Are Human Beings Just Atoms?

In a previous essay on materialism, I discussed the bizarre nature of phenomena on the subatomic level, in which particles have no definite position in space until they are observed. Referencing the works of several physicists and philosophers, I put forth the view that reality consists not of tiny, solid objects but rather bundles of properties and qualities that emerge from potentiality to actuality. In this view, when one breaks down reality into smaller and smaller parts, one does not reach the fundamental units of matter; rather, one is gradually unbundling properties and qualities until the smallest objects no longer even have a definite position in space!

Why is this important? One reason is that the enormous prestige and accomplishments of science have sometimes led us down the wrong path in properly describing and interpreting reality. Science excels at advancing our knowledge of how things work, by breaking down wholes into component parts and manipulating those parts into better arrangements that benefit humanity. This is how we got modern medicine, computers, air conditioning, automobiles, and space travel. However, science sometimes falls short in properly describing and interpreting reality, precisely because it focuses more on the parts than the wholes.

This defect in science becomes particularly glaring when certain scientists attempt to describe what human beings are like. All too often there is a tendency to reduce humans to their component parts, whether these parts are chemical elements (atoms), chemical compounds (molecules), or the much larger molecules known as genes. However, while these component parts make up human beings, there are properties and qualities in human beings that cannot be adequately described in terms of these parts.

Marcelo Gleiser, a physicist at Dartmouth College, argues that “life is the property of a complex network of biochemical reactions . . . a kind of hungry chemistry that is able to duplicate itself.” Biologist Richard Dawkins claims that humans are “just gene machines,” and “living organisms and their bodies are best seen as machines programmed by the genes to propagate those very same genes,” though he qualifies his statement by noting that “there is a very great deal of complication, and indeed beauty in being a gene machine.” Philosopher Daniel Dennett claims that human beings are “moist robots” and the human mind is a collection of computer-like information processes which happen to take place in carbon-based rather than silicon-based hardware.

Now it is true that human beings are composed of atoms that are the basis of chemicals and molecules, that are the basis of chemical compounds, such as genes. The issue, however, is whether describing the parts that compose a human being is the same as describing the whole human being. Yes, human beings are composed of atoms of oxygen, carbon, hydrogen, nitrogen, calcium, and phosphorous. But these atoms can be found in many, many places throughout the universe, in varying quantities and combinations, and they do not have human qualities unless and until they are organized in just the right way. Likewise, genes are ubiquitous in life forms ranging from mammals to lizards to plants to bacteria. Even viruses have genes, though most scientists argue that viruses are not true life forms because they need a host to reproduce. Nevertheless, while human beings share a very few properties and qualities with bacteria and viruses, humans clearly have many properties and qualities that the lower life forms do not.

In fact, recognizing the very difference between life and death can be lost by excessive focus on atoms and molecules. Consider the following: an emergency room doctor treats a patient suffering from a heart attack. Despite the physician’s best efforts, despite all of the doctor’s training and knowledge, the patient dies on the table. So what is the difference between the patient that has died and the patient as he was several hours ago? The quantity and types of atoms composing the body are approximately the same as when the patient was alive. So what has changed? Obviously, the properties and qualities expressed by the organization of the atoms in the human being has changed. The heart no longer supplies blood to the rest of the body, the lungs no longer supply oxygen, the brain no longer has electrical activity, the human being no longer has the ability to run or walk or jump or talk or think or love. Atoms have to be organized in an extremely precise manner in order for these properties and qualities to emerge, and this organization has been lost. So if we are really going to accurately describe what a human being is, we have to refer not just to the atoms, but to the overall organization or form.

The issue of form is what separates the ancient Greek philosophers Democritus and Plato. Both philosophers believed that the universe and everything in it was composed of atoms; but Democritus thought that nothing existed but atoms and the void (space), whereas Plato believed that atoms were arranged by a creator, who, being essentially good, used ideal forms as a blueprint. Contrary to the views of Judaism, Christianity, and Islam, however, Plato believed that the creator was not omnipotent, and was forced to work with imperfect matter to do the best job possible, which is why most created objects and life forms were imperfect and fell short of the ideal forms.

Democritus would no doubt dismiss Plato’s ideal forms as being unreal — after all, forms are not something solid, so how can anything that is not solid, not made of material, exist at all? But as I’ve pointed out, the atoms that compose the human body are found everywhere, whereas actual, living human beings have these same atoms organized in a precise, particular form. In other words, in order to understand anything, it is not enough to break it down into parts and study the parts; one has to look at the whole. The properties and qualities of a living human being, as a whole, definitely do exist, or we would not know how to distinguish a living human being from a dead human being or any other existing thing composed of the same atoms.

The debate between Democritus and Plato points to a difference in ways of knowing that persist to this day: analytic knowledge and holistic knowledge. Analytic knowledge is pursued by science and reason; holistic knowledge is pursued by religion, art, and the humanities. The prestige of science and its technological accomplishments has elevated analytic understanding above all other forms of knowledge, but we remain lost without holistic understanding.

What precisely is “analytic knowledge”? The word “analyze” means “to study or determine the nature and relationship of the parts (of something) by analysis.” Synonyms for “analyze” include “break down,” “cut,” “deconstruct,” and “dissect.” In fact, the word “analysis” is derived from the New Latin word analyein, meaning “to break up.” Analysis is an extremely valuable tool and is responsible for human progress in all sorts of areas. But the knowledge derived from analysis is primarily a description and guide to how things work. It reduces knowledge of the whole to knowledge of the parts, which is fine if you want to take something apart and put it back together. But the knowledge of how things work is not the same as the knowledge of what things are as a whole, what qualities and properties they have, and the value of those qualities and properties. This latter knowledge is holistic knowledge.

The word “holism,” based on the ancient Greek word for “whole” (holos), was coined in the early twentieth century in order to promote the view that all systems, living or not, should be viewed as wholes and not just as a collection of parts or the sum of parts. It’s no accident that the words “whole,” “heal,” healthy,” and “holy” are linguistically related. The problems of sickness, malnutrition, and injury were well-known to the ancients, and it was natural for them to see these problems as a disturbance to the whole human being, rendering a person incomplete and missing certain vital functions. Wholeness was an ideal end, which made wholeness sacred (holy) as well. (For an extended discussion of analytic/reductionist knowledge vs. holistic knowledge, see this post.)

