Science & Religion Spirituality & Health Character Devlopment Free Enterprise Grant Opportunities

Journal of RELIGION & SCIENCE

VOLUME 31, NUMBER 3 - Sept. 1996

PHYSICS: WHAT DOES ONE NEED TO KNOW?

by John R. Albright

Abstract. For the basic areas of physics -- classical mechanics, classical field theories, and quantum mechanics -- there are local dynamical theories that offer complete descriptions of systems when the proper subsidiary conditions also are provided. For all these cases there are global theories from which the local theories can be derived. Symmetries and their relation to conservation laws are reviewed. The standard model of elementary particles is mentioned, along with frontier questions about them. A case against reductionism in physics is presented.

Keywords: classical field theory; classical mechanics; collective phenomena; conservation laws; determinism, dynamical theory; Lagrangian mechanics; Maxwell's equations; Newton's laws; parity; phase transition; physics; quantum mechanics; quark; reductionism; Schrödinger equation; standard model; symmetry; teleology; uncertainty principle; wave function.

Physics is the subject in which we consider the simplest systems and then attempt to analyze them completely. Such a description of physics may seem not entirely serious in view of the reputation for difficulty that the subject has earned. But the systems are really chosen by physicists for their inherent simplicity; the difficulty comes with the attempt at completeness of description, since it has been found that mathematics is the descriptive vehicle that most clearly summarizes what is happening.

In view of this definition, I will examine various fields of physics that have achieved a certain amount of success at the program of complete analysis and understanding. For each special area of physics, certain questions are considered:

  1. What is the historical background of the topic, and how has it been put to practical use?
  2. What constitutes a complete description of a system?
  3. What is the dynamical theory on which a complete description may be based?
  4. What additional conditions are needed to augment the dynamical theory?
  5. Is there an alternative to the dynamical theory, based on a global rather than a local view?

This set of questions emphasizes the importance of the theoretical structure of physics. It must be remembered that physics also is an experimental science, and that theories are not just ideas that have been made up by antisocial recluses who wear baggy sweaters and no socks. Theories are not valuable if they cannot make any connection with the real world. In this essay I shall limit the discussion to theories that have been tested often, with considerable sensitivity, and that have passed the tests: that is, they have been exposed to the risk of falsification and have survived. I shall not be able here to summarize the experimental basis of these theories. I shall also not attempt to survey all of physics, but instead I will concentrate mainly on those theories that Roger Penrose (1989, 152) called "superb" because of their wide-ranging applicability, their accuracy, and their beauty. Some attention will be paid to theories that are not yet complete in the same way that the superb theories are. Because of space limitations I shall not consider cosmology in any detail.

 

CLASSICAL MECHANICS

The oldest superb theory in all of science is classical mechanics, the macroscopic theory of motion from the seventeenth century -- the theory of Galileo Galilei, René Descartes, and Isaac Newton. It does a remarkable job of describing systems whose size corresponds to human scale; only for the very smallest and very largest systems do we need a better theory. Classical mechanics is useful as the basis for mechanical engineering. It enables the precise calculations of the orbits of everything in the solar system; it makes possible the prediction of the motions of the planets years in advance of when they actually happen. Its principles govern much of what we encounter in our daily lives.

A system in classical mechanics is construed as a set of points -- each with its own mass -- that are in motion; a complete description is achieved by getting a set of equations that tells us the space coordinates of each point as it evolves in time. These equations constitute the trajectory of system. This description is complete in the sense that it enables calculation of the past and future motion of the parts of the system. Given enough mathematical prowess, this information can be used to calculate other quantities of interest, such as the velocity, the acceleration, the energy, the momentum, and the angular momentum. Since the time dependence is completely specified, the system is deterministic, with all the connections to religion and philosophy that are implied by that term.

 

The dynamical basis for obtaining the trajectory of a classical system is the set of equations obtained from Newton's laws, the equations of motion. They are differential equations of the second order, so that calculus is needed to set them up and to solve them for the trajectory. Here lies part of the reason why physics has a reputation for difficulty. The differential equations form a local theory, since they are applied to the system at a specific point in space and time. The opposite would be a global theory, in which all points in space and time are considered together.

To find the relevant solution to the equations of motion, we need to specify initial conditions. Like most differential equations, the equations of motion in classical mechanics have an infinite number of possible solutions, only one of which corresponds to the actual motion of the system. It is the initial conditions that render a solution unique and make determinism a possibility. The initial condition for a single point mass is usually a specification of the position and the velocity of the mass at a starting time. At this point it must be made clear that we do not need to know how to solve differential equations in order to teach the relation between science and religion. But it would help if -- in a private setting -- we made use of a computer to follow a simulation of a relatively simple system that evolves according to Newton's laws, starting from a given initial state. By changing the initial conditions, we would see a different time evolution pattern. Some systems are so stable that a slight change in the initial conditions leads to an almost imperceptible change in the development of the system. Others are so unstable (the technical term is chaotic) that their evolution is drastically different if the initial conditions are altered even a little.

Alternate formulations of Newton's laws were developed in the eighteenth and early nineteenth centuries by Moreau de Maupertuis, Leonhard Euler, Joseph Louis Lagrange, and William Rowan Hamilton. I will describe here the approach of Euler and Lagrange. We begin by writing the Lagrangian for a given system. For certain simple and important examples, the Lagrangian depends only on the positions and the velocities of the particles that make up the system; it is calculated by subtracting the potential energy from the kinetic energy. Next we construct the sum of all Lagrangians for all the points between (1) an initial point in space and time and (2) a final point. This summation is called the action.

