embraced by Goff), is based on a rather technical and questionable dodge: the notion that mere individual identity can lead to non-logical differences across qualitative duplicates. Ralph Weir discusses the problems with this notion:

[I]t is implausible that mere identity can make non-logical differences to what is possible for qualitative duplicates. If it is possible for your zombie twin to exist with only its physical properties and no consciousness, then the same is true of that part or aspect of reality that your zombie twin duplicates. And if it is possible for your ghost twin to exist with only its phenomenal properties and no body, then the same is true of that part or aspect of reality that your ghost twin duplicates. The zombie argument is therefore safe from objections that require that mere identity can make non-logical differences to what is possible for qualitative duplicates. But so is the disembodiment argument for mental substances.[8]

The important challenge to Goff is that, if there is a metaphysical possibility for zombie twins—i.e., your physical duplicate without phenomenal consciousness—then your analogous ghost twin is a metaphysical possibility too. Goff’s commitment to qualities commits him to substance dualism, or—at a minimum—a wholly distinct substance of a phenomenal nature that accounts for physical bodies in terms of phenomenal qualities; in other words, some version of idealism that causally or structurally couches the reality of physical bodies in a mind or minds.

 

Theistic panpsychism

There is a smaller contingency of panpsychists defending theism, which underwrites panpsychism with a value-full system. This, however, is a departure from the historical rationale leading someone like Goff to posit qualities in an otherwise naturalist system. It is an odd departure given the recent historical reasons that gave rise to panpsychism. But then again, if panpsychism is just a retrieval of some version of substance dualism or idealism (e.g., Leibniz’s idealism), then it should not be considered a historical outgrowth of naturalist physicalism.

 

Theism: qualities aren’t brute but grounded in sufficient reason

In recent analytic theistic literature, one can notice a growing prevalence of the view that contingent, individual consciousnesses are grounded in some ultimate and necessary reality. Indeed, there is a growing body of literature contending that not only is naturalism insufficient as an explanation for consciousness, but that theism supplies the missing ground. This is certainly a move in the right direction, for as Joshua Rasmussen states,

Instead of positing mindless units beyond all experience, I propose that a first-person, personal reality is fundamental to all other realities. This mind-first picture is simpler and has greater explanatory power than the mindlessness-first picture, or so I argue. The mind-first picture also provides resources for solving the many construction problems, explaining nature and formations of matter, and explaining how there can be any being like us. For these reasons, I arrived at this theory of your ultimate origin: your origin is not based in impersonal, mindless stuff but in the ‘stuff’ of a personal foundation.[9]

Geoffrey Madell makes an argument for a substance that defies any conditioned analysis that can be borrowed from a naturalist or physicalist frame, and points us in the direction of some higher-order cause or ultimate explanation quite apart from the contingency of the physical and emergent consciousnesses:

There is, however, no denying that many people will see grounds for rejecting outright the account of the self which seems to be emerging from what I have said, and that for a fundamental reason. To suggest, as I appear to have done, that there are no criteria for identity of the self over time, and no criteria which have to be satisfied for a state of consciousness to be mine at any one time, leaves one with a sort of free floating ‘I’. On one hand, every attempt to establish criteria for the identity of the self, to tie it logically to some such condition as the continuity of the body or of psychological continuity, or its identity to the notion of origin, seems to break down. But to accept this is to give credence to the idea of the self as an entity which, purely as a matter of chance, alights on a certain set of properties in history but might equally have alighted on any other set. This presents a dilemma of awesome proportions, and we must eventually confront it.[10]

Michael Bitbol highlights this:

Why do I live now, in this special period of history? Why am I me, born in this family, in this place of the world? I was taught that there were many other possibilities: being any person, at any time, or even just not being at all. And yet here I am, in front of you. Me, not you, here, not there, now, not then…What is the reason, if any, of this inescapable singularity? Does the fact that we all live through this mystery alleviate it in any way?[11]

The deeper problem seems to be naturalism, as Farris points out:

This rather obvious truth is so plain that you might think there’s no need to say it. What we long for is the person. Persons are valuable to us. In fact, persons are most cherished above  all other things that we regard as valuable. Sure, there are pets that we love and hold as valuable. We value food, our jobs, our homes, and our cars. We like having things, but if we are honest, it’s persons that we prize more highly than anything else in the world.[12]

While no one would actually claim that technology or science could get in the way of that which is most highly cherished, we are seeing and hearing of developments that promise the possibility to accommodate all of our needs and wants through artificial means. The prospect of constructing individuals that we can interact with that appear to be flesh and blood persons is certainly something that is not outside the imaginative social consciousness of contemporary society… While this might seem a bit far-fetched to some it hits at the heart of what we care about most and what is at stake in the science-engaged theological conversations today.[13]

What these authors make clear is that a naturalist view of subjects is insufficient to account for qualities in the world. What is needed is both a distinct type of substance and a necessary ground for explaining conscious subjects. Theistic panpsychism is better equipped to handle these facts than the sort advanced by Goff, yet it isn’t an alternative to the traditional options (idealism and substance dualism); instead, theistic panpsychism is just a form of idealism. At the end of the day, we need a substance for qualities, which points us in the direction of some ultimate explanation for contingent consciousnesses. Some version of theism supplies that explanation in a way that Goff’s panpsychism does not.

 

References

[1] https://aeon.co/ideas/panpsychism-is-crazy-but-its-also-most-probably-true

[2] https://aeon.co/ideas/panpsychism-is-crazy-but-its-also-most-probably-true

[3] He is not committed to either reductive or non-reductive panpsychism (the latter being the view that there are differing levels of the mind governed by layers of natural laws). And, while, he leans in the direction of non-reductive panpsychism, he is not convinced that we have sufficient knowledge of Nature to make determinate that hierarchical minds could not be reduced to the underlying qualities or mind-lets.

[4] Annaka Harris, Conscious: A Brief Guide to the Fundamental Mystery of the Mind (New York: Harper, 2019), 81.

[5] Thomas Nagel, Mind and Cosmos (Oxford: Oxford University Press, 2008).

[6] Philip Goff, “A Conscious Universe: Panpsychism vs. Idealism,” Institute of Art and Ideas. https://iai.tv/aarticles/conscious-universe-panscyhism-idealism-goff-kastrup-auid-1584?_auid=2020; This is precisely the version Goff endorses when distinguishing it from the ‘analytic idealism’ of Bernardo Kastrup. Goff attempts to avoid theism as well as an ontological idealism that undergirds the natural world. The natural world just is conscious at the most foundational ontological level along with physical particles.

[7] E. J. Lowe, “The Probable Simplicity of Personal Identity.” In Personal Identity: Complex or Simple? edited by Georg Gasser and Matthias Stefan, 137-155 (Cambridge: Cambridge University Press, 2012), 149.

[8] Ralph S. Weir, “Can a Post-Galilean Science of Consciousness Avoid Substance Dualism?”, Journal of Consciousness Studies, 28, No. 9–10, 2021, pp. 212–28, see p. 221. [DOI: 10.53765/20512201.28.9.212]

[9] Joshua Rasmussen, Who Are You, Really? (Downers Grove: IVP, 2023), 268-269).

[10]  Madell, Essence of the Self, 10–11.

[11]  Michel Bitbol, as quoted in Nicholas Humphrey, Soul Dust (London: Quercus, 2011), 151–152.

[12] Ibid., 7.

[13] Ibid., 8.

From points to fluctuating strands: advancing theoretical physics

From points to fluctuating strands: advancing theoretical physics

Reading | Physics

A microscopic view of a simple plaited textile - 3D render

To advance, theoretical physics must stop thinking in terms of points (be it point-sized things like particles, or points of space) and, instead, think in terms of the configurations and fluctuations of strands constituting the underlying fabric of reality, argues Dr. Schiller. While his views still preserve distinct strands, we believe it is a step forward towards a holistic understanding of nature as the behavior of an extended substrate, as opposed to the dynamics of discrete things.

Overview

In contrast to the progress in experimental physics and in all other sciences, progress in fundamental theoretical physics has been slow for the last five decades. Why were so many efforts in vain? There are six reasons. The first reason is the insistence to use points to describe particles and space. Five additional habits of thought slow down fundamental theoretical physics even more.

The present essay argues that all six delaying habits can be eliminated in exactly the same way: points need to be substituted by a different kind of fundamental concept. The new concept is provided by the common constituents of space, black holes, particles, wave functions, and forces. Distilling the correct constituents allows answering the open questions of fundamental physics, including the origin of all colors in nature.

 

The need for a change in fundamental physics

All of modern science, from Galileo to today, can be seen as the quest to describe with precision everything that is observed. Now, in all observations, something moves. The precise description of motion is the subject of physics. Modern physics can be seen as the quest to describe with precision everything that moves. In fact, only two aspects of nature are found to move: particles and the curvature of space.

Due to its drive towards describing particles and space to full precision, science has made—and is still making—steady progress in all its domains, from material science to chemistry, from the geosciences to astronomy, from physics to astrophysics, and in all the life sciences. As a whole, because of this drive, science is a fascinating success story.

In the success story of science, there is one exception. For the last five decades, fundamental theoretical physics—in contrast to the other parts of physics—has stagnated. All known fundamental properties of nature were described in 1973. They agree with all experiments. Nevertheless, there are important open questions. For 50 years, the origin of the interactions, the origin of the elementary particles, and the origin of the fundamental constants remain unknown. For example, there is no explanation for the origin of electricity and radioactivity, no explanation for the origin of electrons and quarks, and no explanation for the strength with which electrons interact with photons. This strength is described by a number, the fine structure constant, measured to have the value 137.0359992 [2], but whose origin is unknown. Because the origin is unknown, also the precise origin of every single color in nature is unknown.

Why is the origin of colors unknown? Why are the origins of particles, forces and their strengths still unclear, despite a large number of researchers working on them? In the following, it is argued that six reasons prevented progress:

  1. The habit of using point particles and points in space.
  2. The hesitance to conceive a model for black holes.
  3. The hesitance to conceive a model for particles and wave functions.
  4. The hesitance to conceive a model for space.
  5. The habit of looking for a unified equation.
  6. The hesitance to conceive a model for forces.

Interestingly, the solution for each impasse is the same: nature must be described with a new type of fundamental constituents.

 

Points do not exist

The two great achievements of twentieth-century physics—general relativity [1–3] and the standard model of particle physics, extended with massive Dirac neutrinos [4]—describe nature with full precision. Even though hints of deviations arise regularly, no hard data contradicts either of the two theories. Despite huge incentives to find deviations, none has been confirmed. The two theories stubbornly insist on being correct and precise.

When the two theories of physics are combined, one finds a striking result. There is, in nature, a smallest measurable length [5–7]. The smallest measurable length is given by the Planck length, about 10−35 meter. The value is incredibly small, and far from any experimental detection. No experiment has reached it, and it appears that no experiment will ever do so. But above all, the value is not zero. In short, between two points in space, there is not always a third one.

Likewise, there is also a smallest measurable time interval in nature. The Planck time is about 10−43 second. Again, the smallest measurable duration is not zero. Between two instants in time there is not always a third one.