Holistic knowledge is not just about ideal physical health. It’s about ideal forms in all aspects, including the qualities we associate with human beings we admire: wisdom, strength, beauty, courage, love, kindness. As mistaken as religions have been in understanding natural causation, it is the devotion to ideal forms that is really the essence of religion. The ancient Greeks worshipped excellence, as embodied in their gods; Confucians were devoted to family ties and duties; the Jews submitted themselves to the laws of the one God; Christians devoted themselves to the love of God, embodied in Christ.

Holistic knowledge provides no guidance as to how to conduct surgery or build a computer or launch a rocket; but it does provide insight into the ethics of medicine, the desirability or hazards of certain types of technology, and the proper ends of human beings. All too often, contemporary secular societies expect new technologies to improve human lives and pay no heed to ideal human forms, on the assumption that ideal forms are a fantasy. Then we are shocked when the new technologies are abused and not only bring out the worst in human nature but enhance the power of the worst.

What Does Science Explain? Part 5 – The Ghostly Forms of Physics

The sciences do not try to explain, they hardly even try to interpret, they mainly make models. By a model is meant a mathematical construct which, with the addition of certain verbal interpretations, describes observed phenomena. The justification of such a mathematical construct is solely and precisely that it is expected to work — that is, correctly to describe phenomena from a reasonably wide area. Furthermore, it must satisfy certain esthetic criteria — that is, in relation to how much it describes, it must be rather simple. — John von Neumann (“Method in the Physical Sciences,” in The Unity of Knowledge, 1955)

Now we come to the final part of our series of posts, “What Does Science Explain?” (If you have not already, you can peruse parts 1, 2, 3, and 4 here). As I mentioned in my previous posts, the rise of modern science was accompanied by a change in humanity’s view of metaphysics, that is, our theory of existence. Medieval metaphysics, largely influenced by ancient philosophers, saw human beings as the center or summit of creation; furthermore, medieval metaphysics proposed a sophisticated, multifaceted view of causation. Modern scientists, however, rejected much of medieval metaphysics as subjective and saw reality as consisting mainly of objects impacting or influencing each other in mathematical patterns.  (See The Metaphysical Foundations of Modern Science by E.A. Burtt.)

I have already critically examined certain aspects of the metaphysics of modern science in parts 3 and 4. For part 5, I wish to look more closely at the role of Forms in causation — what Aristotle called “formal causation.” This theory of causation was strongly influenced by Aristotle’s predecessor Plato and his Theory of Forms. What is Plato’s “Theory of Forms”? In brief, Plato argued that the world we see around us — including all people, trees, and animals, stars, planets and other objects — is not the true reality. The world and the things in it are imperfect and perishable realizations of perfect forms that are eternal, and that continually give birth to the things we see. That is, forms are the eternal blueprints of perfection which the material world imperfectly represents. True philosophers do not focus on the material world as it is, but on the forms that material things imperfectly reflect. In order to judge a sculpture, painting, or natural setting, a person must have an inner sense of beauty. In order to evaluate the health of a particular human body, a doctor must have an idea of what a perfectly healthy human form is. In order to evaluate a government’s system of justice, a citizen must have an idea about what perfect justice would look like. In order to critically judge leaders, citizens must have a notion of the virtues that such a leader should have, such as wisdom, honesty, and courage.  Ultimately, according to Plato, a wise human being must learn and know the perfect forms behind the imperfect things we see: we must know the Form of Beauty, the Form of Justice, the Form of Wisdom, and the ultimate form, the Form of Goodness, from which all other forms flow.

Unsurprisingly, many intelligent people in the modern world regard Plato’s Theory of Forms as dubious or even outrageous. Modern science teaches us that sure knowledge can only be obtained by observation and testing of real things, but Plato tells us that our senses are deceptive, that the true reality is hidden behind what we sense. How can we possibly confirm that the forms are real? Even Plato’s student Aristotle had problems with the Theory of Forms and argued that while the forms were real, they did not really exist until they were manifested in material things.

However, there is one important sense in which modern science retained the notion of formal causation, and that is in mathematics. In other words, most scientists have rejected Plato’s Theory of Forms in all aspects except for Plato’s view of mathematics. “Mathematical Platonism,” as it is called, is the idea that mathematical forms are objectively real and are part of the intrinsic order of the universe. However, there are also sharp disagreements on this subject, with some mathematicians and scientists arguing that mathematical forms are actually creations of the human imagination.

The chief difference between Plato and modern scientists on the study of mathematics is this: According to Plato, the objects of geometry — perfect squares, perfect circles, perfect planes — existed nowhere in the material world; we only see imperfect realizations. But the truly wise studied the perfect, eternal forms of geometry rather than their imperfect realizations. Therefore, while astronomical observations indicated that planetary bodies orbited in imperfect circles, with some irregularities and errors, Plato argued that philosophers must study the perfect forms instead of the actual orbits! (The Republic, XXVI, 524D-530C) Modern science, on the other hand, is committed to observation and study of real orbits as well as the study of perfect mathematical forms.

Is it tenable to hold the belief that Plato and Aristotle’s view of eternal forms is mostly subjective nonsense, but they were absolutely right about mathematical forms being real? I argue that this selective borrowing of the ancient Greeks doesn’t quite work, that some of the questions and difficulties with proving the reality of Platonic forms also afflicts mathematical forms.

The main argument for mathematical Platonism is that mathematics is absolutely necessary for science: mathematics is the basis for the most important and valuable physical laws (which are usually in the form of equations), and everyone who accepts science must agree that the laws of nature or the laws of physics exist. However, the counterargument to this claim is that while mathematics is necessary for human beings to conduct science and understand reality, that does not mean that mathematical objects or even the laws of nature exist objectively, that is, outside of human minds.

I have discussed some of the mysterious qualities of the “laws of nature” in previous posts (here and here). It is worth pointing out that there remains a serious debate among philosophers as to whether the laws of nature are (a) descriptions of causal regularities which help us to predict or (b) causal forces in themselves. This is an important distinction that most people, including scientists, don’t notice, although the theoretical consequences are enormous. Physicist Kip Thorne writes that laws “force the Universe to behave the way it does.” But if laws have that kind of power, they must be ubiquitous (exist everywhere), eternal (exist prior to the universe), and have enormous powers although they have no detectable energy or mass — in other words, the laws of nature constitute some kind of supernatural spirit. On the other hand, if laws are summary descriptions of causation, these difficulties can be avoided — but then the issue arises: do the laws of nature or of physics really exist objectively, outside of human minds, or are they simply human-constructed statements about patterns of causation? There are good reasons to believe the latter is true.