There are many ways to calculate the action for a particular pair of points, depending on the path taken between the beginning and the end. The path that extremizes the action is the one actually used by the system for its motion. The calculus of variations is the branch of mathematics that we use to go from an extremized action to a set of equations, which turn out to be identical to those obtained directly from Newton's laws (Yourgrau and Mandelstam 1968). This Lagrangian procedure is a global approach to mechanics, since it requires (in principle only -- we never really have to calculate all possible action quantities) calculating the action through all possible paths in space and time. As mentioned earlier, a local approach calculates quantities at a single point and uses them to figure out what will happen next. Nevertheless, the global approach leads to the same complete description that we get from the local approach. Global approaches are sometimes referred to as "modern teleology" (Barrow and Tipler 1986) because the system seems to act as though it knows it has to minimize (sometimes "maximize" should be used instead) the action, just as water acts as is it is supposed to minimize its potential energy by flowing toward the sea.

 

CLASSICAL FIELD THEORIES

If instead of a collection of massive points we consider a physical quantity (e.g., pressure, electric field, magnetic field, height of water above or below the mean level) that can be defined for each point in space and time, then this field quantity can be treated in a way analogous to the displacement coordinate of classical mechanics. Theories of acoustics, electricity, magnetism, light, and fluid flow all can be considered in this way. To be specific, I will use James Clerk Maxwell's theory of electro-magnetism as an exemplar for this section. The experiments of Benjamin Franklin, Charles Augustin de Coulomb, Alessandro Volta, Hans Christian Oersted, Michael Faraday, and Joseph Henry all helped guide the way to a synthesis in which electric and magnetic forces are described by field concepts. The electric field is defined at each point by a vector with three components corresponding to x, y, and z, the three space coordinates; the magnetic field has a similar set of three components. Experiments by Oersted, Faraday, and Henry showed conclusively that the electric field and magnetic field are not independent of each other whenever either one of them is changing in time. Maxwell synthesized a theory of both electric and magnetic fields and showed that these inseparable fields can propagate together as waves traveling at the speed of light. They are light. Maxwell's theory describes light considered as any electromagnetic wave, visible or not. Whether the human eye can see the wave depends entirely on the wavelength. X rays, ultraviolet, infrared, microwaves, and radio/television signals are just as much light as visible light. Practical application of Maxwell's electromagnetic theory includes much of electrical engineering -- telephone, radio, television, radar, microwave ovens, and the lighting that allows us to see when it is dark.

A complete description of an electromagnetic system requires that we know the three components of the electric field and the three components of the magnetic field as functions of space and time for all points. The dynamical theory that governs such systems is the set of four equations (as written in their usual vector form) called Maxwell's equations. They constitute the local theory, since all four are valid at each point of space and time. As in the mechanical case, they have an infinite number of solutions, and so the physical description of the fields requires additional conditions, analogous to the initial conditions of classical mechanics. These are called boundary conditions, since they often involve knowing the properties of the fields on a surface that surrounds an interior region; the fields are known on this surface but are unknown in the interior volume. With the given conditions and with Maxwell's equations it is possible to arrive at the complete description.

The Lagrangian approach works for electromagnetism. The Lagrangian must be replaced by a Lagrangian density (i.e., the amount of Lagrangian per unit volume), which is then turned into an action by summing over both space and time. The calculus of variations is next used to extremize the action to a maximum or a minimum. It should be no surprise to learn that the result of all this is the reappearance of Maxwell's equations (Jackson 1975, 597). Once again, the global theory contains the local one.

The general features of electromagnetic theory presented here are typical of other classical field theories. A feature that often appears in such theories, including electromagnetism, is the propagation of waves governed by a wave equation that results from the theory. In all these cases there also is a global form of the theory.

 

QUANTUM MECHANICS

In 1925 and 1926, Werner Heisenberg, P.A.M. Dirac, and Erwin Schrödinger invented three different approaches to quantum mechanics, a new theory of matter in the small. It was quickly recognized that these formulations were consistent with one another. A fourth consistent approach was invented later by Richard Feynman. Quantum mechanics rapidly became the method of choice for describing systems the size of molecules or smaller. The theory is not outrageously difficult, but it is unfamiliar to those who are used to classical physics. It has proved to be extremely practical, since without it there would be little or no understanding of atomic physics: lasers would be a total mystery. Those working in physics of the solid state would have been unable to make the progress that has led to transistors, integrated circuits, chips, and the associated technology of computers and communication. Nuclear technology also depends on quantum mechanics to provide the basis of understanding.

When we use the methods of Schrödinger, the complete description of a quantum mechanical system resides in the wave function, psi, which is a field quantity very similar to those described in the preceding section. If we know psi as a function of time and the space coordinates, we know all that is knowable about the system. It happens that quantum mechanics does not permit us to know some of the things that classical mechanics allows. The celebrated example is that of Heisenberg's uncertainty principle, which states that we cannot know simultaneously both the position and the momentum of a particle to arbitrary accuracy. The best we can do is to calculate probabilities of finding the position within a certain range or the momentum in its range. It follows that Newtonian determinism is impossible in quantum mechanics, since the initial conditions required in classical mechanics to compute the trajectory are made unavailable as a result of the uncertainty principle. It must be emphasized that the uncertainty principle is not a starting point for quantum mechanics. It is a consequence of the basic assumptions of the theory. Heisenberg began by seeking a theory that would emphasize those aspects of atoms that could be directly observable; he was then led to a set of assumptions that constitute the theory. Not until two years later did he find that these assumptions implied the uncertainty principle, and the logic is inexorable. If the uncertainty principle is wrong, then quantum mechanics must be abandoned.

