Читать книгу A Physical Principle of Universal Order - Jaime S. Carvalho - Страница 6

An understandable universe

Оглавление

Although this work is mainly concerned with scientific concepts, we cannot ignore their philosophical and religious foundations, particularly in a topic concerning the understanding of the universe and our role in it. After all, these three disciplines should be considered not isolated branches of knowledge but phases in the continuous development of the human intellect. We cannot fully understand philosophy without religion, and science without the influence of both.

With language not yet developed and very limited thinking ability, our ancestors of 50,000 to 20,000 years ago were guided mostly by their perceptual and instinctual processes. The notion of self—of a person differentiated from nature—was lacking.

By their own experience, they “knew” that nature was in constant change—hotter and colder days, strong and weak winds, occasional storms, flowing rivers, growing and decaying trees, and the like—but had no inkling how this change came about. The elusive concept of change could not be grasped by our forebears and, therefore, the developing language was necessarily built on concepts of permanence.

Unable to understand the external world, they lived an existence of fear, not only of natural phenomena but of hunger, wild beasts, disease, and death. To mitigate these fears, they conceived of entities with super powers—living spirits inhabiting trees, mountains, and rivers—as the rulers of the planet. In times of need, they appealed to those illusory beings for help. As time went by, these concepts found their way into the tradition, leading to the first cultural era of superstition, remnants of which still exist today.

With language evolving and abstract thinking more generalized, the primitive, vaguely defined concept of living spirits was later replaced by a belief in extraterrestrial gods possessing supernatural powers—polytheism. But whereas the conception of spirits was put forth in a context of animism, the conception of gods was developed in a context of anthropomorphism. These humanized gods were the supervisors of the world and everything in it. To propitiate these beings and secure their favors, sacrifices were offered to them. As this belief spread into the culture, the number of gods expanded, eventually resulting in a hierarchical system of deities. The highly emotional idea of all-powerful gods was stabilized by the formation of a special priestly caste that sets itself up as the mediator between the people and the gods.

By 10,000 BC agriculture and stock breeding were well developed, and the guaranteed food supply changed the way of life from nomadic to urban. With urban living, language increased in importance, knowledge expanded faster than ever before, and leisure time was more readily available. Human activity was particularly intense in the eastern Mediterranean regions, Middle East, and India, in places where the soil was most productive.

By 1,000 BC conceptual thought could be expressed in script as well as speech, a social hierarchy was well established, but man continued to be a part of nature. Nevertheless, humans (presumably the Egyptian priesthood) became capable of achieving universal ideas, the most important one being the concept of a single universal god—monotheism. This single authority was the creator of the universe and of everything on earth, and an absolute ruler possessing limitless powers.

After 1,000 BC the material and economic conditions were improving, but their unfair distribution led to social unrest in a society where instinctive behavior was still predominant. The progressive degeneration of customs led the priesthood to develop the concept of a personal god, a god who protects, disposes, rewards, and punishes. God was always present within the individual, in a divine part of the brain called soul, which was considered to survive the death of the body. The old idea of spirits in nature was converted into spirits in human nature—conscious spirits. The tradition carried these concepts forward, ultimately leading to the creation of all the great monotheistic religions. To some extent, religious thought has influenced humans ever since.

With the teachings of monotheism, particularly the idea of guilt, man became aware of sharp divisions within himself—good and evil, reason and instinct, body and soul. He saw himself as a thinking person separated from nature—idea could be clearly distinguished from fact. The concept of “self” had finally been established. Humans could now concentrate their attention on the surrounding world.

By 600 BC, with man increasingly attracted to the novel field of thinking processes in his own nature, philosophical thought started appearing. With the advent of philosophy, the supernatural and mythical explanation of the origin of the natural world was challenged by rational explanations, which were still vague and diffuse but more related to the environment. Religious ideas, however, continued to underlie most of philosophic thought.

In philosophy, the idea of spirits and soul was replaced by a more refined but equally obscure one, that of mind. As first conceived around the fifth century BC, the mind was a material substance but finer and purer than the rest of matter. It was an independent, changeless entity where all knowledge and power resided.

Without any means of experimentation available to them, the philosophers of nature arrived at their conclusions either by close observation through the senses, by introspective “rational” thinking, or by a judicious combination of both. Some philosophers relied on their senses as a reliable instrument of inquiry, others relied more on the judgments of their own minds.