The existence of the smallest measurable length and time implies that a correct description of nature cannot be based on continuous space and time. Space and time are not continuous. There is no way to confirm that space is made of points or that time is made of instants. On the one hand, this result is fascinating. On the other hand, we cannot even think or talk about nature without continuous space and time. In particular, all equations of physics use continuous space and time.

Is there a way out of this dilemma? Yes, there is. Both points of space and point particles must be substituted by something of the Planck size. Space and particles both must have constituents of Planck size. In usual experiments, the Planck size is not detectable, the constituents can be ignored, and we can use points. When exploring general relativity only, the constituents are averaged out; thus they are not important. When exploring quantum theory only, the constituents are again averaged out; again, they are not important. But when general relativity and quantum theory are combined, and only in this situation, at the foundations of space and time, when we want to understand the origin of forces and particles, Planck-scale constituents become inevitable.

 

Black holes are the key to the constituents of nature

50 years ago, Bekenstein and Hawking discovered that black holes have entropy and temperature [8, 9]. This result implies two important aspects: first, the value of entropy (and temperature) is finite; second, the values are given by the ratio of the surface of the black hole and the square of the Planck length. The entropy value implies that black holes are made of constituents that are of Planck size in two directions (nothing can be deduced about the size in the third direction). Black holes are thus made of countable, or discrete, constituents. This confirms the results that arise when general relativity and quantum theory are combined, as expected. But in fact, an additional, new result can be deduced.

Black holes are highly condensed matter. In other words, black holes can be seen as highly condensed or compressed particles. At the same time, black holes are highly curved space. In other words, black holes can be seen as a specific configuration of space. As a result, the constituents of black holes are the same as the constituents of space and of matter. To say it simply: space and particles are made of common constituents.

How can space and particles be made of the same constituents? Space and particles differ: particles are like small stones, small and compact, whereas space is extended, spanning across the universe. The common constituents must therefore be both small and extended at the same time. The properties of black holes imply that the constituents must be tiny in two directions and large in one. In other terms, the common constituents of space and matter and radiation must be strands of Planck radius that are as extended as the universe.

 

Particles and wave functions can be illustrated

Both experiments and quantum theory show that particles are not perfectly point-like. Matter particles, specifically, are described by wave functions. Both experiments and quantum theory state that wave functions are extended and continuous. But are wave functions really continuous? No, they’re not.

When approaching the Planck scale, no experiment can prove that wave functions are continuous. In fact, all experiments suggest that wave functions are strands wiggling around extremely rapidly. As Dirac found out already in 1929, wiggling strands explain the spin ½ of matter particles [10]. He also showed that wiggling strands explain that matter particles are fermions—and the origin of the quantum of action ħ. As Battey-Pratt and Racey discovered in 1980, wiggling strands also explain relativistic wave functions and the Dirac equation [11]. In particular, blurred wiggling strands explain the existence and the spatial extension of wave functions.

If wave functions are blurred wiggling strands, what are particles? There are not many options. Particles must be tangled strands. What we usually call the position of a particle then is the region where the tangled region is located. What we usually call the probability density then is the region around the particle where the tiny strands fluctuate. What we usually call the phase of a particle is the average orientation in space of the tangle. What we usually call particle reactions are changes in the tangling of strands.

Are strands in contrast with the peculiarities of quantum theory? Are they hidden variables? A detailed exploration shows that strands do reproduce quantum theory in all its aspects [12]. Wiggling strands are not hidden variables, because single strands are not observable. Wiggling strands do reproduce quantum superpositions and quantum mechanics. Wiggling tangled strands reproduce entanglement—almost intrinsically, one could say. In short, wiggling strands reproduce quantum theory in all its details at all measurable distances.

In other terms, fluctuating strands explain wave functions. Particles are not small parts; particles are fluctuating tangles. Wave functions arise when strand fluctuations are averaged out. And indeed, classifying the various types of tangles yields the known elementary particles. Exploring the rotation of tangles explains the mass values of the elementary particles. Tangled strands thus show how to tackle the open issues of fundamental physics.

 

Space is composed of extended constituents

The combination of general relativity and quantum theory shows that space cannot be continuous. Space is made of wiggling and fluctuating strands. Empty space contains no particles. Empty space is thus made of untangled strands that randomly change shape. Can all this be correct? There is only one way to find out.

Einstein proved that space can bend and move. Space bends or curves around massive bodies, and the bending moves together with the bodies. In addition, space can bend and move also without nearby bodies: there are gravitational waves in nature. Can strands reproduce all these observations? It turns out that they do.

In the same way that strands form empty space, twisted pairs of strands reproduce the curvature of space. Large numbers of moving twisted pairs reproduce gravitational waves. Densely packed woven strands reproduce black hole horizons. Black holes cannot be compressed further because of the tiny radius of the strands they are made of. A detailed investigation shows that Einstein’s field equations of general relativity arise when the strand fluctuations are averaged out [13]. A further result is obtained: twisted strand pairs are only possible in three spatial dimensions. Strands thus explain why we live in three dimensions.

In other terms, space is made of fluctuating strands. Averaging out strand fluctuations of twisted strand pairs yields general relativity. The averaging reproduces textbook physics.

 

Fundamental physics is not based on equations

Physics describes motion with equations. More precisely, physics describes motion as a consequence of interactions—gravity, electricity, magnetism and the two nuclear interactions. Physics describes all change around us with equations of motion. But on this topic, strands imply a surprising consequence.

Because the Planck limits of space and time are inaccessible, single common constituents of space and particles are unobservable. Therefore, there is no possibility to deduce a unified equation of motion or a unified Lagrangian for single common constituents [14]. Instead, all known evolution equations—such as Einstein’s field equations or Dirac’s equation—emerge from the collective behavior of large numbers of fundamental constituents. The emergence was already found above both for the motion of wave functions and for the motion of empty space.

In other terms, quantum theory and general relativity arise from the averaging over fluctuating tangles of Planck size. All precision in nature arises from the averaging over large numbers of fluctuating tangles. Conversely, the Planck length implies the lack of a unified equation of nature.

 

Forces are due to deformations

The disappointment about the search for a unified equation is more than compensated for by a number of consequences of the strand description that are downright fascinating. In fundamental physics, there are three gauge interactions: the electromagnetic force, the weak nuclear force, and the strong nuclear force. In quantum theory, interactions are processes that change the phase of wave functions. Strands define particle phase as the orientation of the particle tangle, and the wave function as the shape fluctuations of tangles. As a result, strands imply that interactions are deformations of tangles.

Interestingly, almost a century ago, Reidemeister and other mathematicians deduced a way to classify tangle deformations. They found that deformations can be classified into three basic types, called the three Reidemeister moves. Exploring these three moves, one finds that they generate exactly the observed gauge symmetries of the electromagnetic, the weak nuclear interaction, and the strong nuclear interactions. Sets of one, two and three strands produce precisely the gauge groups U(1), SU(2) and SU(3). In short, the three Reidemeister moves on strands explain the origin of the three gauge forces in nature [15, 16].

Why did the discovery of the origin of gauge groups take so long? Strands were introduced by Dirac in 1929. It took until 1980 for Battey-Pratt and Racey to deduce wave functions from strands. In an independent field of research, it took a further decade to discover that shape changes are related to gauge symmetries. But at that time, only commutative gauge symmetries were found, because compact shapes do not yield non-commutative symmetries. Instead, the nuclear interactions are non-commutative: they require tethers. Only 15 years ago did it become clear that strands allow deducing non-commutative gauge groups.

What about the strengths of the interactions? Strands imply that the strengths of interactions are due to the statistics of shape fluctuations in particle tangles. Again mathematics—more precisely, knot theory—leads to progress. Strand tangles imply that charges are quantized, as is observed. And first rough statistical calculations yield coupling constants that agree with measurements.

In other terms, strands imply that the equations of physics do not result from one fundamental equation, but that they all result from the statistics of strand fluctuations. In simple terms, all laws—all interactions and all coupling constants—emerge from wiggling strands. For example, this holds for the mass of elementary particles. Above all, also the famous number 137.0359992… appears to be due to the average number of tangle deformations induced by the first Reidemeister move in electrically charged tangles [12]. Strands thus appear to explain the origin of all colors in nature.

 

Conclusion: renewing fundamental physics

Combining general relativity with quantum field theory—i.e., with the standard model of particle physics—implies that the unified theory of fundamental physics is based on common, discrete fundamental constituents that are both extended and of Planck radius. These strands yield both space and quantum particles. Alas, strands imply that no T-shirts with a unified evolution equation can be designed. Such an equation does not exist. On the other hand, all known evolution equations —such as Einstein’s field equations, Dirac’s equation and the gauge field equations—emerge from the collective behavior of large numbers of fluctuating strands.

Fluctuating strands explain the origin of elementary particles and interactions, and they explain all fundamental constants—elementary particle masses, their mixing angles, the coupling constants, the cosmological constant and the dimensionality of space—from their collective, statistical behavior. In particular, all colors are due to strand fluctuations. Strands and their statistical behavior thus explain all the beauty of nature around us.

Strands also make numerous experimental statements. They have been collected and published. So far, all deduced consequences and predictions agree with all observations.

In total, the use of strands instead of points as fundamental constituents of space and particles yields a new way to describe fundamental physics. Strands explain elementary particles and the four forces. Strands will explain all colors in all details. Strands will allow predicting the mass values of neutrinos.

Even though the universe is filled with fluctuating strands that connect everything, in our everyday life fluctuating strands are not noticed. Chemistry, material science, medicine, biology, geology, astronomy, engineering and most parts of physics are unaffected. In the natural sciences, the exploration of nature will continue independently of strands, and discoveries and inventions will be made independently of strands. Nevertheless, only strands explain particles, space, and all colors around us.

Strands imply that every particle in our body is connected to the rest of the universe by strands. With every movement, we move large numbers of strands. This new description of motion requires quite a change in our thinking habits—though only in fundamental physics. The exploration of nature will remain fascinating for a long time.

 

References

[1] C. M. Will, The Confrontation between General Relativity and Experiment, Living Rev. Rel. 17, 4 (2014), arXiv:1403.7377 [gr-qc].

[2]  LIGO Collaboration and Virgo Collaboration and others, Tests of general relativity with GWTC-3, Physical Review D (Particles, Fields, Gravitation and Cosmology) (2022), arXiv:2112.06861 [gr-qc].

[3]  M. Kramer et al., Strong-Field Gravity Tests with the Double Pulsar, Phys. Rev. X11, 041050(2021), arXiv:2112.06795 [astro-ph.HE].

[4]  R. Workman et al. (Particle Data Group), Review of Particle Physics, Prog. Theor. Exp. Phys. 2022, 083C01 (2022).

[5]  C. A. Mead, Possible connection between gravitation and fundamental length, Physical Review 135, B849 (1964).

[6]  L. J. Garay, Quantum gravity and minimum length, Int. J. Mod. Phys. A 10, 145 (1995), arXiv:gr- qc/9403008.

[7]  S. Hossenfelder, Minimal length scale scenarios for quantum gravity, Living Reviews in Relativity 16, 2 (2013).

[8]  R. M. Wald, The thermodynamics of black holes, Living Reviews in Relativity 4, 1 (2001).

[9]  D.N. Page, Hawking radiation and black hole thermodynamics, NewJournalofPhysics 7, 203(2005).