According to quantum mechanics, a system wherein the energy does not depend on time will get into a stationary state and stay there until it emits a photon (a small package of electromagnetic energy) and makes a sudden transition into a state of lower energy. There may in fact be several such states of lower energy into which the transition may occur. Quantum mechanics says that for any single example of a transition, it is not possible to predict into which state the system will jump; what is possible is to calculate the probabilities for jumping into the various states. If the experiment is repeated often enough, the frequencies of jumping agree statistically with those predicted by quantum mechanics. Again, the explicit denial of determinism is unavoidable.

The dynamical equation that we use to find the wave function is either the Schrödinger equation or its relativistic generalization, the Dirac equation. These equations have both been remarkably successful at enabling the calculation of the properties of atoms and molecules. They have been somewhat successful but less impressive in the analogous task for nuclei and elementary particles. In any event, we need to have boundary conditions in order to find the relevant solution (and discard the irrelevant ones) of the Schrödinger equation or the Dirac equation. The procedures bear a close resemblance to those of classical field theories; mathematical methods devised by Lord Rayleigh for analysis of classical light waves or sound waves have been taken over by quantum mechanics with little alteration. Furthermore, the Lagrangian formalism works in quantum mechanics in a way that resembles classical field theory. We might have expected such a property of the Schrödinger equation or the Dirac equation, since they are specific types of wave equation, mathematically akin to classical field theory.

SYMMETRY AND CONSERVATION PRINCIPLES

  Physicists are very fond of conservation laws. If we can be sure that a certain physical quantity remains unchanged even though the rest of the system is changing drastically, then we say that the quantity is conserved. Such laws appear both in classical and in quantum physics. An important reason for stressing the Lagrangian in the preceding sections is that it is the best way to see the connection between symmetries and conservation laws. Noether's theorem states that if the Lagrangian has a specific type of symmetry, there will be a corresponding conservation law for the system described by that Lagrangian. Already in the case of classical mechanics we can see examples of this truth. If the Lagrangian does not depend explicitly on time, then the energy will be conserved. The symmetry is one of time translation: as time passes, the Lagrangian remains unchanged; therefore, it is symmetric in time. A similar principle arises for each space coordinate: if a change in the x coordinate leaves the Lagrangian unchanged, then the x component of the linear momentum will be conserved. In other words, invariance or symmetry under space translation leads to conservation of momentum. Invariance under rotation through an arbitrary angle implies the conservation of angular momentum.

A more subtle type of symmetry, called gauge invariances, arises in Maxwell's electromagnetism. The invariance already is present in the local form of the theory; if we perform a special type of transformation (a gauge transformation) on the electric and magnetic potentials, the electric and magnetic fields will remain unchanged, and there will be no observable physical consequences of the transformation, since the fields contain a complete description of the system. But if the transformation is done on the Lagrangian, it leads to conservation of electric charge, one of the most revered principles in all of science.

The great respect that physicists have developed for conservation laws might lead one to think that nearly every physical quantity obeys such a law. Such a belief would be mistaken, since there are many complicated combinations of physical quantities that are not conserved. They are unrelated to any symmetry of the Lagrangian.

There are three discrete (as opposed to continuous) symmetries that have been studied in great depth in the second half of the twentieth century: parity (P), charge conjugation (C), and time reversal (T) (Zee 1986). Parity refers to the transformation in which all three space coordinates (x, y, z) have their algebraic sign changed. This transformation incidentally changes the left hand into the right. Charge conjugation changes a particle into its antiparticle (electron into positron, proton into antiproton, and so on). Time reversal, as the name implies, causes the clock to run backward (Sachs 1987).

All of the theories considered above exhibit invariance under C, P, and T separately or in any combination. It was once thought that all of physics had to be invariant under all three symmetries, but the carefully constructed suggestion of T.D. Lee and C. N. Yang in the 1950s led to several experiments that demonstrated conclusively that P is not conserved in the weak nuclear interaction. It was and is still believed that physics is invariant under the product CPT; efforts to construct a theory without this invariance have failed. So it follows that if P is violated, then at least one of the other two symmetries must also be violated. The correct answer is C; when p is violated, C is also. The combination of CP (and therefore T) is conserved most of the time, but in the 1960s an experiment was performed that was the first in a sequence to show that a small violation of T can exist.

These discrete symmetries and their violations may seem unimportant, but they point to large cosmological issues: why is it that we are made of protons, but antiprotons are so rare? Is there an imbalance of matter over antimatter in the universe? Or is the imbalance just a local effect that is averaged to equal amounts of matter and antimatter if we looked at more of outer space? If the imbalance is global, then how did the universe get this way?

Another example of the breaking of a discrete symmetry occurs in biological systems, where a specific preference for right-handed or left-handed molecules occurs in apparent violation of parity. Sugars, amino acids, and DNA all exhibit this effect. On a larger scale, handedness appears in the development of flounders and in the bicameral human brain.

 

ELEMENTARY PARTICLES

The notion that all matter is made of many copies of a small number of building blocks can be traced to Democritus in ancient Greek culture. In the nineteenth century, the advances of chemistry led to the idea that atoms really exist and that they are the fundamental building blocks. The discovery of the electron and of the nucleus of the atom changed all that. The nucleus itself -- after 1932 when James Chadwick discovered the neutron -- is known to consist of a certain number of protons and neutrons. Nuclear physicists saw their task as the elucidation of the forces that hold the neutrons and protons together. To gain insight into the situation, they bombarded nuclei with projectiles, using ever-increasing energy in the mistaken hope that they would get the resolution to "see" the structure of the nucleus. Instead, they saw a great variety of species of particles -- enough to threaten exhaustion of both the Latin and Greek alphabets to find symbols to designate the discoveries. Clearly, not all of these new particles could be fundamental in any meaningful sense.