In a very simplified way, and after a few centuries of wild speculation regarding the nature of the world and our place in it, we can say that there was agreement among the philosophers that change was constantly occurring in nature—it could not be denied—but the process of change was interpreted differently: some view it as simply chaotic, others were able to see a principle of order beneath the chaos. These views were to characterize two distinct types of philosophic thought.

The most influential representative of the first type was Plato (428–327 BC). He rejected the ruthless world of the senses in favor of a transcendent world of permanent ideas. Mind was separated from the material world—a special type of dualism. For him, the static harmony of thought was reality and the confusion of the natural process was illusion. This dualistic-static thought, well fitted to religious concepts, was to survive until our days.

The best representative of the second type of thought was undoubtedly Heraclitus (535–475 BC) who explicitly stated that “everything is in flux,” like the constant flow of a river. He saw a special type of order in change, a continuous state of dynamic equilibrium, of which man was part. Change was conceived as the harmonic interplay between opposites—living and dying, heating and cooling, day and night. The opposition was not antagonistic but complementary: remove day and night goes too. Heraclitus’ thought was of the unitary-process type where nature and man formed a single changing whole, but his concept of change was too primitive and ill-defined to be convincing. Heraclitus influenced many other philosophers but with humans afraid of any kind of change, his ideas slowly became stagnant. It would take about a millennium for them to be revived.

It is interesting to note that the concept of a basic principle as the source and foundation of everything in nature had been present since the beginning of philosophy. It was first conceived in the static terms of a single material substance, such as water, and not in the process terms of a continuous change as envisaged by Heraclitus. Nevertheless, the idea of a single cause underlying the universe is remarkable and must have been motivated by an intense human need to sought order in the confusing complexity of nature.

Another important contribution of Greek philosophy to the understanding of the world, not by way of single principles but by an insight into its matter constituents, was that of the atomists, its best representative being Democritus (460–370 BC). In their view, the universe consisted of empty space and an infinite number of independent, indivisible, and eternal atoms which interacted with each other mechanically. The existence of atoms was later confirmed by scientific physics, though the present quantum mechanical atomic concept is very different from its philosophical model.

With the conquest of Greece by the Romans, Greek philosophy came to a halt and was subsequently greatly obscured by the rise and spread of Christianity. Throughout the Middle Ages, the universe was explained by a blend of religious and concordant philosophical concepts but between AD 1600 and 1650, a dramatic transformation occurred: the establishment of scientific thought.

The tool enabling that historical revolution was the scientific method of investigation, based on quantity. The agents were Kepler (1571–1630) and Galileo (1564–1642), among others. They proclaimed that measurement holds the key to the understanding of nature. With the development of analytical geometry by Descartes (1596–1650), the way was opened for Newton (1642–1726) to discover the laws of motion and universal gravitation, which laid down the foundation for classical mechanics.

Newton derived his laws of motion from the observation of the movement of the planets, and in the mathematics he assumed the reversibility of time. By “reversible” it is meant that the laws governing the process remain unchanged when the direction of time is reversed—when -t is substituted for +t in the equations. In the real world, this means it does not matter to the laws of motion whether the planets rotate from left to right or from right to left. After confirmation of Newton’s laws, it was assumed by physics that all elementary processes in nature were reversible.

But by experience we know that in our universe time is not reversible. We grow older with every breath and cannot get younger. The nonreversibility of time occurs with every process in nature. We cannot, for instance, fry an egg and then put it back raw into the shell. The reason why Newton’s laws of motion work so well is that they deal with simple two-body systems (earth-moon, sun-planet) under conditions where irreversibility effects are negligible. It is in the passage from two-body to many-body, from simple to complex systems, such as proteins and organisms, that the irreversible nature of time becomes manifested. In effect, we can say that all processes in nature are irreversible, the apparent reversible processes being just limiting cases. But so far, all attempts to introduce irreversibility into contemporary physics have been unsuccessful.

During the nineteenth century, following extensive work in electricity and magnetism carried out by several physicists and mathematicians, Maxwell (1831–1879) introduced the concept of electromagnetic waves and, by declaring that all energy resides in the field, set down the basis of field theory.

