Estep on Nature and Nature’s God

In his summary of my book Nature and Nature’s God for the Review of Metaphysics, Caleb Estep writes that I hold that in infinite time all possibilities will be realized. This is not quite what I claim. As I explain in the book, what I hold is that all possibilities that approach arbitrarily close to 1 over time will be realized. In infinite time even something with a small probability of occurring, in any given finite duration, may become nearly certain to occur. This will be the case unless the probability of its occurring diminishes continually such that the sum-total of its probability over infinite time (technically the integral of its probability density from t=zero to t=infinity) is still significantly less than one.

For example, if there is a 1/2 probability of getting heads on a first coin flip, then there is only a 1/4 chance of not getting heads at all on two flips. After three flips there is only a 1/8 chance of not getting heads at all, and so on. However, in some cases the odds decrease so rapidly that there is very little chance of ever getting the desired result. For example, if there is a 1/100 chance of getting result X on the first go, and then a 1/10,000 chance of getting it on the next go, and then a 1/1,000,000 chance of getting it on the third go, and so on, then the odds are against ever getting the result.

Further Thoughts on Act and Potency

We would colloquially say about a composite substance that it is “actually water, not potentially water,” yet it contains within itself both the act of water (the substantial form) and the potency for water (prime matter). The prime matter has not lost the potency for being water by having the form, just as I have not lost the power/potency (potentia) to see when I am actually seeing. When I am not asleep I both potentially see and actually see. When I am asleep I only potentially see.

In book 1 of the Physics, (see chapters 7-9, especially 9) Aristotle is at pains to distinguish matter’s being per se a potential principle from its having per accidens the privation of the form. The matter loses its privation of water-form when it changes into water, but it does not lose its potency to be water. Objective potency always involves a privation, involves not being in some condition. In some cases it is merely a privation, as in the objective potency of the universe to exist “before” God makes it out of nothing. In most cases it is a underlying principle together with a privation.

Aquinas’ statement in the First Way that a thing can’t both be in potency and act at the same time and in the same respect concerns objective potency. Aristotle’s definition of motion (the act of the potential insofar as it is potential) has reference to what Aristotle calls “imperfect potency” as opposed to “perfect potency” (Physics III.2; the distinction is elucidated in the Metaphysics and De Anima). The imperfect potency involved in motion implies privation; that is why it is called imperfect. The power of sight is an example of perfect potency. Hence the context of motion explains Aquinas’ statement. The potency to undergo a change is always an objective potency, because a thing can’t undergo a change to a condition once it is in that condition. But it must retain the subjective potency to be in that condition once it has changed to that condition, otherwise it could not be in that condition. The potency required for motion is a potency joined to a privation.

Prime matter is a subjective potency for substantial forms, and it retains the potency for a certain form when it possesses that form, for it is such a potency essentially, and not accidentally. So too essence is a subjective potency for the act of existence; it retains the potency for that existence when it exists. In Thomism there is a real composition of potency and act.

Feser on Potency and Act

In his reply to Graham Oppy’s criticisms of Thomistic cosmological arguments, Ed Feser denies that creatures exist both actually and potentially at the same time. (“Oppy on Thomistic Cosmological Arguments,” Religious Studies 57 [2021], 503–522, at 509.) He states that the constituents of an object such as a cup of water are potentially a cup of water in the abstract, but in the concrete they are only actually a cup of water. The constituents in question, protons, neutrons, and electrons, are, considered in themselves, potentially a cup of water, but as possessing the form of water they are actually a cup of water and no longer potentially a cup of water.

This doesn’t seem right to me. The real distinction of essence and existence makes Aquinas’ position clear. Without existence an essence is precisely nothing, not any kind of potency at all. But in an actually existing finite being the essence is really distinct from its existence, and is related to its existence as potency to act, as Aquinas argues in De Ente et Essentia. Now the essence is in potency to existence, and not in potency to anything else at all. It is in potency to only one kind of substantial existence, and it is not in potency to non-existence. So either the essence has no potency at all, or it bears the potency to exist when it is actually existing through its composition with the act of existence. In other words, essence is what later scholastics called a subjective potency (the kind of potency that is composed with its act and thus coexists with it) rather than an objective potency (the kind of potency that is opposed to act and ceases to be when its act arrives.) Note that “subjective” does not here mean “relative to someone’s perception” but rather “subject, that is, receiver of act.” “Objective” does not mean “real,” but something like “the objective of a productive activity.”

The same must be true of the relationship between prime matter and substantial form. Certainly the case is less clear here: one could be tempted to think that prime matter, as a principle of potency, was only a potency to receive other substantial forms. But this is not the case. Prime matter’s potency does not undergo any change (see Aristotle, Physics I.9, 192a26–29); it is always a potency for all substantial forms. So when it is actualized by the substantial form of water, it retains the subjective potency to the substantial form of water, as well as potencies for all other substantial forms. Water is a composite of potency and act.

On the Word “Art”

In teaching Aristotle and Thomas one often encounters the word “art.” This is generally translating techne in Greek or ars in Latin. I always remark to students that this word is being used more broadly than the way in which the English word “art” is commonly used. It does not just mean the fine arts such as painting, sculpture, music, or poetry, although it does include the latter. The word “art” as used by Thomas and Aristotle should be taken in connection with the English words “artificial,” “artifact,” and “artisan.”