A consensus has been achieved, called the standard model, in which the fundamental building blocks are called quarks. There are six types of quarks (up, down, charmed, strange, top, and bottom), and only the first two types are needed to explain conventional nuclear physics. A proton is not really an elementary particle; rather, it is made of three quarks (two up-quarks and one down-quark). A neutron is made of two down-quarks and one up-quark. Quarks are held together by gluons (there are eight species). Also fundamental are leptons, of which the six types are parallel to the six types of quarks. Both quarks and leptons obey the Dirac equation of quantum mechanics; therefore, each species has a corresponding antiparticle that has the opposite sign for its electric charge and various other properties as well. Mesons are particles once thought to be elementary but which are now understood to consist of a quark and an antiquark held together by gluons.

The standard model answers certain questions about fundamental particles but leaves others to be answered by future research. For example, why are there exactly six species of quarks? Why do the quarks and leptons have the masses that are observed? Are all these particles really fundamental, or is there a deeper layer for us to probe? Will we learn that there is a much smaller number of really fundamental particles, and that the ones we see are only composites?

ANTIREDUCTIONISM

  Physicists often are accused of being reductionist -- believing that physics could in principle explain all the more complicated sciences, if only the time and the funds to figure out how were available. Dirac (1929) expressed such a vainglorious view: "The underlying physical laws necessary for the mathematical theory of a large part of physics and the whole of chemistry are thus completely known, and the difficulty is only that the exact application of these laws leads to equations much too complicated to be soluble." Such an attitude is not likely to make a physicist popular with colleagues from other departments. But it is misguided for other reasons, since physics contains within itself strong weapons to use against reductionism. Three examples follow.

Exchange symmetry or antisymmetry is the first example of how reductionism is misleading. We can study the properties of a single isolated electron and feel that we have achieved the kind of complete understanding so desired in physics. Such study does not prepare us for what happens when two electrons are in the system. Their wave function is antisymmetric: it must change its algebraic sign every time we exchange the two electrons -- a fact that does not follow from the one-electron theory but which has profound significance for the rest of science. Nuclear bonding, atomic structure, and molecular bonding are all possible because of the exchange antisymmetry of electrons.

A second example is the more general category of collective effects. When large numbers of atoms get together, we can observe phase transition -- solids melting to liquids, liquids boiling to gases, and so on. The sharp discontinuities from these phase transitions are not predictable from the one-atom theory. Condensed matter physics is replete with other examples of phase transitions involving order versus disorder: feromagnetism, superconductivity, superfluid helium, and so on.

A third example is that of symmetry breaking. The parity violations seen in various organic molecules are not derivable from the single-atom theory. Inorganic molecules can show such effects at a simpler level: the water molecule has an electric dipole moment (i.e., one end of the molecule has a positive charge, the other a negative charge), a fact that appears to violate both parity and time-reversal invariance if we know only the properties of hydrogen and oxygen by themselves. Collective interaction yields qualitatively new phenomena. For all of its beauty, elegance, intricacy, practical worth, and closeness to the Mind of the Creator, physics does not necessarily hold the only key to the universe.

 

REFERENCES

Barrow, John D., and Frank J. Tipler. 1986. The Anthropic Cosmological Principle. Oxford: Oxford Univ. Press.

Dirac, P.A. M. 1929. "Quantum Mechanics of Many-Electron Systems."

Proceedings of the Royal Society A, 123: 714-733.

Jackson, John David. 1975. Classical Electrodynamics. 2d ed. New York: Wiley.

Penrose, Roger. 1989. The Emperor's New Mind: Concerning Computers, Minds, and the Laws of Physics, New York: Oxford Univ. Press.

Sachs, Robert G. 1987. The Physics of Time Reversal. Chicago: Univ. of Chicago Press.

Yourgrau, Wolfgang, and Stanley Mandelstam. 1968. Variational Principles in Dynamics and Quantum Theory. Philadelphia, PA.: Saunders.

Zee, Anthony. 1986. Fearful Symmetry: The Search for Beauty in Modern Physics. New York: Macmillan.


CHEMISTRY: WHAT DOES ONE NEED TO KNOW?

by Allen R. Utke

Abstract. The general knowledge and understanding that every teacher of religion and science should have relative to chemistry can be found in the answers to three major questions. In my own response to the first question, How did chemistry emerge as a discipline? I trace the origins, establishment, and subsequent historical significance of cosmology. I contend that chemistry is "the obvious, oldest science" and, as such, has played a key role among the sciences in age long human efforts to understand reality. In my response to the second question, How do chemists currently view (cosmic) reality? I outline three prominent examples in support of my contention that chemistry, despite being "the obvious, oldest science," is seen by some as playing only a tacit role in current efforts to (re)integrate religion and science. In my response to the third question, How do chemists currently view ultimate reality and meaning? I argue that "unifiers" in chemistry can also now play a key role in a reality revolution that is pointing humankind not only toward a possible historical (re)integration of religion and science but also toward a return to cosmology.

Keywords: breakdown of cosmology; chemistry; cosmology; IRAM; reality; (re)integration of science and religion; return to cosmology; URAM.

It can be argued that we humans are conceptual as well as perceptual reality seekers and that this is the most significant difference between ourselves and all other forms of life. However, our seemingly ageless, apparently innate human need to know about reality can be viewed as reducible to two fundamental, polarized, and yet complementary conceptual questions. The first question, How does reality function? is at the heart of our universal, timeless quest to understand immediate reality and meaning, a quest that I designate IRAM. The second question, Why does reality exist? is at the heart of our complementary universal, timeless quest to understand ultimate reality and meaning, a quest I designate URAM.