[10]  M. Gardner, Riddles of the Sphinx and Other Mathematical Puzzle Tales (Mathematical Association of America, 1987) p. 47.

[11]  E. P. Battey-Prattand, T. J. Racey, Geometric Model for Fundamental Particles, Int. J. Theor. Phys. 19, 437 (1980).

[12]  C. Schiller, Testing a conjecture on quantum electrodynamics, J. Geom. Phys. 178, 104551 (2022).

[13]  C. Schiller, Testing a conjecture on the origin of space, gravity and mass, Indian Journal of Physics 96, 3047 (2022).

[14]  C. Schiller, From maximum force to physics in 9 lines towards relativistic quantum gravity, Zeitschrift für Naturforschung A 78, 145 (2023), arXiv:2208.01038 [gr-qc].

[15]  C.Schiller,Testing a conjecture on the origin of the standard model, Eur.Phys.J.Plus 136, 79(2021).

[16]  C. Schiller, Testing a conjecture on quantum chromodynamics, International Journal of Geometric Methods in Modern Physics 20, 2350095 (2023).

Simple code in the mind of God

Simple code in the mind of God

Reading | Computer Science

River Kanies | 2023-06-04

Woman hand touching The metaverse universe,Digital transformation conceptual for next generation technology era.

Insofar as the activity of the mind of nature can be modeled as computation, the complexity of our physical universe is an inevitable, emergent outcome of nature’s computational potentialities, even if its innate, fundamental ‘programs’—basic ‘thoughts’ in the ‘mind of God’—are extraordinarily simple. River Kanies makes this point by leveraging Stephen Wolfram’s notion of ‘ruliad.’

The core tenet of metaphysical idealism is that only consciousness exists irreducibly. It follows that all of reality is generated within, and experienced by, a universal consciousness. We may call this universal consciousness ‘God,’ provided that we avoid anthropomorphizing it. Under idealism, because anything conceivable is possible within imagination—an inherent and spontaneous activity of consciousness—instead of asking ‘what is possible in nature?’ we must, instead, ask ‘why do things seem to follow such rigid laws?’ And ‘why does God seem to be playing dice?’ (as in microscopic quantum phenomena). The purpose of this essay is to hypothesize intuitive answers to these questions, by leveraging Stephen Wolfram’s concept of the ruliad. After all, at the end of the day, it is our intuition that we act upon, not so much our beliefs; and certainly not some sort of ‘objective’ truth, which we as mere humans cannot access directly.

Wolfram defines the ruliad as

the result of following all possible computational rules in all possible ways … it’s something very universal—a kind of ultimate limit of all abstraction and generalization. And it encapsulates not only all formal possibilities but also everything about our physical universe—and everything we experience can be thought of as sampling that part of the ruliad that corresponds to our particular way of perceiving and interpreting the universe.

As such, insofar as we can model the universe as a computational system, the ruliad is the maximum expression of its potentialities.

If we go back to the ‘beginning of the universe’ from the perspective of idealism, we start with the mind of God in the form of infinite untapped potential. We can conceive of the nascent God-mind imagining at whim, no holds barred. Then, at some point, God becomes interested in creating a (computational) structure within which realities can be constructed. There is still infinite potential, but now there are constraints at least within this ‘slice’ of the activity of the God-mind. In a sense, we could say that the idea of computation itself is a sort of structure, or a class of structures (there can be different types of computation, but they all have structure and constraints). If we accept Wolfram’s claim that the ruliad is an “inevitable” formal construct that follows directly from the existence of computation, and must therefore “necessarily exist” (as a concept in the mind of God, let’s say), then as soon as God decides to create a structure, that structure is the a nascent version of the ruliad by default: the collection of all possible computations run on all possible states.

This is still a very amorphous view of the nature of reality. However, Wolfram’s work has more to say on the subject, particularly the concepts of computational equivalence and computational irreducibility, which are related. Wolfram has been able to demonstrate that there is a class of very simple programs that generate truly complex behavior—behavior that cannot be predicted with any formula and can only be determined by running the program in full. For this reason, we can say that the patterns generated by such programs are computationally irreducible. This means that, once we have the ruliad, the next step is to trim it down to the most ‘interesting’ parts—i.e., the computationally-irreducible patterns. Wolfram also demonstrates that all such complexity-generating programs are essentially equivalent in terms of the complexity of the patterns they generate. This means that we can focus on any one of these complexity-generating programs and expect to get comparable insights as with any other.

Furthermore, Wolfram and his team were able to demonstrate that these programs are able to generate high-level patterns that correspond, in certain ways, to both quantum mechanics and general relativity. So we can hypothesize that physical reality is just a high-level construct of one such a program being run at a much lower level of reality. This provides huge insight into the nature of emergent phenomena by demonstrating that a trivial program, when run long enough, can produce patterns that seem fundamentally unrelatable to it and its initial conditions (when viewed by a human observer). If we think of the world we experience as being the pattern that such a program generates, we can understand reality in terms of layers of emergence. We start with a simple program, run it until we get the foundations of physics, and then run it until we get physical reality as we know it.

Looking at the ‘timeline’ of the ‘evolution’ of reality, we start with the infinite potential that is the mind of God. Then God creates some structure—the basis for computation—and we get the ruliad. As God ‘plays around’ with the ruliad, God focuses on the programs that generate complexity. Then let’s say that God picks one and runs the program until an entire universe is generated through layers of emergence, and so we get reality as we know it.

But, as you might suspect, there’s more to the story.

Let’s talk about the concept of an observer. To Wolfram, an observer is a persisting pattern within the ruliad that can be associated with some form of identity, but is not so cohesive as to resist influence from outside of itself. So it takes in external information in the form of representations of other local sub-patterns in the ruliad. Although the pattern as a whole is irreducibly complex, there are pockets of locally-reducible sub-patterns that can be observed and represented by a model inside of an observer within the ruliad. All patterns that we, as observers, can recognize are forms of local reducibility.

Idealism has a few things to say about this interpretation. One would be to emphasize the importance of the distinction between conscious and non-councious observers. If we consider Bernardo Kastrup’s interpretation of quantum mechanics, we understand physical reality as being manifested in relation to a conscious observer. There can be non-conscious observers in the loop, such as measurement devices, but ultimately they are just part of the relationship between the conscious observer and what is being observed.

One could claim that the existence of conscious observers within the ruliad means that consciousness naturally emerges from mere computation. To an idealist, however, this rationale is reversed: consciousness is the original ontological primitive, and computation is a subset of what is possible within the activity of consciousness. So an idealist will interpret Wolfram’s work as suggesting that there are sections in the computationally-generated patterns that represent consciousness, but not that consciousness itself emerges from that pattern. The idealist understanding is that it was the God-mind that generated the pattern in the first place.

One of the implications of Wolfram’s work from a philosophical perspective is that there is a case for faith and optimism in viewing the universe as emerging from a computationally-irreducible pattern. When observing these patterns as they evolve in layers of emergence, over time they only become more complex, not less. This suggests that nature will only continue to grow and expand. There is no suggestion, in Wolfram’s work, that the universe will, for instance, ever collapse in on itself. Furthermore, it seems that every step in the evolution of the universal pattern is necessary for the emergence of higher-level patterns. If, for example, there were some apocalyptic scenario in which all humans were wiped out, that would be a step in the creation of a yet more complex reality.

Wrapped up in this discussion are the implications of computational irreducibility and the interplay between that and optimism. If we understand all of reality, including people, to be computationally-irreducible and computationally-equivalent patterns, then it follows directly that all perspectives are valid, valuable and even necessary. As with all emergence, the higher-level patterns are dependent on all the nuances of the lower-level ones. All perspectives are valid because they are God living through us, as us, and so are inherently divinely interesting.

Perhaps the most powerful concept for bridging the gap between Wolfram’s work and idealism is what Wolfram refers to as universality. Wolfram demonstrates that any of these complexity-generating programs can be set up to simulate any other one. This means that, if we find a complexity-generating program that produces the laws of physics, that program can always be reduced—in principle—to another one, by having the latter simulate the former. No matter what pattern you are looking at, there can always be a deeper pattern from which it is being generated. Computationally speaking, no pattern is then fundamental.

For the idealist, this gives insight into the nature of the relationship between individual consciousness and the patterns generated in the ruliad. We could postulate that the job of individual consciousness is to observe the patterns of reality—identifying their regularities and constraints—and then choose a lower-level pattern to reduced them to, in a manner that conforms to the observed constraints but also enables the fulfillment of other, arbitrary constraints, as determined by the conscious being. This can be understood to be the process of manifestation: from the infinite conceivable realities potentiated within the ruliad, the job of individual consciousness is to actively make choices that steer the development of the patterns of interest.

There is a sense in which this interpretation suggests that every decision made by a conscious being adds an entire new layer to the ‘bottom of the stack’ of our emergent reality. But this should not be understood literally. Although the complex patterns we observe in nature do provide deep insight into its emergent character, they are still only a representation of reality. The claim is thus not that ‘reality is a cellular automaton,’ for example; instead, the purpose of my argument is to provide some intuition for the nature of reality at a deeper level than elementary subatomic particles. My interpretation seeks to reconcile the rigid regularity of nature’s behavior that we experience—i.e., the observed ‘laws’ of nature—with the inherently creative and generative processes of consciousness.

In summary, idealism is entirely consistent with Wolfram’s findings. Furthermore, Wolfram’s work seems to go a long way in answering some of the most difficult questions for the idealist, such as ‘if all is mind, then why do things seem to follow rigid natural laws?’ and ‘why does there seem to be an objective physical reality?’ For someone entrenched in the popular physicalist metaphysics, the goal of this argument is to show that all of physical reality can be explained within the possibility space of the ruliad, which can ultimately be embedded in a universal mind.

The subject beyond the ‘I’: On structural psychoanalysis

The subject beyond the ‘I’: On structural psychoanalysis

Reading | Psychology

Dr. Ludwig Sachs | 2023-05-28

Illustration of man walking on Penrose triangle, surreal concept

A careful investigation of our phenomenal inner life reveals a self indistinguishable from the world and others, and yet impossibly beyond both, argues Dr. Sachs. This realization constitutes a challenge to our need to self-actualize as individuals, with significant psychological (and perhaps even metaphysical) implications.

The challenge of this article will be to create an awareness of the unconscious from the perspective of structural psychoanalysis, which was developed by the French psychiatrist Jacques Lacan1 in the 20th century based on the insights of Sigmund Freud. Ideally, it should also be visually appealing, so to cater to modern vision-oriented sensitivities.

But is it not an impossible task to try to illustrate unconscious structures in any way? Yes, it is! We literally see this impossibility of visualization every day, if we just look closely. And precisely because of this paradoxical circumstance, we can almost ‘see’ the unconscious.