The first thing that needs to be said is that nearly all these so-called laws of nature are actually approximations of what really happens in nature, approximations that work only under certain restrictive conditions. Both of these considerations must be taken into account, because even the approximations fall apart outside of certain pre-specified conditions. Newton’s law of universal gravitation, for example, is not really universal. It becomes increasingly inaccurate under conditions of high gravity and very high velocities, and at the atomic level, gravity is completely swamped by other forces. Whether one uses Newton’s law depends on the specific conditions and the level of accuracy one requires. Kepler’s laws of planetary motion are an approximation based on the simplifying assumption of a planetary system consisting of one planet. The ideal gas law is an approximation which becomes inaccurate under conditions of low temperature and/or high pressure. The law of multiple proportions works for simple molecular compounds, but often fails for complex molecular compounds. Biologists have discovered so many exceptions to Mendel’s laws of genetics that some believe that Mendel’s laws should not even be considered laws.

The fact of the matter is that even with the best laws that science has come up with, we still can’t predict the motions of more than two interacting astronomical bodies without making unrealistic simplifying assumptions. Michael Scriven, a mathematician and philosopher at Claremont Graduate University, has concluded that the laws of nature or physics are actually cobbled together by scientists based on multiple criteria:

Briefly we may say that typical physical laws express a relationship between quantities or a property of systems which is the simplest useful approximation to the true physical behavior and which appears to be theoretically tractable. “Simplest” is vague in many cases, but clear for the extreme cases which provide its only use. “Useful” is a function of accuracy and range and purpose. (Michael Scriven, “The Key Property of Physical Laws — Inaccuracy,” in Current Issues in the Philosophy of Science, ed. Herbert Feigl)

The response to this argument is that it doesn’t disprove the objective existence of physical laws — it simply means that the laws that scientists come up with are approximations to real, objectively existing underlying laws. But if that is the case, why don’t scientists simply state what the true laws are? Because the “laws” would actually end up being extremely long and complex statements of causation, with so many conditions and exceptions that they would not really be considered laws.

An additional counterargument to mathematical Platonism is that while mathematics is necessary for science, it is not necessary for the universe. This is another important distinction that many people overlook. Understanding how things work often requires mathematics, but that doesn’t mean the things in themselves require mathematics. The study of geometry has given us pi and the Pythagorean theorem, but a child does not need to know these things in order to draw a circle or a right triangle. Circles and right triangles can exist without anyone, including the universe, knowing the value of pi or the Pythagorean theorem. Calculus was invented in order to understand change and acceleration; but an asteroid, a bird, or a cheetah is perfectly capable of changing direction or accelerating without needing to know calculus.

Even among mathematicians and scientists, there is a significant minority who have argued that mathematical objects are actually creations of the human imagination, that math may be used to model aspects of reality, but it does not necessarily do so. Mathematicians Philip J. Davis and Reuben Hersh argue that mathematics is the study of “true facts about imaginary objects.” Derek Abbot, a professor of engineering, writes that engineers tend to reject mathematical Platonism: “the engineer is well acquainted with the art of approximation. An engineer is trained to be aware of the frailty of each model and its limits when it breaks down. . . . An engineer . . . has no difficulty in seeing that there is no such a thing as a perfect circle anywhere in the physical universe, and thus pi is merely a useful mental construct.” (“The Reasonable Ineffectiveness of Mathematics“) Einstein himself, making a distinction between mathematical objects used as models and pure mathematics, wrote that “As far as the laws of mathematics refer to reality, they are not certain; and as far as they are certain, they do not refer to reality.” Hartry Field, a philosopher at New York University, has argued that mathematics is a useful fiction that may not even be necessary for science. Field goes to show that it is possible to reconstruct Newton’s theory of gravity without using mathematics. (There is more discussion on this subject here and here.)

So what can we conclude about the existence of forms? I have to admit that although I’m skeptical, I have no sure conclusions. It seems unlikely that forms exist outside the mind . . . but I can’t prove they don’t exist either. Forms do seem to be necessary for human reasoning — no thinking human can do without them. And forms seem to be rooted in reality: perfect circles, perfect squares, and perfect human forms can be thought of as imaginative projections of things we see, unlike Sherlock Holmes or fire-breathing dragons or flying spaghetti monsters, which are more creatively fictitious. Perhaps one could reconcile these opposing views on forms by positing that the human mind and imagination is part of the universe itself, and that the universe is becoming increasingly consciously aware.

Another way to think about this issue was offered by Robert Pirsig in Zen and the Art of Motorcycle Maintenance. According to Pirsig, Plato made a mistake by positing Goodness as a form. Even considered as the highest form, Goodness (or “Quality,” in Pirsig’s terminology) can’t really be thought of as a static thing floating around in space or some otherworldly realm. Forms are conceptual creations of humans who are responding to Goodness (Quality). Goodness itself is not a form, because it is not an unchanging thing — it is not static or even definable. It is “reality itself, ever changing, ultimately unknowable in any kind of fixed, rigid way.” (p. 342) Once we let go of the idea that Goodness or Quality is a form, we can realize that not only is Goodness part of reality, it is reality.

As conceptual creations, ideal forms are found in both science and religion. So why, then, does there seem to be such a sharp split between science and religion as modes of knowledge? I think it comes down to this: science creates ideal forms in order to model and predict physical phenomena, while religion creates ideal forms in order to provide guidance on how we should live.

Scientists like to see how things work — they study the parts in order to understand how the wholes work. To increase their understanding, scientists may break down certain parts into smaller parts, and those parts into even smaller parts, until they come to the most fundamental, indivisible parts. Mathematics has been extremely useful in modeling and understanding these parts of nature, so scientists create and appreciate mathematical forms.