It is far easier superficially to define the word reality (e.g., everything that is, the way things are, and so on) than it is to answer the two penetrating questions about reality just posed. One reason for this is that every attempt to understand reality begins with a serious misunderstanding. For, throughout history, everyone has begun with the obvious "given" perception that matter is the seen, visible, visualizable, tangible, foundational "stuff" of reality. Unfortunately, the perception that unseen, invisible, nonvisualizable energy, space, and time are also major parts of reality has usually been far less obvious to most. In fact, humankind has begun to understand the roles of energy, space, and time significantly, in a seemingly materialistic reality, only in the last two hundred years or so!

 

HOW DID CHEMISTRY EMERGE AS A DISCIPLINE?

  A scientist might be defined as someone who specializes in attempting to answer the question of how reality functions. A chemist might be defined as someone who specializes in attempting to define the role matter plays in the functioning of reality but within the context of the related roles that energy, space, and time also play. Thus, chemistry might be defined as what chemists do within science. (Note: The words chemist and chemistry are modern derivatives of the words Al Khyma, used by Arabian alchemists over a thousand years ago to denote metal working and metal workers. The words science and scientist were apparently coined by William Whewell in 1840. Before the mid-nineteenth century, scientists were usually called natural philosophers.)

Extrapolating current terms and definitions just outlined back through human history, one might ask, Who were the first "scientist" and "chemist"? Interestingly enough, today's terms and definitions argue against the presence of any "scientists" and "chemists" in human history before 10,000 B.C.E. The discoveries of weapons, clothing, utensils, fire, and the wheel were apparently pragmatic solutions to the problem of survival rather than conceptual answers to the question of how reality functions. In fact, survival was such a critical problem (as evidenced by short average life spans) in the first 4 million or so years of human history that the more prominent of the two reality questions asked by early humans was probably why reality exists. And thus it was that early religious conceptual thought (as evidenced in primitive art, ritual, and burial of the dead) apparently characterized early human history rather than early scientific conceptual thought.

From about 10,000 to 500 B.C.E., there apparently were no "scientists" and "chemists" either. It is true that a warmer global climate, coupled with an evolved brain, enabled humans in this period to alter their materialistic reality dramatically with an array of collective societal discoveries, ranging from metallurgy (the consecutive use of gold, silver, copper, tin, lead, bronze, and iron) to medicine and surgery. However, these tremendous accomplishments were pragmatic rather than theoretical in nature and thus did little to answer the question, How does reality function? But, by making life much easier than formerly in terms of survival, the accomplishments did provide human beings with increased time to think about why reality exists.

That increased time, coupled with the unique human ability to use "revealed" myths, fables, legends, and symbols to turn a limited understanding of why reality exits into formalized concepts of ultimate reality and meaning, led to the general development of religious animism, mysticism, idolatry, polytheism, monotheism, and Hinduism prior to the sixth century B.C.E. And in Confucius, Buddha, Zoroaster, Lao-Tzu, and the Jewish prophets, religious thought subsequently reached an anomalous zenith point in the sixth century B.C.E., a zenith point in terms of Judaism, which also later played a major role in the subsequent development of Christianity and Islam.

Of course, once such religious thought had been widely established through prophecy, revelation, and scripture, it usually was verified and codified through faith and ritual. And thus, prior to 500 B.C.E., once the question of why reality exists had seemingly been satisfactorily answered, the answers were invariably deductively extrapolated to also answer the complementary question of how reality functions, if, when and where that question was asked.

There was, however, one place in the world where the rapid general ascendency of religious thought in the sixth century B.C.E. was not only checked but also dramatically altered. That place was known as Ionia. The Ionians originally inhabited Attica in mainland Greece. However, in 1104 B.C.E., a nomadic, barbarous Iron Age people known as the Dorians invaded and conquered Bronze Age Greece. For the next several hundred years Greece remained under the brutal yoke of the Doriann in a reality-shattering period of time known as the Dorian Captivity. About 1000 B.C.E., the Ionians who survived the Dorian Captivity emigrated to the Asia Minor shores of the Aegean Sea, now part of Turkey. By 550 B.C.E., Miletus, the southernmost of the cities known as the "Ionian Twelve," had become an industrialized, wealthy port and center of trade (for both goods and conceptual ideas) of the known world. A highly sophisticated, urbane, educated leisure class had arisen, who were well acquainted with the most recent knowledge to come out of Egypt, Lydia, Babylonia, Phoenicia, and elsewhere.

The time was ripe for an anomalous, fresh reappraisal of the overall nature of reality. About 550 B.C.E., the Ionians became the first people in history to holistically seek both the overall mechanism of reality (how it functions) and also its underlying basis (why it exists) without relying on prior religious thought, myths, poetry, and so on for help and answers.

The Ionians sought the interrelated oneness of reality; their search was undergirded by the conviction that universal reality is a single, integrated reality system, unified and controlled by universal principles and laws, and that all things in universal reality (including humankind) purposively share in a common "good" order. Such a reality system was called the Cosmos by the Ionians, and the study of the cosmos became known as cosmology.

The Ionians, on the basis of today's definitions, developed the first science, the first philosophy, and the first natural theology, and subsequently combined them with various religious concepts to practice the first cosmology. However, they also practiced the first chemistry, by today's definition. For their first major conceptual question about how reality functions was, Is there a fundamental element or "stuff" at the heart of all matter that serves as the underlying, fundamental basis of the material cosmos?