 

The three registers of structural psychoanalysis

Look around you; what do you see? You naturally assume that objects are out there and that you, as an ‘I,’ a human observer with a body, are somehow ‘in your head,’ at a certain distance from the objects. If you’re outside or looking out the window, you may see the blue sky. You know and have learned that you are on Earth, surrounded in all directions by the vastness of the universe, along with your fellow human beings and other living creatures. You also know, of course, that you and everything else exist in a three-dimensional space, which surrounds you and in which you can move. You can simply turn around and see the space that was just behind you, and when you turn back in the original direction, you see the same space again. Perhaps you also have a mirror nearby, in which your body, your ‘I,’ can be reflected. And while you see all of this out there, a certain amount of time passes, which runs independently of you, just as the objects around you and the space in which you move exist independently of you. With this familiar, pictorial conception of space, time and self-consciousness, we have just described the first of three registers or orders of structural psychoanalysis: the Imaginary.

Next step: the things out there will mean something to you. The whole situation is part of your life story and somehow it all relates to you; maybe it even makes sense to you. You can tell your life story using the language you have been taught. You have also been taught what the words and sounds of that language mean. So, the objects out there have names and meanings for you: table, wall, window, magazine, laptop, light, sky, tree, etc. These words, in turn, have further meanings that can be explained with words, which themselves have meanings that can be explained with further words, and so on. For you to be understood, you must follow rules when speaking. In general, everything you see out there seems to be subject to order and structure. Some of these structures may also be hidden or not fully accessible to our descriptions. And yet, what is subject to these orders and laws apparently does not have to be aware of this order itself. One could say that nature has a knowledge of the laws of nature while not being aware of them.

We, too, obey the laws of nature without having to know them. We are subject to the reference system of designations and meanings described above, without knowing its endless network. It is therefore a knowledge of which one is not aware: an unconscious knowledge. There are structures and laws in societies and interpersonal relationships, in living beings, nature, and matter, but also in language, with which we can then speak about these structures. Structural psychoanalysis calls these never-ending structures of words, designations and meanings, which in turn create words, designations and meanings, the Symbolic.

Let’s take a step further. Forget for a moment what you think you know about your situation in space and time, the structure and functioning of your eyes, your brain, and so on. Look again closely. What do you really see? What do you perceive when you simply describe what appears in your consciousness? In front of you, you see an image that looks like a space with certain things in it. But what is behind you? Can you see it without turning around? It is neither dark nor bright, it is simply not conscious to you. Only when you turn around can you observe this image and become aware of it. But what is there before that?

Now imagine that you might not actually be in a three-dimensional space that exists independently of you. After all, you don’t see space as a whole, from all sides. Stay with what you really perceive, not what you supposedly know. As a first approximation, you might imagine that you are in a dark room and your eyes are like flashlights that only illuminate a part of the darkness. What lies in the darkness before you look remains hidden from you. Perhaps the image of space only arises in the moment you look or direct your consciousness towards it. At least this idea would correspond more to what you perceive: an image in front of you that becomes blurrier towards the edges, and behind you, in the dark, a ‘nothing’ or a ‘something’ that is not conscious to you.

But is it really true that there is a front and a back? Or an inside and an outside? And am ‘I’ really in between front and back? Or inside? Let’s take a closer look.

Forget for a moment what you think you know about your body in space, optics, the structure of your eyes, and so on, and assume that you have no problems with seeing. Then the image that appears in front of you is reasonably sharp in the center, directly opposite you, and becomes less clear, less sharp and blurred towards the edge. In contrast to the edge of the image, you can focus on individual objects in the center and thus isolate them from the overall image. One could also say: your ‘I’ is in the center facing the isolated object. If you look at the image more closely, you will also notice something peculiar: the image converges exactly to this center. You can observe this particularly well if you position yourself exactly in line with an elongated object, such as a straight path or the end of a long table. No matter where you look, there is always a central point in the distance, a point towards which everything seems to converge, exactly opposite you.

But wait a minute; that’s strange: ‘I’ should actually be the center of my perception, the (central) point from which I perceive space and the image, not that point over there, in the distance! Why does the image always center there, exactly opposite my ‘I’?

Stay with your perception, not what you think you know about distances and optics. So, if the image converges at that point over there, and not here, then maybe ‘I’ am not really where I thought I was. ‘I’ am actually another—another point, not here, but over there!

But then isn’t inside also outside? Isn’t here and there actually one point, which converges with the one in the distance, in infinity? A place where ‘I’ repeatedly face different objects? And what about the ‘I’ of others? Doesn’t that also mean that we are all one ‘I,’ at the same point, in the same place, in the same hole? But that’s impossible! And with this somewhat frightening approach to the impossible, we have experienced the third register of structural psychoanalysis: the Real.

The three registers or orders (the Real, Symbolic, and Imaginary, abbreviated as RSI) are not independent of each other but, from a psychoanalytic perspective, are interconnected like the so-called Borromean Rings. These three rings are so interconnected that their links fall apart when one ring is opened. In an analogous manner, all human experiences irretrievably disintegrate into this RSI structure.

Given these frightening impossibilities and abyssal infinities, you will quickly retreat into your habitual conception of space, time and self-consciousness, that is, into the Imaginary, using your psychological defense mechanisms. But try to linger for a moment with these paradoxes and impossibilities. What does all this mean? Can it also be visualized?

The human—in psychoanalytic terminology, the subject—is apparently decentered. One might say, ‘I’ am not where I actually am, where I myself am; in a sense, ‘I’ and ‘self’ are disconnected. Let’s stay in the visual field to illustrate this again. We must focus only on what we really see and set aside our beloved knowledge for a moment. Can you see yourself when you see? No, you cannot see yourself. You may see parts of your body, your legs, arms, etc., but all of these are only objective parts of you, not yourself. Usually, one would say, “I have arms, legs, etc.,” and not “I am arms, legs, etc.” But perhaps we can see ourselves through detours? For example, you could stand in front of a mirror to see your reflection. But that is not you either! That is only your reflection, an image that you identify with and say, “That’s me!” So ‘I’ am actually someone else. And when you are standing in front of a mirror, very close to the mirror, do you see what gradually emerges there, in your pupil, in this black ‘hole’? What is looking at you from the darkness, from within you, from this hole? What appears when you want to see yourself very clearly? Your reflection! I am someone else! And in that person’s eyes, your reflection appears again, in which you are reflected again, and so on.

So ‘I’ am condemned to be someone else. Again and again, facing other objects, reflecting in them to reflect them, trying in vain to fill the hole in the same place, this impossibility; the hole in the same place where I cannot be myself; an infinite mirror cabinet without ever coming to myself.

How did we come to this inherent sense of lack? What traumatic event assailed us in our past? Is there perhaps a story about it? Yes, there is. Most people will know it. This story has been told since ancient times, whispered from generation to generation, to tell of the unimaginable, the trauma of humanity in distant times.

Once upon a time, people were completely themselves and one with what made them whole and holy. So they lived in peace with themselves and in harmony with life, and knew the word and name of everything alive. But then the unthinkable happened, a traumatic experience so terrible that humans had to repress the memory of it for all time. It was said to be a fatal mistake, a great guilt that humanity took upon itself. And so humans were banished and expelled from themselves. Since then, one wanders through space and time like Isis in search of the dismembered Osiris, hoping for redemption, driven by an unconscious knowledge to find the lost and remedy the lack.

 

The unconscious is structured like a language

Humans are characterized by an insurmountable lack. This lack of being is referred to as the splitting/dissociation of the subject and is constitutive of human existence. This means that the subject is not capable of grasping itself or being itself, but only ever being another. The split leads to the fact that the subject has no access to a part of itself. This part, which Freud called the unconscious, is however indispensable for the constitution of the subject and its psychic structure. Paradoxically, the human subject can only constitute itself as a subject through its split or dissociation.

The split or dissociation of the human subject is based on a traumatic experience that is fundamental to human existence. However, this experience has been irreversibly repressed and was referred to by Freud as Urverdrängung, the primal repression. The primal repression remains unconscious and cannot be interpreted, therefore it makes no sense.

Structural psychoanalysis regards the split of the subject to be a consequence of its submission to language. The subject is not language itself but arises through the loss or lack that language brings about. The linguistic structures that constitute the subject are called signifiers. Unlike a sign, the signifier does not simply represent something for someone, but is what represents the subject for another signifier. In short, the subject is represented by the relationship between signifiers.

Thus, structural psychoanalysis does not describe the subject as being characterized by a particular meaning, but rather by an infinite referral of meaning and significance. This referral structure is called the Symbolic and consists of an infinite, mostly unconscious structure of designations and meanings. Structural psychoanalysis calls the elements of this structure signifiers, which means that the transitions between them are discontinuous, i.e., abrupt or discrete; in the language of quantum physics: quantized, rather than continuous.

Structural psychoanalysis asserts that the signifiers influence and transform the subject’s natural needs into the so-called signified: the designated or object of meaning. It postulates the primacy of the signifier over the signified, a concept that was adopted by the French ethnologist and anthropologist Claude Lévi-Strauss. This means that the signifying element precedes the signified and that the signified cannot exist independently of the signifying element.

The transformation of needs by the signifier has an overall effect: the lack. In psychoanalytic theory, this lack is symbolized by the phallus signifier. Therefore, the subject has an interest, a desire, a longing for the signified, i.e., the object of meaning.

The chaining of signifiers creates a dynamic of lack and desire that refers back to itself without coming to a halt, without coming to itself. This iterative, self-referential search leads to emergence, self-similarity and self-organization, or fractal patterns/geometries as observed in complex systems.

This dynamic arises from the fact that the signifier is not identical with itself; instead, it is always distinguished from itself. Therefore, the signifier cannot repeat its original meaning. However, the subject desires to rediscover the original meaning, that is, ‘itself,’ since it is represented by the relationships between the signifiers. Consequently, the signifier is repeated, and only through this repetition—which must occur at least once—does it become a signifier at all. Nonetheless, the repetition is futile because the same thing is impossible due to the marking of the subject by the signifiers. Hence, the repetition is repeated again, creating an endless loop of repetitions.

Structural psychoanalysis refers to this impossibility, that which is impossible to symbolize, as the Real. The Real is the traumatic, the causeless, the unfathomable, and the logically impossible, which cannot be imagined or symbolized. The Real is also what ‘keeps coming back in the same place.’

The Real can also be found in the sciences, where in every formalized, symbolized field, such as mathematics, logic, or physics, there are areas that cannot be proven but must still serve as a basis. In these areas of knowledge, the Real functions as an axiom. The Real can also be accessible through logical contradictions that may arise during formalization. In quantum physics, the Real manifests itself as a random event. And in psychoanalysis, it appears at the points where the procedure of free association and interpretation fails.

 

The Phantasm

Imaginary objects serve as illusory compensations for the subject’s unconscious lack of being. In psychoanalytic terminology, the subject split/barred by signifiers desires an object. This object designates and represents what is missing and is therefore desired by the subject. The search for the object in physics can thus be understood as a prototypical expression of this desire.

These objects have an imaginary character, as they are oriented towards the idealized body image. Structural psychoanalysis calls the order of the body image the Imaginary. By referring to the idealized, closed and unitary image of the body, the split subject can conceive of itself as an imaginary unity, as the ‘I.’ We rely on this ideal of wholeness. It is necessary for us to mask our fragmentation/dissociation, which we could not bear.