Religion, on the other hand, tends to focus on larger wholes. The imaginative element of religion envisions perfect states of being, whether it be the Garden of Eden or the Kingdom of Heaven, as well as perfect (or near perfect) humans who serve as prophets or guides to a better life. Religion is less concerned with how things work than with how things ought to work, how things ought to be. So religion will tend to focus on subjects not covered by science, including the nature and meaning of beauty, love, and justice. There will always be debates about the appropriateness of particular forms in particular circumstances, but the use of forms in both science and religion is essential to understanding the universe and our place in it.

What Does Science Explain? Part 4 – The Ends of the Universe

Continuing my series of posts on “What Does Science Explain?” (parts 1, 2 , and 3 here), I wish today to discuss the role of teleological causation. Aristotle referred to teleology in his discussion of four causes as “final causation,” because it referred to the goals or ends of all things (the Greek word “telos” meaning “goal,” “purpose,” or “end.”) From a teleological viewpoint, an acorn grows into an oak tree, a bird takes flight, and a sculptor creates statues because these are the inherent and intended ends of the acorn, bird, and sculptor. Medieval metaphysics granted a large role for teleological causation in its view of the universe.

According to E.A. Burtt in The Metaphysics of Modern Science, the growth of modern science changed the idea of causation, focusing almost exclusively on efficient causation (objects impacting or affecting other objects). The idea of final (goal-oriented) causation was dismissed. And even though the early modern scientists such as Galileo and Newton believed in God, their notion of God was significantly different from the traditional medieval conception of God. Rather than seeing God as the Supreme Good, which continually draws all things to higher levels of being, early modern scientists reduced God to the First Efficient Cause, who merely started the mechanism of the universe and then let it run.

It was not unreasonable for early scientists to focus on efficient causation rather than final causation. It was often difficult to come up with testable hypotheses and workable predictive models by assuming long-term goals in nature. There was always a strong element of mystery about what the true ends of nature were and it was very difficult to pin down these alleged goals. Descartes believed in God, but also wrote that it was impossible to know what God’s goals were. For that reason, it is quite likely that science in its early stages needed to overcome medieval metaphysics in order to make its first great discoveries about nature. Focusing on efficient causation was simpler and apt to bring quicker results.

However, now that science has advanced over the centuries, it is worth revisiting the notion of teleological causation as a means of filling in gaps in our current understanding of nature. It is true that the concept of long-term goals for physical objects and forces often does not help very much in terms of developing useful, short-term predictive models. But final causation can help make sense of long-term patterns which may not be apparent when making observations over short periods of time. Processes that look purposeless and random in the short-term may actually be purposive in the long-term. We know that an acorn under the right conditions will eventually become an oak tree, because the process and the outcome of development can be observed within a reasonable period of time and that knowledge has been passed on to us. If our knowledge base began at zero and we came across an acorn for the first time, we would find it extremely difficult to predict the long-term future of that acorn merely by cutting it up and examining it under a microscope.

So, does the universe have long-term, goal-oriented patterns that may be hidden among the short-term realities of contingency and randomness? A number of physicists began to speculate that this was the case in the late twentieth century, when their research indicated that the physical forces and constants of the universe can exist in only a very narrow range of possibilities in order for life to be possible, or even for the universe to exist. Change in even one of the forces or constants could make life impossible or cause the universe to self-destruct in a short period of time. In this view, the evolution of the universe and of life on earth has been subject to a great deal of randomness, but the cosmic structure and conditions that made evolution possible are not at all random. As the physicist Freeman Dyson has noted:

It is true that we emerged in the universe by chance, but the idea of chance is itself only a cover for our ignorance. . . . The more I examine the universe and study the details of its architecture, the more evidence I find that the universe in some sense must have known that we were coming. (Disturbing the Universe, p. 250)

In what way did the universe “know we were coming?” Consider the fact that in the early universe after the Big Bang, the only elements that existed were the “light” elements hydrogen and helium, along with trace amounts of lithium and beryllium. A universe with only four elements would certainly be simple, but there would not be much to build upon. Life, at least as we know it, requires not just hydrogen but at a minimum carbon, oxygen, nitrogen, phosphorus, and sulfur. How did these and other heavier elements come into being? Stars produced them, through the process of fusion. In fact, stars have been referred to as the “factories” of heavy elements. Human beings today consist primarily of oxygen, followed by carbon, hydrogen, nitrogen, calcium, and phosphorous. Additional elements compose less than one percent of the human body, but even most of these elements are essential to human life. Without the elements produced earlier by stars we would not be here. It has been aptly said that human beings are made of “stardust.”

So why did stars create the heavier elements? After all, the universe could have gotten along quite well without additional elements. Was it random chance that created the heavy elements? Not really. Random chance plays a role in many natural events, but the creation of heavy elements in stars requires some precise conditions — it is not just a churning jumble of subatomic particles. The astronomer Fred Hoyle was the first scientist to study how stars made heavy elements, and he noted that the creation of heavy elements required very specific values in order for the process to work. When he concluded his research Hoyle remarked, “A common sense interpretation of the facts suggests that a superintellect has monkeyed with physics, as well as with chemistry and biology, and that there are no blind forces worth speaking about in nature. The numbers one calculates from the facts seem to me so overwhelming as to put this conclusion almost beyond question.”

The creation of heavier elements by the stars does not necessarily mean that the universe intended specifically to create human beings, but it does seem to indicate that the universe somehow “knew” that heavy elements would be required to create higher forms of being, above and beyond the simple and primitive elements created by the Big Bang. In that sense, creating life is plausibly a long-term goal of the universe.

And what about life itself? Does it make sense to use teleology to study the behavior of life forms? Biologist Peter Corning has argued that while science has long pursued reductionist explanations of phenomena, it is impossible to really know biological systems without pursuing holistic explanations centered on the purposive behavior of organisms.

According to reductionism, all things can be explained by the parts that they are made of — human beings are made of tissues and organs, which are made of cells, which are made of chemical compounds, which are made of atoms, which are made of subatomic particles. In the view of many scientists, everything about human beings can in principle be explained by actions at the subatomic level. Peter Corning, however, argues that this conception is mistaken. Reductionism is necessary for partially explaining biological systems, but it is not sufficient. The reason for this is that the wholes are greater than the parts, and the behavior of wholes often has characteristics that are radically different from the parts that they are made of. For example, it would be dangerous to add pure hydrogen or oxygen to a fire, but when hydrogen atoms and oxygen atoms are combined in the right way — as H2O — one obtains a chemical compound that is quite useful for extinguishing fires. The characteristics of the molecule are different from the characteristics of the atoms in it. Likewise, at the subatomic level, particles may have no definite position in space and can even be said to exist in multiple places at once; but human beings only exist in one place at a time, despite the fact that human beings are made of subatomic particles. The behavior of the whole is different from the behavior of the parts. The transformation of properties that occurs when parts form new wholes is known as “emergence.”

