Book Read Free

Time Loops

Page 20

by Eric Wargo


  By convention, most graphs put time along the horizontal x-axis, and intuitively in our culture we often think of time as “running” from left to right. So to visualize such a system, mentally rotate your simple optical eye away from the vertical y-axis of space so it is directed instead horizontally, facing to the right, along the x-axis. Instead of an in-falling rain of light being constrained by a narrow aperture to form a coherent image on a recessed surface, the noisy back-flow of influence from the future needs to be constrained, or post-selected, later in time, on the right, to form a coherent “image” at an earlier time point, on the left, when it is (er, was —we have to fudge our tenses here) measured initially, perhaps via some form of weak measurement. What would be the post-selection parameter acting as the temporal “aperture” in this time eye? Most simply, it could be existence at that later time point: survival , in other words … if not the survival of the organism as a whole, at least the persistence of the molecular apparatus doing the measurement.

  Weak measurement is a sophisticated tool used in experimental situations to (among other things) detect possible retrocausal influences. But with billions of years of trial and error in the primordial soup, it may not have been necessary to engage in something as subtle as that. All a quantum presponsive circuit, a molecular future detector, needs to have done is learn to somewhat reliably detect a difference between groups of identical quantum-entangled objects (particles, atoms, even molecules) whose only difference is what happens to them next (e.g., in a few milliseconds). An array of entangled qubits in a molecular quantum computer, representing multiple options in a decision space (such as moving to the left versus moving to the right), could serve as a precognitive guidance system, orienting the organism generally toward positive outcomes ahead in its timeline. 21

  My hand-waving makes it sounds nice and easy—but do we actually know of any molecular structures or systems in nature that could be capable of detecting optimal outcomes (most basically, survival) in their future? Maybe . It just so happens that the ongoing search for the roots of consciousness in quantum biology has turned up an excellent candidate—if not for consciousness, then at least for cellular quantum computation, and with it perhaps the temporal shenanigans that would make a time eye possible.

  The cytoskeleton or internal structure of all complex cells is formed from tiny tubular polymers called microtubules . Originally thought to be merely the bones of cells, giving them their shape, these highly dynamic structures are now known to drive cellular movement and shape-changing and to control cell division. They seem to act as the “brains” of cells. Thus, their information storage and computation abilities have attracted a great deal of attention in recent years. 22 Interest in these structures as computing devices is often associated with the work of Stuart Hameroff. As an anesthesiologist at the University of Arizona in the 1980s, Hameroff noticed that anesthetics seem to cause unconsciousness via their actions in microtubules, which are particularly numerous and complexly arrayed in neurons. Thus, he hypothesized that these structures may be the quantum computational basis for consciousness. 23

  A quantum computer, remember, is a matrix or lattice of entangled atoms or other particles that are kept isolated from their surroundings to preserve their entanglement and can act as quantum bits or qubits. Whatever you do to one of the particles affects all the others simultaneously—although again, entanglement may really reflect a zig-zagging connection across time. Conveniently, microtubules are perfect tubular lattices built from individual cup-shaped proteins called tubulin. Proteins like tubulin have multiple ways they can fold, called conformational states . After several researchers had suggested that quantum mechanics may play a role in determining how proteins select one versus another state, Hameroff suggested that tubulin molecules could play the role of qubits through their variable conformational states. 24 The lattice structure of microtubules places the individual tubulin molecules, and electrons in bonds within them, at distances that would enable entanglement to occur. Hameroff and his colleagues subsequently confirmed that microtubules transmit electricity according to quantum principles: Like semiconductors, they offer no resistance. This makes it increasingly promising that his hypothesis—at least about their quantum computational abilities, if not their role in consciousness—could be right. 25

  The possible role of microtubules as the central information processors of complex cells also interested the pathbreaking evolutionary biologist Lynn Margulis. 26 According to Margulis, complex nucleated cells (eukaryotes ) formed originally over two billion years ago from the endosymbiosis or merger of bacteria having different lifestyles and able to make different contributions to the collective. For instance, the engulfment of specialized bacteria gave eukaryotes their oxygen-burning mitochondria and the chloroplasts that enable plant cells to photosynthesize. Both of these structures still retain their own DNA separate from that found in the cell’s nucleus, proving they were once independent-living organisms. 27 Establishing the origins of microtubules has been more difficult, but Margulis argued that those structures, as well as the cilia and flagella that facilitate motion in many eukaryotic cells, are the inheritance of an early engulfment of the distant ancestors of today’s spirochetes (a group that includes the pathogens that cause modern syphilis and Lyme disease). These bacteria were little corkscrews that distinguished themselves from other early bacteria by speedily moving from place to place. The undulatory movement of spirochetes, as well as the motors that drive the movement of cilia and flagella in more complex cells, are made possible by the dynamism of microtubules.