Electrical concepts differ from those of mechanics in a very significant way: they are not supported by direct representations of visual observations. The concept of electric charge is a good illustration of this point. Presumably it holds some essential clues to fundamental structure, but this fact has not been definitely established and until it does, electrical charge remains an abstract concept. The introduction of abstract concepts into fundamental theory has compromised its objectivity and the reliability of its mathematical expressions. This tendency has increased ever since.

In mid-nineteenth century, with the invention of the steam engine, it was recognized that heat always moves by itself from hotter to colder bodies. And it soon became evident that all forms of energy move in a unique direction, from a higher to a lower energy state. The inability of the mechanical physics of the time to explain these observations led to the introduction of thermodynamics theory in late nineteenth century.

Thermodynamics is the branch of physics concerned with heat and temperature and their relation to energy and work. It applies to a wide variety of topics in science and engineering and, since organisms are heat-producing systems, it also applies to us. In effect, its first law, which embodies the principle of the conservation of mass-energy, forms the basis of present biology.

But it is its second law that possesses a unique position among the laws of physics: it is the first law to impose on time a sense of direction. It states that in any closed system, there must be either conservation or increase in entropy. In other words, entropy production can only be positive, which means that entropy increase is irreversible. Entropy has been loosely correlated with the degree of randomness or disorder within a system (water molecules in steam have an entropy level higher than in liquid water). Applied to the universe as a whole, considering it as a closed system, the second law predicts that “the entropy of the universe always increases.” In fact, it was in these terms that Claudius (1822–1888), one of the founders of thermodynamic theory, formulated the second law. Since then and for over a hundred years, physical scientists have paid much attention to processes moving toward states of greater dynamical disorder, leading them to even suggest that the universe displays only one tendency: toward disorder.

In the early twenty century, a complex mathematical theory —relativity theory—was developed by Einstein (1879–1955). It was to change the course of theoretical physics. Since the concern of this book is the place of human beings in the vast system of nature, we just want to address two pertinent assumptions of relativity.

In the special theory, it was postulated that physical laws can best be expressed if it is assumed that space and time are so similar that physics can make no absolute distinction between them—the spacetime concept. Under these conditions the symmetry of space involves the symmetry of time, and therefore the reversibility of the physical laws. As in Newtonian theory, time in relativity continues to be reversible and, in addition, space becomes less real: a four-dimensional continuum does not fit the space surrounding us. We perceive spacetime as space and time, the spacetime notion being meaningless to us. The other assumption concerns the origin of gravity, the central idea of the general theory. While for Newton it is mechanical—an invisible force attracting two material bodies—for Einstein it becomes geometrical—a curvature of the very fabric of spacetime, a distortion induced by mass itself. This concept of gravity has proved very difficult to reconcile with other theories. It thus appears that relativity theory is only applicable to a narrow range of relatively simple systems, at the level of stars and galaxies. Certainly, it does not apply to complex systems, such as human beings.

Ever since Newton, the nature of light—particle, wave, or both at once—was a subject of intense interest to physicists. In 1900, Planck (1858–1947) noted that a blackbody emitted electromagnetic radiation (which includes light) in small discrete packets, later called quanta, rather than as a continuum emission—energy was quantized. His law, giving the distribution of the radiated energy, formed the basis of quantum theory. Einstein later theorized that a beam of light is not a wave propagating through space but a collection of discrete wave packets, which he called photons, whose energy content is proportional to the wave frequency. He then demonstrated that one sufficiently energetic photon can transmit its energy to a single electron in a metal, ejecting it. In 1923, De Broglie (1892–1987) showed that electrons can be diffracted in a similar way to light: that is, particles can act as waves—the wave-particle duality. Photon-electron—or more generally radiation-matter—interactions form the basis of modern quantum mechanics theory.

To comply with the requirements of wave-particle duality, in quantum mechanics an electron is represented by a complex quantity called a wave function, based on the conservation of energy and momentum. The wave function cannot be identified with a single physical property of the electron. Unlike particles, waves do not have an easily defined spatial and temporal position. Thus, for an electron to have wavelike properties, its position cannot be determined with certainty and all its physical attributes cannot be expressed in a deterministic manner, requiring statistical treatment instead. As a result, in quantum mechanics all physical laws are inherently statistical. Quantum reality shows that at a more fundamental level, the world is not Newtonian but it is governed by notions such as chance, probability, and uncertainty. Moreover, the theory insists that we cannot make a measurement without influencing what we measure—the observer becomes an active part of reality. The irreversibility of time is admitted but it is introduced into theory by way of the fictitious observer and is not intrinsic to matter—it is anthropogenic, not materialistic.