Corning notes that when one incorporates analysis of wholes into theoretical explanation, there is goal-oriented “downward causation” as well as “upward causation.” For example, a bird seeks the goal of food and a favorable environment, so when it begins to get cold, that bird flies thousands of miles to a warmer location for the winter. The atoms that make up that bird obviously go along for the ride, but a scientist can’t use the properties of the atoms to predict the flight of these atoms; only by looking at the properties of the bird as a whole can a scientist predict what the atoms making up the bird are going to do. The bird as a whole doesn’t have complete control over the atoms composing its body, but it clearly has some control. Causation goes down as well as up. Likewise, neuropsychologist Roger Sperry has argued that human consciousness is a whole that influences the parts of the brain and body just as the parts of the brain and body influence the consciousness: “[W]e contend that conscious or mental phenomena are dynamic, emergent, pattern (or configurational) properties of the living brain in action . . . these emergent pattern properties in the brain have causal control potency. . . ” (“Mind, Brain, and Humanist Values,” Bulletin of the Atomic Scientists, Sept 1966) In Sperry’s view, the values created by the human mind influence human behavior as much as the atoms and chemicals in the human body and brain.

Science has traditionally viewed the evolution of the universe as upward causation only, with smaller parts joining into larger wholes as a result of the laws of nature and random chance. This view of causation is illustrated in the following diagram:

reductionism

But if we take seriously the notion of emergence and purposive action, we have a more complex picture, in which the laws of nature and random chance constrain purposive action and life forms, but do not entirely determine the actions of life forms — i.e., there is both upward and downward causation:

reductionism_and_holism

It is important to note that this new view of causation does not eliminate the laws of nature — it just sets limits on what the laws of nature can explain. Specifically, the laws of nature have their greatest predictive power when we are dealing with the simplest physical phenomena; the complex wholes that are formed by the evolutionary process are less predictable because they can to some extent work around the laws of nature by employing the new properties that emerge from the joining of parts. For example, it is relatively easy to predict the motion of objects in the solar system by using the laws of nature; it is not so easy to predict the motion of life forms because life forms have properties that go beyond the simple properties possessed by objects in the solar system. As Robert Pirsig notes in Lila, life can practically be defined by its ability to transcend or work around the static patterns of the laws of nature:

The law of gravity . . . is perhaps the most ruthlessly static pattern of order in the universe. So, correspondingly, there is no single living thing that does not thumb its nose at that law day in and day out. One could almost define life as the organized disobedience of the law of gravity. One could show that the degree to which an organism disobeys this law is a measure of its degree of evolution. Thus, while the single protozoa just barely get around on their cilia, earthworms manage to control their distance and direction, birds fly into the sky, and man goes all the way to the moon. (Lila (1991), p. 143.

Many scientists still resist the notion of teleological causation. But it could be argued that even scientists who vigorously deny that there is any purpose in the universe actually have an implicit teleology. Their teleology is simply the “laws of nature” themselves, and either the inner goal of all things is to follow those laws, or it is the goal of the laws to compel all things to follow their commands. Other implicit teleologies can be found in scientists’ assumptions that nature is inherently simple; that mathematics is the language of nature; or that all the particles and forces in the nature play some necessary role. According to physicist Paul Davies,

There is . . . an unstated but more or less universal feeling among physicists that everything that exists in nature must have a ‘place’ or a role as part of some wider scheme, that nature should not indulge in profligacy by manifesting gratuitous entities, that nature should not be arbitrary. Each facet of physical reality should link in with the others in a ‘natural’ and logical way. Thus, when the particle known as the muon was discovered in 1937, the physicist Isidor Rabi was astonished. ‘Who ordered that?’ he exclaimed. (Paul Davies, The Mind of God: The Scientific Basis for a Rational World, pp. 209-10.

Ultimately, however, one cannot fully discuss the goals or ends of the universe without exploring the notion of Ideal Forms — that is, a blueprint for all things to follow or aspire to. The subject of Ideal Forms will be discussed in my next post.

What Does Science Explain? Part 3 – The Mythos of Objectivity

In parts one and two of my series “What Does Science Explain?,” I contrasted the metaphysics of the medieval world with the metaphysics of modern science. The metaphysics of modern science, developed by Kepler, Galileo, Descartes, and Newton, asserted that the only true reality was mathematics and the shape, motion, and solidity of objects, all else being subjective sensations existing solely within the human mind. I pointed out that the new scientific view was valuable in developing excellent predictive models, but that scientists made a mistake in elevating a method into a metaphysics, and that the limitations of the metaphysics of modern science called for a rethinking of the modern scientific worldview. (See The Metaphysical Foundations of Modern Science by Edwin Arthur Burtt.)

Early scientists rejected the medieval worldview that saw human beings as the center and summit of creation, and this rejection was correct with regard to astronomical observations of the position and movement of the earth. But the complete rejection of medieval metaphysics with regard to the role of humanity in the universe led to a strange division between theory and practice in science that endures to this day. The value and prestige of science rests in good part on its technological achievements in improving human life. But technology has a two-sided nature, a destructive side as well as a creative side. Aspects of this destructive side include automatic weaponry, missiles, conventional explosives, nuclear weapons, biological weapons, dangerous methods of climate engineering, perhaps even a threat from artificial intelligence. Even granting the necessity of the tools of violence for deterrence and self-defense, there remains the question of whether this destructive technology is going too far and slipping out of our control. So far the benefits of good technology have outweighed the hazards of destructive technology, but what research guidance is offered to scientists when human beings are removed from their high place in the universe and human values are separated from the “real” world of impersonal objects?

Consider the following question: Why do medical scientists focus their research on the treatment and cure of illness in humans rather than the treatment and cure of illness in cockroaches or lizards? This may seem like a silly question, but there’s no purely objective, scientific reason to prefer one course of research over another; the metaphysics of modern science has already disregarded the medieval view that humans have a privileged status in the universe. One could respond by arguing that human beings have a common self-interest in advancing human health through medical research, and this self-interest is enough. But what is the scientific justification for the pursuit of self-interest, which is not objective anyway? Without a recognition of the superior value of human life, medical science has no research guidance.