  The link between microtubules and movement could be consistent with the hypothesis that microtubules were the first cellular guidance systems, even the first future detectors. You do not need to “decide” much if you are a relatively stationary bacterium floating in muck or clinging to some surface. And most types of bacteria do not have microtubules. (Today, bacteria are thought to navigate mainly by orienting toward or away from chemicals in their environment, called chemotaxis .) But if you are a speedy mover, a way of making informed choices about whether to move to the right or move to the left (or up or down) could come in very handy. If microtubules are quantum computers presponsive to post-selected outcomes (i.e., survival) in addition to perhaps encoding a record of past successes, then there you have it: Engulfed spirochetes may have endowed eukaryotes not only with motility but also with quantum pre-sense, and perhaps simple learning ability, via their microtubules. If such a molecular quantum computer could detect the relative favorability of multiple decisional options, even a few milliseconds into the future, it would obviously have conferred a valuable selective advantage on any cell equipped with it. It would have given that cell the ability to bind time .

  Time binding was a term originally coined by the philosopher Alfred Korzybski to denote our species’ unique ability to transmit information to later generations and thereby pursue goals that transcend the span of an individual’s lifetime. For Korzybski, who influenced many 20th -century science fiction writers like Philip K. Dick, Robert A. Heinlein, and Frank Herbert, time binding was implicitly higher than both space binding , the activity of animals who live in an eternal present and are dominated by the imperative to forage and hunt for food in their environment, and energy binding , the activity of plants that, via their chloroplasts, convert energy from the sun. 28 Korzybski wrote before the era of quantum biology, and he was not thinking in terms of any ability to use information from an organism’s future as well as its past. But something like time binding, mediated perhaps by a molecular future detector along the lines I have proposed, may be as old as those energy- and space-binding functions of life, truly a “first sight.” Earth’s primordial soup may have been a precognitive soup.

  The Big (6-Inch-in-Diameter) Picture

  Starting right around the time Charles Howard Hinton wrote his A New Era of Thought , the Spanish pathologist Santiago Ramón y Cajal was using paint and ink to depict the animal neuro
ns he saw under his microscope with great detail, revealing that these cells were like trees with often hundreds of branches and countless bud-like projections, each making a connection to another neuron. At that time, it was not yet widely believed that neurons were the building blocks of thought, but their importance rapidly became apparent to Ramón y Cajal. Around the turn of the century, he drew and painted even more mindbogglingly complex human neurons, vast and sublimely intricate despite being so tiny. He called neurons “mysterious butterflies of the soul whose beating of wings may one day reveal to us the secrets of the mind.” 29 Ramón y Cajal’s work paved the way for 20th and 21st century neuroscience.

  With recent visualization technologies, even a one cubic millimeter spec of mouse cortex looks like a dense Amazonian rainforest—a vast jungle of trees with roots and tendrils making millions of connections. The human brain, an object a little bigger than a grapefruit, contains about 86 billion of these cells, each one making about 1,000 synaptic connections with neighboring cells, amounting to about 80 trillion connections across the brain. It is sometimes said that there are more possible paths that a signal can take through this structure than there are atoms in the entire universe. Neuroscientist Christof Koch famously declared that the brain is “the most complex object in the known universe.” 30 This is true even just at the “macro” scale neuroscientists can easily study with present-day imaging technology, and trends in several research fields promise to exponentially increase our knowledge of the brain’s complexity in coming decades.

  The most publicized controversy in neuroscience and philosophy today concerns the brain’s role in relation to consciousness—whether experience and awareness arise solely from brain processes, whether consciousness is an “emergent property” that rests on those processes yet cannot be predicted by them, or, again, whether it may be somehow more basic and universal in nature. Some, like Koch, suggest that consciousness is intrinsic to complexity itself, and that while the brain may be super-conscious, even a rock may possess a tiny bit of that ineffable quality. Critics of Enlightenment materialism (including many parapsychologists) are particularly keen to reject neuroscience assumptions that consciousness is a product only of brain processes or even that it is some kind of higher-order emergent property. 31 (Again, Rupert Sheldrake has called this assumption “promissory materialism.” 32 ) Alternative theories have long been proposed, such as that the brain merely acts as a kind of prism or radio receiver for consciousness—a view argued by Frederic Myers, Henri Bergson, and the psychologist William James, for example. 33 (Although today’s scientific psychologists often consider James the father of their science, they choose to overlook his interest in psychical phenomena and his anti-materialistic views on consciousness.) The novelist Aldous Huxley later used the metaphor of the brain as a “reducing valve” for consciousness or what he called “Mind at Large.” 34 The position that consciousness is actually fundamental and irreducible in nature is not unrelated to this idea. Panpsychism, for instance, is the position that matter is just a manifestation of mind—an idea with deep roots in Eastern philosophy and advanced in different ways by prominent 20th -century thinkers like Bergson, philosopher Alfred North Whitehead, and Carl Jung. 35 A good argument can also be made that consciousness is really an ill-defined term that marks the breakdown of language and symbolization at certain boundaries and margins of knowledge, rather than a well-formed problem that either the sciences or philosophy could ever hope to solve definitively on their own. 36