The ‘I’ is the representation of oneself as a body or consists in the identification with the image of one’s own body. This identification process—i.e., the consciousness of the ‘I’—can be impaired, as in the case of depersonalization or out-of-body experiences. As a simple experiment, the so-called rubber hand illusion2 can be performed, in which one can experience an illusory spatial localization of the subject. This has been experimentally demonstrated by the philosopher and neuroscientist Thomas Metzinger.3

The relationship to the body image also structures our relationship to space, by orienting ourselves towards the opposition of inside and outside. This is also reflected in physical concepts, such as closed systems and internal/external observers. Freud initially described the psychological apparatus as a kind of sphere, in which the ‘I’ is structured as a modified surface of the id, shaped by the influence of the external world.

Structural psychoanalysis revised Freud’s concept of space in order to reduce the attachment to the Imaginary. Here, the spatial structure of the psyche is no longer viewed from the surface of a sphere, but is examined using topologies such as the torus, the Klein bottle, and the projective plane/sphere with a cross-cap, in order to critically question the imaginary inside/outside opposition. It is important to note that structural psychoanalysis does not view the topological structure as a mere metaphor, but instead understands the structures of the subject as the topological structure.

In topological structures that have self-referential properties, the Real becomes more apparent than in (imaginary) everyday life. The circle is the geometric paradigm for such a structure because, as we imaginarily draw a line to form a circle, we know that the line refers to itself from the beginning: the curve should close in the end, going back to where it started.

However, the circle is fraught with impossibilities that manifest themselves mathematically, as the Real, in the form of the impossibility of squaring the circle, as well as the transcendental number pi—the ratio between the circle’s perimeter and diameter—that encompasses an infinite sequence of digits, thereby being impossible to pin down with complete precision. That is why psychoanalysis does not illustrate the futile attempt to return to oneself as a circle.

Another example is prime numbers, i.e., numbers that are only divisible by one and themselves. From a psychoanalytic perspective, dividing by oneself is only possible imaginarily, because a repetition of the same is impossible.

The unity of the subject is only possible imaginarily, which is represented mathematically as the imaginary unit i: the impossible square-root of -1 in the space of complex numbers. And so the subject circulates in that space between the missing (the zero) and the unattainable unity (the whole number 1). The process unfolds around a center, namely the object of desire, which is lost and impossible to find, thus lying in the Real. In that center, prime numbers arise with every circling attempt to return to oneself. But they are impossible in terms of complete self-congruence and therefore lie in the Real, outside the number space. They can only be written imaginarily, in a continuous way of futile repetition.

Therefore, a purely symbolic, i.e., mathematical description cannot advance further here. A more comprehensive, holistic view is necessary, called a ‘biopsychosocial’ view in psychosomatic medicine.

 

The Apprentices of Sais

It remains to be hoped that “sisters of deep structures,” such as mathematics and psychology, will one day come together again and allow us to understand the mysterious cipher script that Novalis speaks of in The Apprentices of Sais:

Manifold are the paths people take. Whoever pursues and compares them will see strange figures emerge; figures that seem to belong to that great cipher script that one sees everywhere, on wings, eggshells, in clouds, snow, crystals, and rock formations, on freezing waters, inside and outside of mountains, in plants, animals, humans, in the lights of the heavens, on touched and stroked pieces of pitch and glass, in filings around a magnet, and in strange conjunctions of chance. In them, one senses the key to this miraculous script, its grammar…

 

Notes

  1. e.g. https://www.lacanonline.com/ or https://lacan-entziffern.de/ (German)
  2. e.g. Youtube https://youtu.be/xdxlT68ygt8
  3. Lenggenhager B, Tadi T, Metzinger T, Blanke O. Video ergo sum: manipulating bodily self-consciousness. Science. 2007 Aug 24; 317 (5841): 1096-9. doi: 10.1126/science.1143439. PMID: 17717189.

Spinoza versus humanism (The Return of Metaphysics)

Spinoza versus humanism (The Return of Metaphysics)

Reading | Metaphysics

Young,Confident,Super,Businessman,In,Mask,And,Cape

Now that metaphysics has been rehabilitated in our philosophical discourse, Spinoza’s thought can help us overcome our last idol of delusion: humanism, the notion that humans are in some innate sense special and separate from the rest of nature. This essay is the final installment of our The Return of Metaphysics series, developed in collaboration with the Institute of Art and Ideas (IAI). It has been first published by the IAI on 12 May 2023.

A few years ago, I took part in a conference at Bochum University, and partly out of respect for the university which hosts the illustrious Hegel-Archiev, I decided to present a paper in which I argued that, not unlike Hegel’s Encyclopedie, one can profitably read Spinoza’s Ethics as a circular text which ends just where it begins. The Q&A session after my talk was lively and joyous, but one specific question still resonates with me. “This was an interesting talk. Still, if I may. What is the point of engaging again with Spinoza’s metaphysics? Has not Kant already proved that dogmatic metaphysic is obsolete?” asked one of the participants, a senior Kant scholar and philologist. Since the question was genuinely well-meant and fair, I asked the scholar whether he would mind repeating Kant’s proof, so that we could consider it jointly. At which point I was rewarded with a warm, generous, smile followed by a hand gesture which I interpreted as indicating certain desperateness or recognition that some ‘dogmatic’ philosophers are just incorrigible.

While few contemporary philosophers would be as blunt as this colleague and friend, one must admit that the underlying view that Kant killed off metaphysics dominated modern philosophy until very recently. This perception of Kant as the “all-crushing [Alleszermalmer]” destroyer of metaphysics, is not groundless.

Today, the main resistance to Spinoza’s return—but also the reason why it is urgent and necessary—is less the dismissal of metaphysics as dogma and more an attachment to a dogma of a different sort: humanism. To the extent that Anglo-American philosophy has opened itself to rigorous metaphysical thought, Spinoza has been well and truly rehabilitated—much of contemporary monist metaphysics is directly indebted to Spinoza. Overcoming the attachment of (not just analytic philosophy) to humanism—the belief that human beings occupy a uniquely prominent place in nature—has proved even harder than rehabilitating metaphysics. But given the crises that this ideology has given birth to, overcome we must, and Spinoza is the philosopher to help us do it.

As far as the return of Spinozist thought in metaphysics is concerned, the two centuries that followed Kant witnessed two major waves of attempts to rehabilitate metaphysics by overcoming Kant’s various bifurcations of reality: first, the German Idealism of the early nineteenth century, and then, toward the end of that century, British Idealism. The two movements were closely related and Spinoza’s metaphysical monism and systematicity were a major source of inspiration for both of them. Most of the key figures in both German and British Idealism presented themselves as either Spinozists or as attempting to provide an improved, or elevated, Spinozism.

As has been argued in previous installments of The Return of Metaphysics series, analytic philosophy began as a rebellion against what was seen as the metaphysical extravagance of the British Idealists. As a consequence, throughout most of the twentieth century, the shares of Spinoza’s philosophy plunged just as the entire discipline of metaphysics suffered from severe disrepute. Analytic philosophy had a lukewarm attitude toward Spinoza in spite of the fact that, from its very inception, analytic philosophy aspired to approximate the precision and clarity of mathematics, an aspiration that could hardly be better matched in any other major philosophical work than in Spinoza’s axiomatized Ethics.

It was just natural that the recent rehabilitation of metaphysics in the Anglo-American world over the past thirty or so years has been accompanied by a huge surge of engagement with Spinoza’s philosophy, first in North America, but more recently, also in all other quarters of Anglo-American philosophy. Apart from the re-emergence of metaphysics as a central discipline of philosophy in the Anglo-American world, one can also discern a few additional causes for this recent surge, not the least of which is the emergence of Spinoza as an icon in the French philosophical scene during the 1970s and 1980s. This latter event itself took place partly due the wide impact of the work of a few outstanding philosophers—most of all, Martial Gueroult (1891-1976), but also, Alexandre Matheron (1926-2020), and Gilles Deleuze (1925-1995)—and partly due to certain developments in the French left and the attempt to establish a Marxism free from humanistic illusions. While, traditionally, Marxism was built upon Hegelian conceptual foundations, the new French Marxism of the 1960s and 1970s attempted to offer a Marxism grounded in Spinozist metaphysics, and as such, free from the illusions of teleologically structured history, subjectivity, and humanism.

The impact of the work of a few outstanding philosophers was also crucial in the rise of Spinozism on the American shores of the Atlantic, and here, the works of Don Garrett and Michael Della Rocca had the most decisive role. Both figures showed that a careful and accurate reconstruction of Spinoza’s arguments reveals a metaphysical system which might well seem odd to us (at least at first sight), yet is still highly impressive in terms of its precision, ingenuity, boldness, and consistency.

It goes without saying that the development of various forms of monism in recent analytic philosophy—think, for example, of the works of Jonathan Schaffer and Galen Strawson, each of which could be considered as offering a certain kind of Spinozism (not Spinoza’s Spinozism, but Spinozism nonetheless)—provided auxiliary winds to the sails of the new Spinozist armada.

Overcoming the dogma that metaphysics must be dogmatic—i.e., inadequately motivated, and obscure—was one of the main challenges for the rehabilitation of Spinozism. To the extent that contemporary analytic metaphysics is open to philosophy that is precise yet bold, and to the extent that it is no longer captivated by blind obedience to common sense, Spinozism can and will flourish. But there is another dogma—one which is not unique to current Anglo-American philosophy—the overcoming of which is likely to result in an even more radical upheaval.

Humanism—in a nutshell, the view that humanity occupies a uniquely prominent place in nature (if it is part of nature at all), and that the human perspective should be justly considered as constituting the boundaries and structure of reality—has deep philosophical roots, going back as far as Protagoras’ dictum: “Man is the measure of all things: of those that are—that they are, and of the things that are not—that they are not.” Humanism dominated the mainstream of modern philosophy both before and after the advent of secularization. Figures as diverse as Pico Della Mirandola, Descartes, Leibniz, Hegel, and Sartre pledged their allegiance to its tenets. In a sense, we are all humanists today: we believe in the miracle of free will (even if we do not believe in any other miracle), we adore human dignity as an innate determination of all humans (war criminals included), and we do our best to point out allegedly unique features of human beings—consciousness, self-consciousness, unity of the self, freedom, rationality, the ability to act morally, or whatever—features that may help us justify our attitude toward other animals, as mere things. Our default humanism functions as a genuine ideology: for the most part it is invisible, seamless, and taken for granted, and one needs to train herself in defamiliarization in order to recognize the arbitrariness of these deep-seated convictions.

Of all modern philosophers, Kant was the most sophisticated, most systematic, most resourceful, and most influential advocate of humanism. In the opening lines of his Anthropology from a Pragmatic Point of View, Kant writes: “The fact that the human being can have the ‘I’ in his representations raises him infinitely above all other living beings on earth. Because of this he is a person, and by virtue of the unity of consciousness through all changes that happen to him, one and the same person—i.e., through rank and dignity and entirely different being from things, such as irrational animals, with which one can do as one likes.” For Kant, one of the reasons humans “are entirely different beings” from other animals is due to the fact that humanity takes part in the noumenal realm, and, as such, is free and not constrained by the strict determinism that governs experience, or nature. Moreover, while Kant argued that we cannot have knowledge of God, he still emphatically maintained that it is rational to have faith both in a benevolent God and in the immortality of the human soul as moral postulates. In this sense, Kant expelled anthropomorphic religion from the main entrance, while inviting it back from the back door. It is thus not a surprise that for many (post-)Christians, Kant’s philosophy provided an easy soft-landing in the face of secularization, a soft-landing that allowed them to stick to notions such as evil, free-will, and even, life after death.