Or consider this: right now, astronomers are developing and employing advanced technologies to detect other worlds in the galaxy that may have life. The question of life on other planets has long interested astronomers, but it was impossible with older technologies to adequately search for life. It would be safe to say that the discovery of life on another planet would be a landmark development in science, and the discovery of intelligent life on another planet would be an astonishing development. The first scientist who discovered a world with intelligent life would surely win awards and fame. And yet, we already have intelligent life on earth and the metaphysics of modern science devalues it. In practice, of course, most scientists do value human life; the point is, the metaphysics behind science doesn’t, leaving scientists at a loss for providing an intellectual justification for a research program that protects and advances human life.

A second limitation of modern science’s metaphysics, closely related to the first, is its disregard of certain human sensations in acquiring knowledge. Early scientists promoted the view that only the “primary qualities” of mathematics, shape, size, and motion were real, while the “secondary qualities” of color, taste, smell, and sound existed only in the mind. This distinction between primary and secondary qualities was criticized at the time by philosophers such as George Berkeley, a bishop of the Anglican Church. Berkeley argued that the distinction between primary and secondary qualities was false and that even size, shape, and motion were relative to the perceptions and judgment of observers. Berkeley also opposed Isaac Newton’s theory that space and time were absolute entities, arguing instead that these were ideas rooted in human sensations. But Berkeley was disregarded by scientists, largely because Newton offered predictive models of great value.

Three hundred years later, Isaac Newton’s models retain their great value and are still widely used — but it is worth noting that Berkeley’s metaphysics has actually proved superior in many respects to Newton’s metaphysics.

Consider the nature of mathematics. For many centuries mathematicians believed that mathematical objects were objectively real and certain and that Euclidean geometry was the one true geometry. However, the discovery of non-Euclidean geometries in the nineteenth century shook this assumption, and mathematicians had to reconcile themselves to the fact that it was possible to create multiple geometries of equal validity. There were differences between the geometries in terms of their simplicity and their ability to solve particular problems, but no one geometry was more “real” than the others.

If you think about it, this should not be surprising. The basic objects of geometry — points, lines, and planes — aren’t floating around in space waiting for you to take note of them. They are concepts, creations of the human brain. We may see particular objects that resemble points, lines, and planes, but space itself has no visible content; we have to add content to it.  And we have a choice in what content to use. It is possible to create a geometry in which all lines are straight or all lines are curved; in which some lines are parallel or no lines are parallel;  or in which lines are parallel over a finite distance but eventually meet at some infinitely great distance. It is also possible to create a geometry with axioms that assume no lines, only points; or a geometry that assumes “regions” rather than points. So the notion that mathematics is a “primary quality” that exists within objects independent of human minds is a myth. (For more on the imaginary qualities of mathematics, see my previous posts here and here.)

But aside from the discovery of multiple mathematical systems, what has really killed the artificial distinction between “primary qualities,” allegedly objective, and “secondary qualities,” allegedly subjective, is modern science itself, particularly in the findings of relativity theory and quantum mechanics.

According to relativity theory, there is no single, objectively real size, shape, or motion of objects — these qualities are all relative to an observer in a particular reference frame (say, at the same location on earth, in the same vehicle, or in the same rocket ship). Contrary to some excessive and simplistic views, relativity theory does NOT mean that any and all opinions are equally valid. In fact, all observers within the same reference frame should be seeing the same thing and their measurements should match. But observers in different reference frames may have radically different measurements of the size, shape, and motion of an object, and there is no one single reference frame that is privileged — they are all equally valid.

Consider the question of motion. How fast are you moving right now? Relative to your computer or chair, you are probably still. But the earth is rotating at 1040 miles per hour, so relative to an observer on the moon, you would be moving at that speed — adjusting for the fact that the moon is also orbiting around the earth at 2288 miles per hour. But also note that the earth is orbiting the sun at 66,000 miles per hour, our solar system is orbiting the galaxy at 52,000 miles per hour, and our galaxy is moving at 1,200,000 miles per hour; so from the standpoint of an observer in another galaxy you are moving at a fantastically fast speed in a series of crazy looping motions. Isaac Newton argued that there was an absolute position in space by which your true, objective speed could be measured. But Einstein dismissed that view, and the scientific consensus today is that Einstein was right — the answer to the question of how fast you are moving is relative to the location and speed of the observer.

The relativity of motion was anticipated by the aforementioned George Berkeley as early as the eighteenth century, in his Treatise Concerning the Principles of Human Knowledge (paragraphs 112-16). Berkeley’s work was subsequently read by the physicist Ernest Mach, who subsequently influenced Einstein.

Relativity theory also tells us that there is no absolute size and shape, that these also vary according to the frame of reference of an observer in relation to what is observed. An object moving at very fast speeds relative to an observer will be shortened in length, which also affects its shape. (See the examples here and here.) What is the “real” size and shape of the object? There is none — you have to specify the reference frame in order to get an answer. Professor Richard Wolfson, a physicist at Middlebury College who has a great lecture series on relativity theory, explains what happens at very fast speeds:

An example in which length contraction is important is the Stanford Linear Accelerator, which is 2 miles long as measured on Earth, but only about 3 feet long to the electrons moving down the accelerator at 0.9999995c [nearly the speed of light]. . . . [Is] the length of the Stanford Linear Accelerator ‘really’ 2 miles? No! To claim so is to give special status to one frame of reference, and that is precisely what relativity precludes. (Course Guidebook to Einstein’s Relativity and the Quantum Revolution, Lecture 10.)

In fact, from the perspective of a light particle (a photon), there is infinite length contraction — there is no distance and the entire universe looks like a point!

The final nail in the coffin of the metaphysics of modern science is surely the weird world of quantum physics. According to quantum physics, particles at the subatomic level do not occupy only one position at a particular moment of time but can exist in multiple positions at the same time — only when the subatomic particles are observed do the various possibilities “collapse” into a single outcome. This oddity led to the paradoxical thought experiment known as “Schrodinger’s Cat” (video here). The importance of the “observer effect” to modern physics is so great that some physicists, such as the late physicist John Wheeler, believed that human observation actually plays a role in shaping the very reality of the universe! Stephen Hawking holds a similar view, arguing that our observation “collapses” multiple possibilities into a single history of the universe: “We create history by our observation, rather than history creating us.” (See The Grand Design, pp. 82-83, 139-41.) There are serious disputes among scientists about whether uncertainties at the subatomic level really justify the multiverse theories of Wheeler and Hawking, but that is another story.