It is generally acknowledged that Thales of Miletus was the first scientist, chemist, philosopher, natural theologian, and thus cosmologist, in history. On what is his nomination based? About 550 B.C.E., he proposed that water was the fundamental universal form of matter, he discovered magnetism and static electricity, he discovered Thales' Proposition (the earliest principle of occidental mathematics), he hypothesized that reality is permanent, and he hypothesized that God is immanent in reality!

In the next seven hundred years, Thales of Miletus was followed by such Greek cosmologists, philosopher-scientists, and philosophers as Anaximander, Anaximenes, Pythagoras, Parmenides, Heraclitus, Leucippus, Democratus, Empedocles, Anaxagoras, Diogenes, the Stoics, the Epicureans, the Sophists, Socrates, Plato, Aristotle, Ptolemy, and many others.

A detailed discussion of the reality-shattering, conceptual accomplishments of these Greek scholars, and their subsequent impact on human history, is beyond the scope of this essay. However, it would be misleading to summarize that impact broadly, as many have, by saying only that Greek thought helped both to set the historical stage and to write the script for the scientific revolution and the development of the scientific method in premodernity (before 1600-1650), the subsequent development of modern science and technology since then, and the currently unfolding play of modernity. Such an oversimplified depiction of history leaves three very important summations unsaid.

First of all, the nature of the fundamental material element, or "stuff," of reality remained a central question at the heart of Greek thought. The nomination of water as the fundamental element eventually evolved into the four elements (fire, water, earth, and air) and four qualities (hot, cold, wet, and dry) of Aristotle's and others' theories. Democritus was the first "scientist" and "chemist" in history to propose that atoms were the fundamental form of both matter and reality (atoms plus the "void"), but his atomic theory was rejected by most other Greeks. The four element and four quality theory was subsequently employed in a two-thousand-year pragmatic quest to turn the base metals into gold and silver. That quest, which overall became known as alchemy, was practiced as Eastern alchemy from about 300 B.C.E. to 600, then as Arabic alchemy from about 600 to 1200, and finally as European alchemy from about 1200 to 1600.

Second, it was Eastern and Arabic alchemy that subsequently kept early Greek "science" and cosmology alive (after the Greeks were conquered from without by the nonscientific Romans, the Romans were conquered from within by the nonscientific Christians, and the Christians passed through the Dark Ages) until they were rediscovered by Europeans in Arabic lands during the Crusades.

A subsequent medieval attempt to unite Greek thought with Christian belief became known as Scholasticism and was carried out after about 1100 in the first newly established European universities. However, the incorporation of experimentation into early Greek science by Roger Bacon, about 1250, and subsequently by others (e.g., Tycho Brahe, Johannes Kepler, Galileo, Copernicus, and Francis Bacon) created a rift and then a widening chasm between "early empirical science" and religion over the next five hundred years. But, buttressed by other major contemporary intellectual and social developments, beleaguered science gradually matured and grew in acceptance, influence, and power. Despite the fact that it lost the early conceptual reality battles with religion, it finally won the overall war about 1600-1650 and helped usher in the modern age.

Third, the evolving, accelerating, exponential success and power with which modern science increasingly has been able to answer the question of how reality functions now stands in sharp contrast with the traditional, seemingly static answers that religion and philosophy continue to give to the question of why reality exists. This increasing dualism and polarity, and resultant decreasing complementarity, has already resulted in the breakdown and fragmentation of Ionian cosmology in our age. In other words, that which the Ionians holistically put together in a premodern age that increasingly emphasized URAM has now been reductionistically rent asunder in our own modern age, which increasingly emphasizes IRAM.

In summary, I have now answered, even if briefly, the question, How did chemistry emerge as a discipline? However, beyond that, I also have attempted to superimpose and briefly defend my unusual hypothesis that chemistry is "the obvious, oldest science" and, as such, has played the longest and perhaps even the key role among all of the sciences in human attempts to understand reality. And, of course, overall, I have also attempted to "set the stage" for asking the remaining two questions, which follow.

HOW DO CHEMISTS CURRENTLY VIEW (COSMIC) REALITY?

Space limitations in this paper preclude, not only a detailed answer to this question, but even a broad overview. The story of the interrelated, accelerating growth of an increasingly specialized chemical view of reality, within the context of a similarly growing and yet increasingly fragmented modern scientific view of reality, is so extensive and so complex that by necessity it must remain largely untold here. However, three perspectives in particular illuminate current developments in scientific thought about the nature of immediate reality and meaning (IRAM).

Holism. First of all, it should be pointed out that an increasing number of scholars, from diverse backgrounds, have been contending in the latter half of the twentieth century that they have discovered a "new reality" within modern reductionistic science. In that "new reality" the universe is viewed as being a single, orderly, integrated, holistic reality system, unified and controlled by universal laws and principles, within which humankind plays a significant role. In other words, more and more scholars are beginning to realize that the Ionians were right! In support of the aforementioned scholars, and out of respect and admiration for the Ionians, I will hereafter refer to the universe as the cosmos and the "new reality" as cosmic reality.

Some scholars are even now maintaining that cosmic reality can, or even should, serve as a basis for a historical reappraisal of the current relationship between science and religion, a possible reintegration of the two, and perhaps even a "return to cosmology." Growing evidence for cosmic reality (and the associated dramatic contentions it is prompting) is currently being drawn from twentieth-century science, particularly the "new physics," biology, genetics, and the neurosciences. The "new physics" is a term applied collectively to relativity, quantum theory, and recent discoveries about the origin, nature, and functioning of the cosmos, as provided by astronomy, astrophysics, and particle physics.

Chemistry generally seems to be relegated to a supportive role in most of the evidence currently being presented for cosmic reality. That's odd, because if chemistry really is "the obvious, oldest science" as I have claimed, doesn't it also have its own unique story to tell about cosmic reality, and shouldn't it possibly have a bigger role in this regard in the evidence being presented?