Humanism had its own detractors and critics from ancient Xenophanes, through (the late) Maimonides and Hume, to Nietzsche, but none was as bold and methodical as Spinoza.

Spinoza’s attack on anthropomorphic religion is well-known. His sharp criticism of most other aspects of humanism is less recognized (at least in the Anglo-American philosophical arena). Both in the Ethics and in his Political Treatise, Spinoza scolds those who uphold the common perception of humanity as a “dominion within a dominion,” i.e., as constituting an autonomous realm of beings that disturb rather than strictly follow the laws of nature by virtue of their alleged, unique endowment with free will. Whatever exclusive qualities the humanists claim bestow humanity with a unique status—elevated above the rest of nature—Spinoza would either argue that the belief that humans have this quality is a pleasant fairy-tale (as in the case of free will, a miraculously causeless event that obtains out of thin air), or he would deny that human beings are unique in having these qualities. Spinoza would equally reject the dogma that human subjectivity and reason can be better known than the world (since our knowledge of the causes of human subjectivity is still far less developed than our knowledge of the mechanics of billiard balls, and their like).

Next to Nietzsche and Luis Althusser, Hegel was one of the few readers who recognized the scale of Spinoza’s attack on humanism, but in spite of his appreciation, perhaps even attraction to Spinoza’s anti-humanism, Hegel still argued, pace Spinoza, that the “absolute must be a subject, not only a substance” (to which Spinoza would have responded by saying that if Hegel were a triangle, rather than a human being—i.e., “a subject”—he would have surely insisted “that the absolute must be triangular”).

The temptations of anthropomorphism and anthropocentrism are obvious, and it is not a coincidence that so many philosophers (and theologians) conceived the ultimate being (God, the Absolute, etc.) in their own image. Equally clear is the motivation to conceive human beings as free agents: our notions of justice, reward, and the entire package of our bourgeois values, require it desperately. But if humanity ever sets its step beyond the safe boundaries of humanism, and begin doubting this cult of our last and most magnificent idol—Man—at that very day we should recognize our debt to Spinoza, “the all-crushing destroyer” of our ultimate idol (Spinoza himself, to be clear, is not at all a misanthrope: in fact, his entire Ethics is about charting the path toward a blessed human life, though one which is free from humanistic illusions).

Would that day ever shine? Having no access to crystal balls, I can only paraphrase the words of a sage (yes, the very same sage!): since human affairs are by nature contingent and changeable, I would believe that someday, given the right conditions, this may well happen.

Why evolutionary theory contradicts materialism

Why evolutionary theory contradicts materialism

Reading | Evolutionary theory

Richard Oxenberg, PhD | 2023-04-30

Human evolution, natural selection, from monkeys to modern humans. Anthropology and genetic heritage, against the background of the starry sky,milky way

Evolutionary theory not only fails to account for the putative emergence of consciousness from a non-conscious, material substrate, but it also outright contradicts materialism by implying that subjective states have causal powers in and of themselves, argues Dr. Oxenberg. His argument is explicit, conceptually clear, original, compelling, and we could not find a way to refute it. It is an argument not against evolutionary theory, but precisely based on it. Dr. Oxenberg then goes on to conclude that “the truth of evolutionary theory is consistent with a fully informed and rational spiritual faith.”

My thesis in this paper is a fairly simple one, and one, I believe, that is fairly simple to support on rational grounds—although I imagine it will prove controversial among some dedicated to a philosophy of ‘scientific objectivism.’ But the thesis can be stated simply enough: A materialist interpretation of evolutionary theory cannot account for the subjective dimension of life and, in particular, cannot account for the desire for physical survival, which it presupposes. When we analyze evolutionary theory with care we discover, somewhat astonishingly, that it is actually inconsistent with a philosophy of metaphysical materialism. A full recognition of this undermines the claims of evolutionary materialists (such as Richard Dawkins and others) who have advanced and popularized the notion that the truth of evolutionary theory implies the falsity of religious belief.

But before considering my thesis let me first of all clarify it. I want to be quick to say that when I speak of the limits of evolutionary explanation I do not mean this in the sense in which Intelligent Design (ID) theorists say that natural selection cannot account for irreducible complexity. The ID theorists argue that irreducibly complex organic systems cannot result from natural selection, as each element of the system would need to be selected for independently. This is a more or less straightforward rational argument based upon a key premise of natural selection theory: that only what bestows a survival advantage is selected for. If, say the ID theorists, the elements of an irreducibly complex system do not each bestow a survival advantage, then the logic of natural selection itself implies that they cannot have resulted from it. On this basis, it is argued that such systems must be accounted for by something else; in particular, some deliberately acting intelligence. This is the basic argument of ID theory.

Evolutionists answer by denying that there are, in fact, ‘irreducibly’ complex systems. They maintain that the examples of irreducible complexity provided by ID theorists can indeed be shown to be reducible to elements that would have bestowed survival advantage when they first appeared. In some cases, this can be demonstrated more or less directly. In others, say the evolutionists, it is reasonable to assume it on the basis of the general success of natural selection theory. In other words, the evolutionists claim that the explanatory gaps ID theorists profess to find in natural selection theory simply do not exist.

My argument differs from this. It does not seek to demonstrate gaps in the explanatory power of evolutionary theory, but to show that evolution’s explanations do not reach as far as the evolutionary materialists (e.g., Richard Dawkins, Daniel Dennett, E. O. Wilson, Richard Lewontin, etc.) suggest. Indeed, I maintain that a careful consideration of natural selection theory itself calls into question the assumptions of metaphysical materialism. To see this, however, we will need to attend closely to the logic of natural selection theory.

That logic is straightforward and can be expressed simply: In order for a trait to be ‘selected for’ it must meet two criteria: (1) It must exist. It must be present in some entity. Nothing can be selected for that does not already exist. (2) It must be such as to bestow a survival advantage on the entity that possesses it.

Criterion (1) alone is sufficient to indicate the explanatory limits of natural selection theory. This theory is often said to be a theory of origins (of course, Darwin’s revolutionary work was entitled The Origin of the Species), but in fact natural selection as such does not explain the ‘origin’ of anything. Given that only traits that already exist can be selected for, the origin of those traits must be attributed to something other than selection. This is obvious enough and would be uncontroversial if it were not for the way that evolutionary theorists often slip into a casual, teleological mode of speaking, as if this or that trait were intentionally produced by natural selection for the purpose of fulfilling this or that survival aim. Among the most flagrant examples of this sloppy mode of speaking is the title of Richard Dawkins’ popular bestseller, The Selfish Gene. The word ‘selfish’ suggests an intentionality that is contrary to the actual logic of evolutionary theory. Of course, the evolutionists themselves understand this to be a shorthand—and scientifically sloppy—way of speaking. It is, however, more than sloppy; it is misleading. It involves the logical blunder of mistaking an effect for a cause.

According to natural selection theory, survival advantage is the effect of traits that are then selected for because they have that effect. The giraffe’s long neck, for instance, is not produced in order for the giraffe to reach the high leaves; rather, the giraffe is able to reach the high leaves because it has a long neck. What then is the origin of the long neck? Natural selection theory cannot answer this question. Evolutionary theory in general answers ‘random mutation.’ But what kind of an ‘origin’ is this? What is ‘random mutation’?

Mutation simply means change. Of course, in order for something to change, randomly or otherwise, it must first of all exist, and in two distinct ways. The medium that undergoes the mutation must exist as the substratum of mutation, and that into which that medium changes must exist in potentia as a real possibility of that substratum.

For example, in order for the gene that determines the length of the giraffe’s neck to mutate into a gene that produces a longer neck, that gene must, first of all, already exist, and, second of all, must have the potential to mutate in just the way it does. Of course, it is entirely possible, logically, that genes in general would not have that particular potential. Even more to the point, neither the existence of genes nor their potentialities have their ‘origin’ in random mutation. The term ‘random mutation’ describes a process; it tells us nothing of the origin of that which undergoes the process. It does not answer the question of how the universe comes to be a place that makes possible such genes. Carefully considered, then, neither the theory of natural selection nor the theory of random mutation provide insight into the origin of living traits and processes.

These two ideas—natural selection and random mutation—constitute the core of Darwinian evolutionary theory. Neither provide an explanation of the origin of living entities. Thus, evolutionary theory is not actually a theory of ‘origins’ at all. It describes processes, but yields no insight into the origin of that which undergoes these processes. It describes how, it does not explain wherefrom. But it is this latter that we would need to understand if we wanted to understand the nature of life in its essence.

At this point, however, we can hear the evolutionary theorist object. The question evolutionary theory answers is why organic systems have the functional organization they do. Before Darwin, this organization was thought to be the work of a supreme intelligence who crafted these systems in just these ways for a divinely ordained purpose. After Darwin, we were better able to explain this organization as a function of strictly natural, i.e., material, processes. So, the evolutionist says, Darwin’s theory does indeed explain what was unexplained or poorly explained before, and does so through appeal to strictly natural, i.e., material, processes.

True enough. But a great deal hangs on just what we are to understand by the word ‘explain.’ What is explained by Darwinian theory is the manner in which living systems become organized. What remains unexplained is the ultimate nature of these living systems. It is this nature we must understand if we are to assess the ultimate potentialities of life. And it is these potentialities we must understand in order to assess the rationality of the religious impulse. Evolutionary theory does not address this. For sure, the evolutionary materialists contend that evolutionary theory implies that the nature of life will be (or can be) understood in entirely material terms.

Were this true it would indeed undermine most religious interpretations of the meaning of life. My contention, however, is that evolutionary theory implies no such thing; indeed, it implies the opposite. We can see this most clearly by applying our analysis of the limitations of evolutionary explanation to the non-material, subjective features of life we are familiar with through introspection.

It is a commonplace of evolutionary discourse to speak of the ‘competition for,’ or ‘struggle for,’ survival. Given the intensive struggle for survival we see among living systems, we are told, only those that are well adapted to the conditions of survival can persist from generation to generation. Those less well adapted fall by the wayside. This is the logic that drives natural selection. What is seldom noted with respect to this discourse, however, is that the very use of the term ‘struggle for survival’ takes us beyond the conceptual bounds of metaphysical materialism. Material systems as such may affect one another, but they do not struggle with one another. The word ‘struggle’ implies intentionality and purpose. A ‘struggle for survival’ implies a desire to survive. We have immediate knowledge of the existence of such desire, for we experience it within ourselves. What is its origin? How has it arisen? What is its essential nature and ultimate thrust? As we’ve already noted, evolutionary theory cannot answer these questions.

But, given the sloppy way evolutionary theory is often expressed, even by those who know better, it behooves us to put a finer point on this. Surely the desire to survive is selected for, some might say. Those who have a desire to survive are much more likely to behave in ways that will further their survival than those who do not, hence the desire to survive confers a clear survival advantage.