Although waves may not have objective existence, the abstract wave equations of quantum mechanics have provided excellent approximations for a vast range of systems, from crystals to atoms. But the theory can only make accurate predictions for systems at the subatomic and atomic level, where energy is low and, above all, the number of components is small. The physical principles involved in the theory appear to be inappropriate for the description of very complex systems with a great number of degrees of freedom such as those found in biological structure. Quantum mechanics is thus an incomplete, complex, and indirect description of reality.

Theoretically speaking, modern physics possesses two major doctrines of space and time, general relativity and quantum mechanics, each applying to extremes of the magnitude scale of the universe—the macro and the micro world—but lacks a theory for the world in between, the one we inhabit. Attempts have been made—and continue to be made—to generalize, reconstruct, or build a new and more comprehensive quantum mechanics, but the various approaches so far pursued in these directions have been unsuccessful. The unification of quantum mechanics and general relativity has also been intensively sought, but the so-called “theory of everything” has not been achieved. A major obstacle in these attempts appears to be the passage from simple to complex systems.

In the first half of the twentieth century, it was recognized that the concept of a purely entropic universe did not make sense. Human beings living in such a universe would be in a constant struggle against nature, pushing things uphill and powerful laws bringing them down over and over again. Under these circumstances, life could not be sustained. This is against the order observed over billions of years of successful organic evolution. Among the scientists who recognized this paradox was the quantum physicist Schrödinger (1887–1961), the composer of the wave equation, who was looking at the phenomenon of life from the point of view of physics. To describe those obviously orderly processes, he invented the term “negative-entropy.” Order in organisms was maintained by an intake of “negative-entropy” from the environment, contained in the food they ingested. In this way, the universe became a mixture of order and disorder.

Since the latter part of the last century, the introduction of powerful electron microscopes and telescopes made it evident that this world contains countless ordered or partly-ordered spatial units—atoms, molecules, macromolecules, organelles, cells, organs, organisms, stars, solar systems, galaxies, clusters of galaxies—distributed all over the visual space of the organic and inorganic realms. Although less evident, these spatial units are organized in complex hierarchies in every organism and in the universe as a whole. To explain the origin of these units and hierarchies of units and their development over time, a new kind of general theory, a one-way (irreversible) science of changing structure accounting for all the patterns in the universe, is required. To be successful, this theory should be built on the geometry of tridimensional space—the space we really see. It should be simple and capable of explaining all known partial theories, including quantum mechanics and thermodynamics. Is such a grandiose theory possible?

It is not only possible but its blueprint already exists, although present physics has not recognized it as a legitimate theory. To openly admit the irreversibility of natural processes, physics would have to renounce the whole system of Newtonian concepts on which the ideas of quantum theory and relativity are rooted, and so far it has been reluctant to do so. Matter, energy, forces, interactions, and wave properties are not appropriate for the description of irreversible effects. A whole new set of concepts is required to deal with one-way processes.

Contrary to all available theories of classical physics, this unique one-way field theory was primarily derived from observations of biological structure. It is a theory of change built on a sound scientific foundation. Its basic concepts were developed over the centuries, from Heraclitus and Aristotle (384–322 BC) all the way to Ernst Mach (1838–1916), Pierre Curie (1859–1906) and Bertrand Russell (1872–1970), but it was only in the middle of last century that they were put together by the Scottish physicist Lancelot Law Whyte (1896–1972) and later extended by the American physicist and psychologist Leo John Baranski (1926–1971). The theory is the most general ever conceived, in fact universal, and therefore applicable to all physical, organic, and social systems. Its realm is not that of quantity but of order.

For the purpose of this work, we are only interested in those systems where there is structural change over time with formation and extension of ever more complex structural patterns. The concept of order described below refers to the order found in these systems.

A Physical Principle of Universal Order

Подняться наверх