Hierarchical Matter, Entropy, and Negentropy. As another facet of the largely untold story of how chemists and scientists in general interrelatedly view cosmic reality, it might be pointed out that not only have chemists always viewed matter as being the obvious major part of reality, but they currently also view it as being the only hierarchical major part or the only major part that consists of parts within parts within parts within parts....

From that unique chemical perspective, the unfolding, overall account of cosmic reality as it is currently being related by the "new physics," biology, genetics, and neurosciences takes on a chemical essence that may have been previously unclear or even overlooked. The present account, ranging from the Big Bang to the historical appearance of human beings capable of conceptually seeking cosmic reality, clearly becomes more than merely a depiction of a cosmos that continually becomes more disorderly and less informational (more entropic) in terms of energy, in accordance with the second law of thermodynamics. The account also becomes a depiction of a cosmos that, in a counter trend sort of way, also becomes more orderly and more informational (more negentropic) in terms of matter, in accordance with what might be termed the cosmic evolution of matter.

Thus, the existence of hierarchical matter (electrons, quarks, neutrons, protons, atomic nuclei, atoms, molecules, objects, and living objects) clearly reveals the presence of a complementary, evolutionary, dual directionality in the cosmos as a major answer to the question of how reality functions. Furthermore, it can then be pointed out that DNA is the most orderly informational molecule known in the cosmos (there may be as much as 100,000 volumes of information stored in the DNA of a human fertilized ovum). It also can be pointed out that the human brain is the most orderly informational object known in the cosmos (the 100 billion information-processing, conceptualizing neurons present in the human brain can be connected in more possible ways than there are atoms in the universe). And thus it can be argued that the emergence of life and human conceptual thought may actually be the ultimate purpose of the functioning of the evolving cosmos. At a minimum, the hierarchical nature of matter raises the provocative question, Why is the human brain the only object known in the cosmos that is aware of the cosmos?

Matter, Energy, and Periodicity. As a third facet of the aforementioned, largely untold story of how chemists and scientists interrelatedly view cosmic reality today, it might be pointed out that very few scholars and lay persons outside of chemistry seem currently to have a clear understanding of how matter and energy are complementary cosmic partners in the periodic law and the periodic table. In order to see why such a lack of understanding is so significant, it's necessary first to review several related segments of chemical history.

In 1649, Robert Boyle, the "skeptical chemist," dramatically refined and redirected Ionian thought and also historically sealed the coffin of alchemy by postulating that iron, sulfur, copper, and other substances are actually the fundamental elements of matter, rather than fire, water, earth, and air -- which had been considered the fundamental elements for the previous two thousand years.

In 1869, Dmitry Mendeleev discovered that every such element belongs to one of seven "families" of elements, and each family is characterized by similar physical and chemical properties. In fact, Mendeleev furthermore found that each family was divisible into an A and a B subfamily. This summarization of the nature of matter became known as the periodic law, and Mendeleev's graphic representation of the periodic law became known as the periodic table. Like the Ionians, Mendeleev never knew how all matter was related, for the underlying basis of the periodic table, the electron, wasn't discovered until the end of the nineteenth century by J. J. Thomson and others.

In 1912, Lord Rutherford conducted his famous gold foil experiment, and on this basis formulated the solar system model of the atom, in which electrons revolve around a nucleus composed of protons. In 1932, the nucleus was found also to contain neutrons. In 1913, one of Rutherford's students, Niels Bohr, in a single stroke of conceptual genius, combined all previous understanding of matter with all previous understanding of energy to propose the existence of a complementary universal pattern of quantized (set) energy levels around the nucleus of every atom in the universe. Bohr further postulated that each electron in an atom, in violation of the laws of classical physics, remains in its own place in the pattern in one of the energy levels unless it is promoted to a higher level with an injection of energy from outside the atom. However, when the electron strangely "jumps" back down to where it belongs, in what might be described as a "cosmic dance," it emits the energy it previously absorbed as light of a certain set wavelength and frequency. In other words, Bohr discovered that light originates on earth and often elsewhere through electron "jumps" in atoms!

In the 1920s, Bohr and many other investigators, notably P. A. M. Dirac, Werner Heisenberg, and Erwin Schrödinger, developed quantum mechanics, a mathematical extension of Bohr's work. In one application of quantum mechanics, one can calculate the "address" and shape of every "electron orbital" in any atom. Overall, it can then collectively be seen that Bohr's originally proposed electronic pattern actually consists of intricate energy levels within levels within levels, and thus order within order within order.

When one fits into Bohr's pattern the differing number of electrons (one to ninety-two) that each of the ninety-two natural elements in the universe has in its atoms, one sees another amazing repetitive pattern emerge. Eight families (each comprising A and B subfamilies) of elements arise, with all of the atoms of the elements in any particular family having the same number of outer-level electrons, albeit in a different outer energy level for each element. Since all of the families also are related by sequential numbers of outer-level electrons, ranging from an "extreme" of one electron to an "extreme" of eight, every element in the universe has a unique but related role to play, relative to all of the other elements. In other words, by revealing the underlying basis of the periodic table and the periodic law (discovered in 1869 by Dmitri Mendeleev), quantum mechanics actually revealed the cosmic blueprint underlying all matter in the universe.