This is true enough (although its implications are not at all what the materialists might suppose, as we shall see in a moment). But, again, to suppose that this explains the desire to survive is to mistake an effect for a cause. The survival advantage conferred by the desire to survive is an effect of that desire, not its cause. The selection that then takes place is the effect of the survival advantage. The cause of the desire is not at all given by natural selection theory. It is not even addressed by it. Nor, for reasons stated above, does the idea of ‘random mutation’ provide a causal explanation. The origin of this desire is a pure mystery.

More broadly, the origin of the subjective in general is a pure mystery. No insight into this is provided by evolutionary theory. No one has, or can, explain the mechanism through which a material gene, however randomly mutated, results in thinking, feeling, caring, hoping, loving, etc. Neither the origin nor ultimate nature of these subjective states are given by evolutionary theory. True enough, evolutionary theory tells us that if subjective states are traceable to our genetic structure, then those states will be subject to selective pressures. But this tells us nothing at all of their origin or ultimate nature. Nor does it shed any light on how the subjective can emerge from the material.

As some will note, this is simply the mind-body problem considered in the context of evolutionary theory. My point in bringing it up in this context is simply to indicate that evolutionary theory in no way resolves it. This again highlights the limits of evolutionary explanation, and in a particularly striking way. After all, what we are subjectively is what we are most intimately and immediately. Our subjectivity is our immediate presence to ourselves. Without it, we would not be what we are. If we wish to understand our origin and ultimate nature, then, it is the origin of this subjectivity we must understand. Evolutionary theory provides no insight into this.

But we can take this point a step further. Indeed, we can turn the materialist argument altogether on its head. How, we might ask, can the desire to survive—a subjective state—confer a physical survival advantage? It can do so only by affecting physical behavior. The natural selection argument in this instance goes something like this: animals who desire survival are, for that reason, more likely to act in ways that further survival than those who do not. Hence, they are more likely to be winners in the struggle for survival and the desire for survival is ‘selected for.’ But this will only be true to the extent that desire—again, a subjective state—can have physical efficacy. Somehow this mind-state must be able to reach into matter, the matter of an animal’s body, and affect its behavior. To the extent that we were to understand evolutionary theory as grounded in, or as implying, metaphysical materialism, this would make no sense. How can a mere subjective state physically move a material body? On materialist grounds it cannot. This leads us to a startling conclusion: In order for the desire to survive to have any bearing on natural selection at all, metaphysical materialism must be wrong.

We can flesh this out by examining what we are calling the ‘desire for survival’ in closer detail. It might be pointed out that the phrase ‘desire for survival’ is an abstraction. In fact, we don’t desire ‘survival’ per se, but the furtherance of pleasurable states and the avoidance of painful states. It just happens that natural selection associates pleasure with survival and pain with threats to survival. If the opposite were the case we would desire our demise. Fair enough. But how are we to understand how such an association occurs? In order for the subjective states of pleasure and pain to become associated by natural selection with survival and threats to survival, those states must have the power to induce survival-relevant behavior. After all, it is behavior that is either selected for or selected out by tending to promote or undermine survival. In order for an association of pleasure with survival-promoting behavior to be selected for, the subjective state of pleasure itself must have the power to modify behavior. If subjective states couldn’t affect behavior, they would have no bearing on selection. But how, on materialist grounds, can subjective states modify physical behavior?

What we would expect, on materialist assumptions, is that there would be no subjective states at all. Of course, we know this not to be the case. In the face of the undeniable fact of subjective states, then, materialism maintains that these states are mere ‘epiphenomena,’ i.e., irrelevant to the causal nexus of the material world. But this is just what natural selection shows is not the case. If subjective states were irrelevant to material causality, they would have no correlation with survival-relevant behavior. We would expect a random distribution of these states, and expect them to have no affect on, or association with, behavior.

But we do not find a random distribution of subjective states; rather, we find that positive (i.e., desirable) states are, by and large, associated with survival-promoting behavior and negative states with survival-threatening behavior. Again, this could only result from natural selection if these states can reach into the physical world and change it. But if they can, then this means the subjective—i.e., the non-material—must have physical efficacy. This altogether contravenes the assumptions of materialism.

The materialist would likely counter that these states are not merely subjective, but have material counterparts in the brain or nervous system. Neurological study leaves little doubt that this is so. On this basis, the materialist would argue that it is not the subjective quality of the state that affects physical behavior but its material, neurological counterpart. This would allow all causality to remain on the material plane. Fair enough. But then how to account for the fact that these subjective states have just these subjective qualities? Why, for instance, does the material counterpart of pleasure feels good subjectively? Why does pain, or the physical counterpart of pain, hurt? The materialist can give no plausible explanation for this. If everything results from strictly material interactions to which the quality of subjective states is irrelevant, then the subjective qualities of pleasure and pain would have no efficacy. Not only would there be no need for these subjective qualities to be as they are, but, even more to the point, we could find no plausible explanation for how these qualities come to be associated with physical states and physical behavior relevant to survival. Can it be a mere coincidence that sticking a hand in a fire hurts, or that sexual activity feels good? Such a suggestion strains credulity.

The natural selection argument must be that the subjective quality of hurt is selected for because this subjective quality itself induces beneficial avoidance behavior. But, again, this means that the hurt itself—that is, the subjective quality of hurt—must be able to somehow reach into the physical and modify it. It must somehow be able to cause (or induce) our bodies to behave in certain ways. Only to the extent that this is the case does it make sense to believe that natural selection favors an association of hurt with avoidance behavior.

But if this is so then the materialist thesis that reality—or, at least, all that is efficacious in reality—has a material basis, must be wrong. We arrive at the astonishing conclusion that evolutionary theory, far from favoring a materialist metaphysic, actually undermines it, by indicating that the subjective has causal efficacy in the material world. This suggests that the subjective—consciousness—cannot be a mere byproduct or ‘epiphnomenon’ of matter.

What, then, is it? Where does it come from? What is its essential nature? Evolutionary theory does not—indeed cannot—answer these questions.

Of particular relevance to religion is the question of what potentialities the subjective has. Can it exist only in association with the material world as we know it, or might there be subjective domains of reality that do not have or require a material basis or association? Nothing in evolutionary theory yields an answer to this question. Are the values that arise from subjective desire—of which the value we place on physical survival is one—restricted to values that promote physical survival, or might the valuation of physical survival be but one aspect of a broader teleological trajectory, inherent to the subjective as such, that would find its ultimate consummation in what the religious person might call ‘communion with God’?

Again, given that evolutionary theory tells us nothing about the origin of the subjective, and hence nothing of its ultimate nature or thrust, it provides no answer to this question. In order for living systems to survive over generations they must engage in behavior that will effectively pass their genes along, and—to the extent that our subjective valuations have physical efficacy—this implies that behaviors that do will tend to be valued. But this in no way suggests that these are the only values inherent to the subjective.

The materialist might point out that we have evidence for Darwinian evolution and the ‘struggle for survival’ it suggests, but no evidence (other than the subjective testimonies of the faithful) for a religious teleology. But such a point would only be relevant to the extent that belief in the former contradicted belief in the latter, such that we had to choose between them. The point of my argument is to indicate that it does not. Does evolution raise difficulties for theology insofar as it highlights the dog-eat-dog (or animal-eat-animal) structure of the natural world, a structure, some might think, inconsistent with belief in a loving God? Perhaps, but then we did not need Darwin to point out this aspect of nature to us. Isaiah was aware long ago that, in the world we occupy, the wolf does not lie down with the lamb (except to eat it!). What in theology is called the ‘problem of evil,’ both natural and moral, has long been a difficulty for religion. Faith demands the ability to believe that this problem is not finally fatal to the rationality of theism. This is certainly a subject worthy of consideration, but it is a subject we will have to leave for another time.

In closing, I wish to make clear, again, that what I have presented here is no ‘God-of-the-gaps’ argument. What is at issue is whether evolutionary theory makes it unreasonable to believe that human life, at its core, is characterized by a teleology that finds its true consummation in some version of ‘communion with God.’ This is suggested by evolutionary materialists such as Dawkins and Dennett. My argument is not that they are mistaken because evolutionary theory is wrong or incomplete (as the ID theorists maintain), but that they are mistaken because evolutionary theory does not have the implications they suppose. In particular, it does not preclude the possibility of a ‘spiritual’ teleology. Indeed, when we consider evolution’s own recognition that life ‘struggles’ for survival—a struggle manifest in the subjective desire for survival—we see that it actually presupposes a teleological dimension to life.

My claim is that there is nothing in evolutionary theory that would require us to believe that life’s teleology is fully expressed in the pursuit of terrestrial survival. What evolution highlights is that terrestrial life includes this pursuit (something, of course, we already knew), but it neither proves nor implies that it is wholly defined by it.

Let us suppose, then, as a thought experiment, that some form of theism is right. What would we expect to see? We would expect to see the emergence of some creature whose desires are indeed favorable to terrestrial survival (otherwise the creature would not survive) but not wholly satisfied with such survival—a creature straining for something beyond mere material, terrestrial goods, in ways that yield no obvious selective advantage. I maintain that this is just what we do see in the human struggle for beauty, truth, justice and the spiritual in general. In this way religion serves as some evidence for its own validity. The human spirit aspires to more than mere material continuance. The suggestion on the part of evolutionary materialists that it does not—that, indeed, it cannot—lies at the heart of the hostility many religious people feel toward evolutionary theory. But when we separate evolutionary theory from the philosophy of metaphysical materialism it is often used, or misused, to advance, we see that there is no need for such hostility. The truth of evolutionary theory is consistent with a fully informed and rational spiritual faith.

Fitting theism in the natural frame

Fitting theism in the natural frame

Reading | Theology

The courtyard and stream at the Salk Institute, San Diego at sunset. With its ritualistic, brutalist design. Designed by the famous Kahn.

The ‘hard problem of consciousness’ demands an expansion of the naturalist understanding of reality that may allow for some form of reconciliation between science and theology, argues Rev. Dr. Joshua Farris.

I’ve recently attended a summer study and conference on the topic of “Naturalism and the Religious Worldview” in Croatia, put on by the Humane Philosophy Project. As you might imagine if you are familiar with debates in science and religion, these are often framed as incompatible partners. But not only is there ongoing interest in exploring the relationship between the two, in fact there is a recent movement afoot that attempts to take a step further by reconciling the two perceptibly incompatible worldviews. The picture that comes to mind is of my wife’s Boxer dog trying to fit in the doggie door (intended for small dogs); it just doesn’t seem to work. Let’s consider the development of Naturalism and Religion.

 

Theology

Theology you might take to be the study of God’s nature and, especially, his acts in creation. In a classical understanding, the study of God boils down to the study of primary and secondary nature. All things are related to God as the primary referent of study, and his creation is a study of that which is a secondary act, which only analogously tells us something about the divine nature (his characteristics; or in an Eastern understanding, his ‘energies’). But this is strikingly distinct from the purported object of study as conceived in Naturalism.

Granted, anyone familiar with the study of Naturalism—and my students will attest to this fact—knows that in contemporary studies it—Naturalism—is notoriously hard to define. And it is not hard in the way that all philosophical concepts are difficult to define, but it is uniquely so because the boundaries have been so expanded that the term has become rather porous.