Nevertheless, despite the obsolescence of the metaphysical premises of modern science, when scientists talk about the methods of science, they still distinguish between the reality of objects and the unreality of what exists in the mind, and emphasize the importance of being objective at all times. Why is that? Why do scientists still use a metaphysics developed centuries ago by Kepler, Galileo, and Newton? I think this practice persists largely because the growth of knowledge since these early thinkers has led to overspecialization — if one is interested in science, one pursues a degree in chemistry, biology, or physics; if one is interested in metaphysics, one pursues a degree in philosophy. Scientists generally aren’t interested in or can’t understand what philosophers have to say, and philosophers have the same view of scientists. So science carries on with a metaphysics that is hundreds of years old and obsolete.

It’s true that the idea of objectivity was developed in response to the very real problem of the uncertainty of human sense impressions and the fallibility of the conclusions our minds draw in response to those sense impressions. Sometimes we think we see something, but we don’t. People make mistakes, they may see mirages; in extreme cases, they may hallucinate. Or we see the same thing but have different interpretations. Early scientists tried to solve this problem by separating human senses and the human mind from the “real” world of objects. But this view was philosophically dubious to begin with and has been refuted by science itself. So how do we resolve the problem of mistaken and differing perceptions and interpretations?

Well, we supplement our limited senses and minds with the senses and minds of other human beings. We gather together, we learn what others have perceived and concluded, we engage in dialogue and debate, we conduct repeated observations and check our results with the results of others. If we come to an agreement, then we have a tentative conclusion; if we don’t agree, more observation, testing, and dialogue is required to develop a picture that resolves the competing claims. In some cases we may simply end up with an explanation that accounts for why we come up with different conclusions — perhaps we are in different locations, moving at different speeds, or there is something about our sensory apparatus that causes us to sense differently. (There is an extensive literature in science about why people see colors differently due to the nature of the eye and brain.)

Central to the whole process of science is a common effort — but there is also the necessity of subduing one’s ego, acknowledging that not only are there other people smarter than we are, but that the collective efforts of even less-smart people are greater than our own individual efforts. Subduing one’s ego is also required in order to prepare for the necessity of changing one’s mind in response to new evidence and arguments. Ultimately, the search for knowledge is a social and moral enterprise. But we are not going to succeed in that endeavor by positing a reality separate from human beings and composed only of objects. (Next: Part 4)

What Does Science Explain? Part 2 – The Metaphysics of Modern Science

In my previous post, I discussed the nature of metaphysics, a theory of being and existence, in the medieval world. The metaphysics of the medieval period was strongly influenced by the ancient Greeks, particularly Aristotle, who posited four causes or explanations for why things were. In addition, Aristotle argued that existence could be understood as the result of a transition from “potentiality” to “actuality.” With the rise of modern science, argued Edwin Arthur Burtt in The Metaphysical Foundations of Modern Science, the medieval conception of existence changed. Although some of this change was beneficial, argued Burtt, there was also a loss.

The first major change that modern science brought about was the strict separation of human beings, along with human senses and desires, from the “real” universe of impersonal objects joining, separating, and colliding with each other. Rather than seeing human beings as the center or summit of creation, as the medievals did, modern scientists removed the privileged position of human beings and promoted the goal of “objectivity” in their studies, arguing that we needed to dismiss all subjective human sensations and look at objects as they were in themselves. Kepler, Galileo, and Newton made a sharp distinction between the “primary qualities” of objects and “secondary qualities,” arguing that only primary qualities were truly real, and therefore worth studying. What were the “primary qualities?”: quantity/mathematics, motion, shape, and solidity. These qualities existed within objects and were independent of human perception and sensation. The “secondary qualities” were color, taste, smell, and sound; these were subjective because they were derived from human sensations, and therefore did not provide objective facts that could advance knowledge.

The second major change that modern science brought to metaphysics was a dismissal of the medieval world’s rich and multifaceted concept of causation in favor of a focus on “efficient causation” (the impact of one object or event on another). The concept of “final causation,” that is, goal-oriented development, was neglected. In addition, the concept of “formal causation,” that is, the emergence of things out of universal forms, was reduced to mathematics; only mathematical forms expressed in the “laws of nature,” were truly real, according to the new scientific worldview. Thus, all causation was reduced to mathematical “laws of nature” directing the motion and interaction of objects.

The consequences of this new worldview were tremendous in terms of altering humanity’s conception of reality and what it meant to explain reality. According to Burtt, “From now on, it is a settled assumption for modern thought in practically every field, that to explain anything is to reduce it to its elementary parts, whose relations, where temporal in character, are conceived in terms of efficient causality solely.” (Metaphysics of Modern Science, p. 134) And although the early giants of science — Kepler, Galileo, and Newton — believed in God, their conception of God was significantly different from the medieval view. Rather than seeing God as the Supreme Good, the goal or end which continually brought all things from potentiality to actuality, they saw God in terms of the “First Efficient Cause” only. That is, God brought the laws of nature into existence, and then the universe operated like a clock or machine, which might then only occasionally need rewinding or maintenance. But once this conception of God became widespread, it was not long before people questioned whether God was necessary at all to explain the universe.

Inarguably, there were great advantages to the metaphysical views of early scientists. By focusing on mathematical models and efficient causes, while pruning away many of the non-calculable qualities of natural phenomena, scientists were able to develop excellent predictive models. Descartes gave up the study of “final causes” and focused his energies on mathematics because he felt no one could discern God’s purposes, a view adopted widely by subsequent scientists. Both Galileo and Newton put great emphasis on the importance of observation and experimentation in the study of nature, which in many cases put an end to abstract philosophical speculations on natural phenomena that gave no definite conclusions. And Newton gave precise meanings to previously vague terms like “force” and “mass,” meanings that allowed measurement and calculation.