What's the overall significance of the periodic law, the periodic table, and periodicity (the many orderly trends present in the periodic table)? I believe that it's only when one significantly understands the universal electronic energy blueprint based on light underlying all matter, as briefly outlined, that one can truly appreciate the "awesome" complementary way in which matter and energy function in the cosmos. For example, one then understands how the cosmic evolution of matter in general and the cosmic evolution of the elements takes place in the cosmos. One also then understands why carbon is the "cosmic elemental star" (the most unusual element) in the periodic table and the only element that can produce life and allow the evolution of life within the larger scheme of cosmic evolution.

An understanding of the periodic law, the periodic table, and periodicity also enables one truly to appreciate as well as simply to explain both the orderly way in which the cosmos unites atoms through the formation of chemical bonds to form molecules and the subtle chemical "tricks" the cosmos plays to bestow peculiar properties on certain molecules -- "tricks" involving hydrogen, multiple and coordinate covalent bonds, bond angles, molecular geometry, molecular polarity, and so on. For example, "tricks" involving carbon allow the formation of carbon dioxide, amino acids, proteins, and DNA and thus the formation of both terrestrial and extraterrestrial life. In this regard, carbon dioxide and DNA might be called "cosmic molecular stars."

However, water could be viewed as being the biggest "cosmic molecular star" of all, and an understanding of the periodic law, the periodic table, and periodicity enables one to understand why. In that regard, whenever and wherever two hydrogen atoms combine with one oxygen atom to form a molecule (particle) of water, the uniqueness of the resulting molecule is synergistically greater than the sum of the uniqueness of each of the two elements. The unique electron pattern in oxygen atoms dictates that water molecules are bent in terms of their two intramolecular chemical bonds rather than linear. And hydrogen's elemental uniqueness dictates that water molecules have a much greater intermolecular attractability than might be expected because of a rare type of chemical bond that only hydrogen atoms can form (with only three elements, including oxygen), known as a hydrogen bond. The remarkable story of chemical bonds, one of the four forces that hold the cosmos together, and the role of light in those bonds will unfortunately have to be omitted here because of space considerations.

Working in unison, the bent shape and unusually high attractability of water molecules produce liquid water rather than gaseous water at room temperature and also more than twenty other highly unusual properties of water, properties that, when added to those of carbon, make possible life in the cosmos.

In summary, I have been able to outline only briefly three isolated fragmentary facets of the vast, still largely untold story of how chemists and scientists currently interrelatedly view cosmic reality. However, I nonetheless hope that the reader has at least caught some significant glimpses of the important and yet often overlooked contributions that chemistry can make to the story.

 

HOW DO CHEMISTS VIEW ULTIMATE REALITY AND MEANING?

Ironically, chemistry ("the obvious, oldest science") has played a major role both in the establishment of cosmology about twenty-five hundred years ago and also in the polarization of science and religion and the breakdown and fragmentation of cosmology in our own modern age. Is it possible that the pendulum of history is now swinging back toward a (re)integration of science and religion and even a "return to cosmology"? And is it possible that chemistry could once again play a major role in such a historic swing?

The answers to these two questions depend, of course, on how a chemist (or any scientist) views the question of ultimate reality and meaning (URAM). Some chemists (and other scientists) might be called scientific diversifiers in this regard, for they largely are interested only in the polarized, reductionistic, immediate reality and meaning side (IRAM) of chemistry (or science) and thus only in the question of how reality functions. However, other chemists (and other scientists) might be called scientific unifiers, for they are also interested in the URAM side of science. They view the questions of how reality functions and why it exists as complementary polarities in the human need to know about reality.

The currently unfolding account of cosmic reality seems to be defining an increasingly mathematical, nonvisualizable, holistic, interrelated, complementary, unified, systematic, finely tuned, informational, lawlike, recursive, temporal, nonlocalized, dynamic, creative, evolving, and even mindlike cosmos in which humans play a significant role and may even be its purpose. It's a cosmos much as the Ionians generally viewed it However, if the overall holistic oneness of the cosmos increasingly is being revealed, is the oneness itself ultimate reality and meaning, or does the oneness actually point to deeper reality and meaning, a One?

Scientific diversifiers still tend to agnostically or atheistically view the cosmos just outlined as being a random, accidental, anthropic "universe." However, scientific unifiers tend to see a planned, designed, theistic, anthropic cosmos. Overall, it's an exciting time to be alive, for the scientific unifiers seem to be gaining in number and conceptually gaining historical ground. In fact, the questions of a possible (re)integration of science and religion, and even of a possible "return to cosmology," seem now to be coming into increasingly sharp focus on the frontiers of thought. And once again, chemistry seems to be playing a major role in a "reality revolution" that is pointing us toward the future. The Ionians were right!

 

REFERENCES

Anderson, Byron D., and Nathan Spielberg. 1995. Seven Ideas That Shook the Universe: New York: Wiley.

Brown, Theodore L., Bruce E. Bursten, and H. Eugene LeMay. 1991. Chemistry: The Central Science. Englewood Cliffs, N.J.: Prentice-Hall.

Clarke, R. E. D. 1961. The Universe: Plan or Accident? Exeter: Paternoster Press.

Cobb, Cathy, and Harold Goldwhite. 1995. Creations of Fire: Chemistry's Lively History from Alchemy to the Atomic Age. New York: Plenum.

Davies, Paul, and John Gribbon. 1992. The Matter Myth. New York: Simon and Schuster

Horigan, James E. 1978. Chance or Design? New York: Philosophical Library.

Life in the Universe. Special Issue. 1994. Scientifc American 271, no. 4.

Murchie, Guy. 1980. The Seven Mysteries of Life. Boston: Houghton-Mifflin


|  TOP OF PAGE  |  CONTACT JTF | SIGN UP FORM  |  SITEMAP |  TFP HOME | JTF HOME  |

Web Design by Genesis Computing, Inc. 1998