 

A survey of naturalism

In the early twentieth century, American analytic philosophers defined it as a thesis committed to the natural world of biological organisms, particularly Darwinian evolution, often with a commitment to the exclusion of any ‘spooky’ entities like God, spirits, angels, and things like ghosts. Ancillarily to this, there is often a commitment to what is called the “scientific method”—perceived as it is to be the rigorous process of empirical testing of phenomena. One might describe the view as falling under three broad types. The first is Metaphysical Naturalism, the second Epistemological Naturalism, and the third Methodological Naturalism—i.e., the scientific method.

In short, Metaphysical Naturalism is “the view that everything is or is grounded in the kind of entities studied by fundamental physics.” Epistemological Naturalism is “the investigative practice of studying the natural world by prizing the a-posteriori approach”[1] [Editor’s note: a-posteriori knowledge is knowledge that can only be gathered through empirical experience, and not logically deduced a priori.]. Methodological Naturalism is the study of the objects of physics and biological organisms.

Metaphysical Naturalism is standardly defined as affirming that: only nature exists; nature has always existed or simply popped into existence; nature is deterministic at one level; nature is purely physical (the natural entailment from its physical basis in particles); and finally, nature is a self-explanatory system. Correspondingly, Evolutionary Naturalism is defined as that which is physical and explained by physical causes and effects through adaptation, genetic drift, and other evolutionary mechanisms. Theism, in this definition, is immediately ruled out as misfit. It just doesn’t fit, like the boxer trying to fit in a small doggy door.

But then, there is Epistemic Naturalism, a practice guided by a set of methodological principles that takes it that the surest way of knowing is funneling everything through the rigorous empirical method (Francis Bacon often comes to mind as the earliest proponent of such a view). In such a view, not only is the a-posteriori method prized, but the a-priori method is suspect.

One way of defining Naturalism, generally, comes from the famous philosopher Thomas Nagel. He defines it along similar lines as Evolutionary Naturalism, given above, particularly highlighting the non-teleological nature of Nature. He states:

The profoundly nonteleological character of this modern form of naturalism is concealed by the functional explanations that fill evolutionary accounts of the characteristics of living organisms. But any reference to the function or survival value of an organ or other feature is shorthand for a long story of purposeless mutations followed, because of environmental contingencies, by differential reproductive fitness—survival of offspring or other relatives with the same genetic material. It is in the most straightforward sense false that we have eyes in order to see and a heart to pump the blood. Darwinian natural selection could be compatible with teleology if the existence of DNA had the purpose of permitting successive generations of organisms to adapt through natural selection to changes in the environment—but that, of course, is not the naturalistic conception. That conception, far from offering us a sense of who we are, dissolves any sense of purpose or true nature that we may have begun with. The meaning of organic life vanishes in the meaninglessness of physics, of which it is one peculiar consequence. It is widely thought that, without knowing the details, we now have every reason to believe that life arose from a lifeless universe, in virtue of the basic laws of particle physics or string theory or something of the kind, which did not have life or us “in mind.” If this is the case, then it is difficult, nigh impossible, to see how God could fit into this picture.[2]

However, not so fast! Some naturalists have expanded its boundaries so much that it fails to fit so neatly into the definitions given already. Lynne Rudder Baker describes an expansive, non-reductionist Naturalism she calls Liberal Naturalism, which loosely impacts epistemic access: “I suggest that liberal naturalists not locate their naturalism in entities lacking causal powers but rather espouse naturalism only in the sense of ‘pertaining to the natural realm.’ In that way, they could be more liberal than nonreductivists like Kornblith and still be naturalistic.”[3] The question with which Baker leaves the reader in her fascinating book, Naturalism and the First-Person Perspective, is whether Naturalism is disenchanted or optimistic.

There are myriad sets of interrelated problems with Naturalism, generally, like the problem of definition. But there is a more fundamental problem with it: namely, what some have called the intuitive problem of folk psychology.

 

The intuitive problem with the naturalist attitude

The common problem with Naturalism, or the naturalistic attitude as I call it, is that it fails to account for that which is most obvious to us in our experience of the world. This is what some will call the intentional attitude, consciousness, or simply experience of what it is like to be in relation to the world. Instead, it either eliminates the intentional attitude or it eschews it to such an extent that it becomes irrelevant to knowledge and no longer holds the primacy of place in how we come to know the world.

The intuitive problem is, simply put, a problem of reconciling the natural or scientific attitude giving primacy of place to the empirical method with that which is most apparent to consciousness. Consciousness is a term for what it is to be aware of. It is inclusive of properties of what it is like to experience in general. In other words, it is the difference made between that which constitutes the green grass in my yard (e.g., the properties of particle physics and ecology that come to comprise the grass and all that include it; namely, the dirt, the roots, and the blades themselves) and the experience of that grass, often described by what it is like to see the shape, the color, and all that surrounds it. Thomas Nagel calls this “subjective appearances.” The reality of our subjective appearances is so undeniable that it takes primacy of place over the constituent parts that undergird the physical reality studied by science.

David Chalmers deems these “subjective appearances” irreducible to physics or biology, a problem he refers to as the “hard problem of consciousness.”[4] In this way, consciousness is a phenomenon that biology and physics are unable to capture. We need more than physics and biology to capture it.

Thankfully, we are beginning to see this turn in science where consciousness (i.e., the intuitive move or the folk turn) is taking up space, once again, in the literature. Ironically, the naturalist attitude is re-presenting itself, but whether it can make sense of the “intentional attitude” is again up for debate in a fresh new way.

 

The theological turn in science

The theological turn in science is characterized by this move to reconsider not only the “intentional stance” often characterized by consciousness, but more than that, the openness to theology itself as a paradigm shift toward God as having, once again, an explanatory role in our understanding of the world. Some in the theological turn are attempting to re-envision the naturalist attitude in a way that gives credence to the natural world, the study of the objects of physics, and the empirical (i.e., a-posteriori) method as primary. They are now expanding the boundaries of Naturalism to include God and consciousness. Are they successful? That is yet to be seen. But I have serious reservations, since this privileges the following:

  1. The scientific method (i.e., Methodological Naturalism).
  2. Natural law.
  3. Biological generation as an originative explanation for all organisms.

This also means not privileging the mind, intentionality, the first-person perspective, or consciousness.

In fact, some of the main proponents of this new endeavor are clear that the common privileging of the mind, consciousness, the first-person perspective, etc., in philosophy and theology needs re-envisioning because this characteristic is often what has created insuperable problems for progressing the discussion between science and religion. These proponents include most notably Sarah Lane Ritchie, Christopher Knight, and others. They believe that what underlies the challenges for integrative science and religion projects follows from the dubious ‘intuition’ that the mind and brain are dualistic entities. In other words, the supposed hard problem of consciousness stands underneath the tension between science and religion as reflected so clearly in the recent accounts to find God acting at the joints between consciousness and the natural world (i.e., causal joint theories). By bringing God and nature closer together, they suppose we can construct newer theories that avoid God’s messing about ‘unnaturally’ in nature’s regular law-like way. These rely on the unnecessary assumption of some sort of non-physicalism regarding God, the mind, and action.

Without offering a full-blown critique of these projects, it is worth advancing some concerns. First, the challenge of intuitions calls into question the veridicality of intuition itself. Ironically, the affirmation that the scientific method (i.e., Methodological Naturalism) should be privileged in some way concerning knowledge depends on intuition. Second, by not privileging the mind, consciousness, and the first-person perspective, the leaders of Theological Naturalism not only import intuitive ideas of God, humans, and actions that, on the surface, appear inconsistent with their projects, but these rather important terms become insufficient designations for which we have little to say. Third, with all the problems customarily associated with Naturalism, as suggested above, it’s hard to conceive of theology being fitted to a Naturalism frame. Once again, my wife’s boxer trying to fit into a small doggy door just will not work. But is that what we are left with? I don’t think so, and I certainly hope not.

 

A better way

Surely, there is a better way than fitting a square peg in a circular opening. Science construed along naturalist lines, according to the naturalist attitude, obscures that which is central to the world. Minds are central and deserve not only a place at the scientific table but must be brought into our engagements with scientific endeavors. This requires, however, an expansion not of the boundaries of nature so as to capture mind, agency, and God, but to understand nature in light of the mind, agency, and God.

Some in recent science and religion literature have argued that, at best, Methodological Naturalism could be operational with a small set of data studied in a lab (what you might think of as operational science), but even the findings here would be quite limited in its explanatory value.

An alternative approach advanced by Benedikt Paul Gocke is that we should expand our understanding of science in such a way that both philosophy and theology fall under a broader definition of science, which takes into account all aspects of proper rational principles and values in the process of systematically making sense of the world. This, of course, affirms that minds are central and the often-disregarded a-priori method must be re-employed as we consider the natural world because, after all, mind is fundamental to nature.[5]

 

For a new project pursuing this sort of reflection, see here: https://designtheology.org/conference-registration

 

Notes

[1] These working descriptions are common in the literature. For further explanation, see the helpful article: Papineau, David, “Naturalism”, The Stanford Encyclopedia of Philosophy (Summer 2021 Edition), Edward N. Zalta (ed.), https://plato.stanford.edu/archives/sum2021/entries/naturalism/.

[2] Thomas Nagel, Secular Philosophy and the Religious Temperament (Oxford: Oxford University Press, 2010), 15, 16. Also see: Papineau, David, “Naturalism”, The Stanford Encyclopedia of Philosophy (Summer 2021 Edition), Edward N. Zalta (ed.), https://plato.stanford.edu/archives/sum2021/entries/naturalism/. [accessed on August 10, 2022].

[3] Lynne Rudder Baker, Naturalism and the First-Person Perspective (Oxford: Oxford Uinversity Press, 2013), 17. She generally agrees with De Caro and Voltolini’s ( 1 ) the claim that no entities can violate the laws of nature, and (2) the clam that “ontology should be shaped by natural science alone.” To not allow entities that violate the laws of nature is to not allow entities with causal powers without coming from the natural science. See also De Caro, Mario and Alberto Voltolini, “Is Liberal Naturalism Possible?” in Naturalism and Normativity ed. by Mario De Caro and David Macarthur (New York: Columbia University Press, 2010), 76. Of course, also see David Papineau in his Stanford Article on “Naturalism”. Also see the following select key figures: Pereboom, Quine, Steven Pinker, Galen Strawson. This is a long and fascinating discussion here. In more recent mainstream literature Ray Kurzweil and Novel Harrari develop optimistic accounts through the lens of transhumanism. Meghan O’Gieblyn develops a somewhat agnostic view with a disenchanted outlook that pervades most of the writing in God, Human, Animal, Machine (New York: Doubleday, 2021).

[4] David Chalmers, ‘Facing up to the Problem of Consciousness’ in J. Shear, ed., Explaining Consciousness: The Hard Problem (Cambridge: MIT Press, 1997), 18-19. Josh Weisberg, “The Hard Problem of Consciousness,” Internet Encyclopedia of Philosophy. https://iep.utm.edu/hard-problem-of-conciousness/ [accessed on Friday October 7, 2022].

[5] Thanks to the Humane Philosophy Project for inspiring this short article and to the Alexander Von Humboldt Foundation and Göcke, the Chair of Philosophy of Science and Philosophy of Religion for the time and space to work on this.