The mistake that these early scientists made, however, was to elevate a method into a metaphysics, by proclaiming that what they studied was the only true reality, with all else existing solely in the human mind. According to Burtt,

[T]he great Newton’s authority was squarely behind that view of the cosmos which saw in man a puny, irrelevant spectator . . . of the vast mathematical system whose regular motions according to mechanical principles constituted the world of nature. . . . The world that people had thought themselves living in — a world rich with colour and sound, redolent with fragrance, filled with gladness, love and beauty, speaking everywhere of purposive harmony and creative ideals — was crowded now into minute corners in the brains of scattered organic beings. The really important world outside was a world hard, cold, colourless, silent, and dead; a world of quantity, a world of mathematically computable motions in mechanical regularity.  (pp. 238-9)

Even at the time this new scientific metaphysics was being developed, it was critiqued on various grounds by philosophers such as Leibniz, Hume, and Berkeley. These philosophers’ critiques had little long-term impact, probably because scientists offered working predictive models and philosophers did not. But today, even as science is promising an eventual “theory of everything,” the limitations of the metaphysics of modern science is causing even some scientists to rethink the whole issue of causation and the role of human sensations in developing knowledge. The necessity for rethinking the modern scientific view of metaphysics will be the subject of my next post.

What Does Science Explain? Part 1 – What is Causation?

In previous posts, I have argued that science has been excellent at creating predictive models of natural phenomena. From the origins of the universe, to the evolution of life, to chemical reactions, and the building of technological devices, scientists have learned to predict causal sequences and manipulate these causal sequences for the benefit (or occasionally, detriment) of humankind. These models have been stupendous achievements of civilization, and religious texts and institutions simply cannot compete in terms of offering predictive models.

There remains the issue, however, of whether the predictive models of science really explain all that there is to explain. While many are inclined to believe that the models of science explain everything, or at least everything that one needs to know, there are actually some serious disputes even among scientists about what causation is, what a valid explanation is, whether predictive models need to be realistic, and how real are some of the entities scientists study, such as the “laws of nature” and the mathematics that are often part of those laws.

The fundamental issues of causation, explanation, and reality are discussed in detail in a book published in 1954 entitled: The Metaphysical Foundations of Modern Science by Edwin Arthur Burtt. According to Burtt, the birth and growth of modern science came with the development of a new metaphysics, that is, the study of being and existence. Copernicus, Kepler, Galileo, and Newton all played a role in creating this new metaphysics, and it shapes how we view the world to this day.

In order to understand Burtt’s thesis, we need to back up a bit and briefly discuss the state of metaphysics before modern science — that is, medieval metaphysics. The medieval view of the world in the West was based largely on Christianity and the ancient Greek philosophers such as Aristotle, who wrote treatises on both physics and metaphysics.

Aristotle wrote that there were four types of answers to the question “why?” These answers were described by Aristotle as the “four causes,” though it has been argued that the correct translation of the Greek word that Aristotle used is “explanation” rather than “cause.” These are:

(1) Material cause

(2) Formal cause

(3) Efficient (or moving) cause

(4) Final cause

“Material cause” refers to changes that take place as a result of the material that something is made of. If a substance melts at a particular temperature, one can argue that it is the material nature of that substance that causes it to melt at that temperature. (The problem with this kind of explanation is that it is not very deep — one can then ask why a material behaves as it does.)

“Formal cause” refers to the changes that take place in matter because of the form that an object is destined to have. According to Aristotle, all objects share the same matter — it is the arrangement of matter into their proper forms that causes matter to become a rock, a tree, a bird, or a human being. Objects and living things eventually disintegrate and perish, but the forms are eternal, and they shape matter into new objects and living things that replace the old. The idea of formal causation is rooted in Plato’s theory of forms, though Aristotle modified Plato’s theory in a number of ways.

“Efficient cause” refers to the change that takes place when one object impacts another; one object or event is the cause, the other is the effect. A stick hitting a ball, a saw cutting wood, and hydrogen atoms interacting with oxygen atoms to create water are all examples of efficient causes.

“Final cause” refers to the goal, end, or purpose of a thing — the Greek word for goal is “telos.” An acorn grows into an oak tree because that is the goal or telos of an acorn. Likewise, a fertilized human ovum becomes a human being. In nature, birds fly, rain nourishes plants, and the moon orbits the earth, because nature has intended certain ends for certain things. The concept of a “final cause” is intimately related to the “formal cause,” in the sense that the forms tend to provide the ends that matter pursues.

Related to these four causes or explanations is Aristotle’s notion of potentiality and actuality. Before things come into existence, one can say that there is potential; when these things come into existence they are actualized. Hydrogen atoms and oxygen atoms have the potential to become water if they are joined in the right way, but until they are so joined, there is only potential water, not actual water. A block of marble has the potential to become a statue, but it is not actually a statue until a sculptor completes his or her work. A human being is potentially wise if he or she pursues knowledge, but until that pursuit of knowledge is carried out, there is only potentiality and not actuality. The forms and telos of nature are primarily responsible for the transformation of potentiality into actuality.

Two other aspects of the medieval view of metaphysics are worth noting. First, for the medievals, human beings were the center of the universe, the highest end of nature. Stars, planets, trees, animals, chemicals, were lower forms of being than humans and existed for the benefit of humans. Second, God was not merely the first cause of the universe — God was the Supreme Good, the goal or telos to which all creation was drawn in pursuit of its final goals and perfection. According to Burtt,

When medieval philosophers thought of what we call the temporal process it was this continuous transformation of potentiality into actuality that they had in mind. . . . God was the One who eternally exists, and ever draws into movement by his perfect beauty all that is potentially the bearer of a higher existence. He is the divine harmony of all goods, conceived as now realized in ideal activity, eternally present, himself unmoved, yet the mover of all change. (Burtt, The Metaphysical Foundations of Modern Science, pp. 94-5)

The rise of modern science, according to Burtt, led to a radical change in humanity’s metaphysical views. A great deal of this change was beneficial, in the sense that it led to predictive models that successfully answered certain questions about natural processes that were previously mysterious. However, as Burtt noted, the new metaphysics of science was also a straitjacket that constricted humanity’s pursuit of knowledge. Some human senses were unjustifiably dismissed as unreliable or deceptive and some types of causation were swept away unnecessarily. How modern science created a new metaphysics that changed humanity’s conception of reality will be discussed in part two.