  Whatever the case—whether the material brain produces consciousness or merely receives or filters it—there is much less debate over the importance of the brain in shaping experience, controlling the body, and even in encoding an individual’s personality and memory, all the forms of information that we find meaningful in defining our selves and in helping us act successfully in the world. Damage to specific brain areas produces very predictable and often catastrophic deficits in functioning and impairments in meaning-making, as Oliver Sacks showed in his prolific writings. The brain as described by contemporary neuroscience is a jaw-droppingly complex mechanism of interacting parts and functions—a vast, hyperfast system coordinating sensory inputs, motor responses, and involuntary processes throughout the body from heartbeat to breathing to digestion as well as preserving or “storing” a record of experiences (although the old computer metaphor for the brain, with its storage and retrieval operations, has largely given way to new metaphors drawn from weather and other chaotic systems). As discussed previously, it is also an imaging system with unbelievable resolution, able to generate realistic pictures and sounds and words in the inner workbench of thought—images based on real-life experiences as well as ones that are wholly new and original. It makes sense that if a full-on quantum computer exists in nature, the brain (or components of it) would be the most exciting and promising place to look.

  Cognition is increasingly recognized to be “quantum-like” in numerous ways. For instance, when viewing an ambiguous image like a Necker cube or a duck/rabbit, the viewer only sees one aspect at a time, not both—an either/or that oscillates from moment to moment. 37 Words or other items learned in memory experiments have multiple potential links to other cues—akin to superposition—until a test is administered, which effectively forecloses nonrelevant associative links. 38 Various fallacies and heuristics in probability judgment also seem best modeled quantumly. For instance, the order in which information is presented to a test subject constrains the outcome of that person’s decision. 39 Thus the latest thing in cognitive psychology is the framework of “quantum cognition,” describing the processes of perception, memory, or judgment in quantum computer terms, along with a bold disclaimer that “quantum” is just a convenient and suggestive metaphor. But plenty of researchers have been keen to prove that the metaphor is more than a metaphor.

  In the 1980s, physicist Roger Penrose argued that the brain must literally be a quantum computer after pondering mathematician Kurt Gödel’s incompleteness theorems : Any formal mathematical system will contain statements that are unprovable within the system, and no such system can prove its own consistency internally. Penrose reasoned that only a quantum computer could arrive at the idea that computation can never be complete, and thus the brain—or at least Kurt Gödel’s brain—must be such a device. 40 His reasoning was analogous to Hinton’s argument that only a brain with “molecules” that reached across the fourth dimension could think in four-dimensional terms. Like Hinton, Penrose was clearly ahead of his time. Since Stuart Hameroff had already identified a likely quantum culprit with his microtubules, Penrose collaborated with the anesthesiologist to formulate what they call the “orchestrated objective-reduction” (Orch-OR) hypothesis, in which neuronal microtubules create consciousness through brain-wide entanglement. 41

  The narrow channels in neuronal walls that control the movement of ions into and out of the cell—and thus the cell’s action potential (the electrical charge that passes down it when it fires)—have also attracted attention, as these are potentially sites where particles are protected from environmental interference long enough that they could become entangled. 42 Already in the 1970s, physicist Evan Harris Walker had proposed that consciousness depended on quantum tunneling at the synaptic cleft, the narrow gaps where molecules carry signals from neuron to neuron. 43 There is no reason why more than one mechanism could not play a role, or that there may be others that no one has even imagined yet. The brain’s quantum computation could take multiple forms. The problem with existing theories, however, is that there is no known way that quantum entanglement could be preserved throughout or across the whole brain, which intuitively might seem to be a requirement for “quantum consciousness.” 44

  Again, though, it may be that the search for the physical roots of consciousness per se, even in quantum biology, is a hopeless task, simply because of the rift I mentioned earlier between the necessarily reductive methods of objective science and subjective descriptions of experien
ce. No matter how fine-grained neuroscientists’ understanding of the brain becomes, it may never map convincingly onto qualia , or “what it feels like” to be a conscious entity. This is what philosopher David Chalmers called the “hard problem” of consciousness. 45 Nevertheless, as a MacGuffin spurring competitive efforts from various researchers in different fields, a byproduct of efforts to discover the roots of consciousness in the brain may be the discovery of quantum processes in or between neurons that might enable the brain to process information across its timeline.

  In the spirit of Charles Howard Hinton, here’s a sketch of what a brain-based account of precognition might look like in the coming era of quantum hyperthought.

  We can start with what is already generally agreed upon about how memory works in the brain. “Memories” do not exist whole and discrete, like fossils of past experience tucked in the brain’s folds. Different experiences activate, and thus can share, many of the same neurons and connections, across many areas of the brain. Although it is a simplistic (even simple-minded) metaphor, individual neurons could be thought of as “pixels” in our mental life and experience; just as a TV screen reuses the same pixels in different ways from instant to instant to create sequential still images in a changing picture, the brain reuses the same cells and circuits in different configurations from moment to moment to generate the flow of our thoughts and experiences. What seems to distinguish one memory from another, or one experience from another, is a unique spatial and temporal pattern of neuronal activation across the brain. That pattern is determined by the variable strengths of synaptic connections among all those linked neurons.

 

‹ Prev