While I think this is true, I’ve realized lately that more must be said; in a way, but only in a way, the word techne is opposed to the word “art” used in reference to the fine arts. For techne or ars means a know-how, a knowledge of how to produce or do something, and it characteristically gives a person the ability to produce or do this thing over and over again. The carpenter, through the art of carpentry, can make many tables with the same characteristics.

But although know-how is certainly present and active in the case of the fine arts, what characterizes them is a kind of moment of genius that is not reproducible or entirely under the control of the artist. Depending on the level of genius, an artist may be able to make many masterpieces, or just one in a lifetime. But in neither case is there a formula or process that can be followed to automatically guarantee the production of a new masterpiece.

What word do the Greeks have for this? One’s first thought is poiesis, the root word for the English word “poetry,” but this can’t be right. Aristotle uses poiein as one of his ten categories, and gives as examples “to lance” and “to cauterize,” definitely not specifically poetic productions. Poiein in philosophical Greek just means “to make or do.” The word one is looking for is mousike, an art in the sense of something metaphorically inspired by the muses. This is what we call “fine art” in English, something that may make use of techne/technique, but which is a non-reproducible product of genius.

Precis to Nature and Nature’s God

Last week Alex Pruss, Rob Koons, and Ed Houser graciously joined me for a panel discussion of my book Nature and Nature’s God. I enjoyed everyone’s comments and objections, and found the event productive. What follows is the precis I delivered to start the discussion.

In my book Nature and Nature’s God I try to do two things: to set the historical record straight about Aquinas’ unmoved mover argument, and to defend a contemporary version of the argument in light of the developments of empirical science. Over the course of two decades I have felt that most presentations of Aquinas’ First Way have read much into the argument that just isn’t in Aquinas’ texts, and at the same time have drawn too much conclusion out of too little argumentation. Although St. Thomas certainly isn’t infallible, he was a brilliant man, and I take him seriously enough to think that understanding him on his own terms is worth my time.

My project on Aquinas’ First Way began when I learned something interesting about his philosophy of nature from Fr. Weisheipl, namely, that on his view heavy bodies fall of their own accord, with nothing impelling them downwards. I knew that Aquinas builds his First Way around the premise that everything in motion is moved by something else, which I thought meant that everything in motion must be currently impelled by some mover. If that mover was in motion, it too was currently impelled by some other mover, until one gets to an unmoved mover that impels them all: God. I found the argument, understood this way, unconvincing. But now it was obvious that this account of the First Way did not square with Aquinas’ own understanding of natural motion. Perhaps a much more interesting unmoved mover argument was hiding underneath a conventional interpretive cover.

Through a careful reading of Aquinas’ more elaborate presentation of the case for an unmoved mover in the Summa contra Gentiles, as well as books seven and eight of Aristotle’s Physics and Aquinas’ commentary on those books, an exciting physical as opposed to metaphysical argument for God’s existence emerged. Aquinas’ argument is clearly rooted in Aristotle’s, but Aristotle’s argument is quite foreign to us. We are used to cosmological arguments that first try to prove that the universe has a beginning, and then that God must exist to get it started. But Aristotle begins his whole argument by attempting to prove that the universe cannot possibly have a beginning. That proposition then becomes the key premise on which his entire argument for God’s existence is based: only an infinite power could cause a motion of infinite duration, and no physical being can have infinite power. Thus an immaterial being exists.

Aristotle’s argument proved troublesome for Jewish, Christian, and Muslim Aristotelians. It is a central tenet of traditional western monotheism that the universe had a beginning in time, a moment of creation. But the culminating piece of Aristotle’s philosophical corpus was built around a denial of this point of religious doctrine. Moses Maimonides solved the problem, and Aquinas adopts his solution in the Summa contra Gentiles. Aristotle’s proof is placed inside a disjunctive wrapper: although we can’t prove that the universe has a beginning, clearly it either did or it didn’t. If it did, then since nothing brings itself into being God must exist to get the universe started. If it didn’t have a beginning, then it has been in motion for an infinite duration, and then Aristotle’s argument applies: God must exist to sustain the universe forever in motion.

But why must a motion of infinite duration be sustained by an immaterial mover? Motion can either flow from a thing’s nature, or be imposed by an external agent. In Aquinas’ view motion that flows from a thing’s nature is directed, as to an end, to the thing’s natural place and condition. When the body reaches its natural place it naturally comes to rest and does not undergo any further motion unless acted on by an external agent. Hence natural motion is of finite duration. But external agents can only act insofar as they are moved or energized by the motion of a heavenly body. Otherwise the would-be agent and would-be patient simply undergo motion to their natural conditions and rest in equilibrium with each other. Only a voluntary, rational agent can continue to move other things in a cycle indefinitely, and such an agent, if physical, will be of finite power and wear down over time.

According to St. Thomas, all things on earth would move towards their natural place, rest there, cease to act on one another, and remain forever motionless if it were not for the fact that the cyclical motions of the heavenly bodies continued to affect terrestrial things and change them out of their natural conditions, stirring the pot so to speak. But the motions of the heavenly bodies cannot themselves flow from the nature of the heavenly bodies, since such motion is cyclical and does not head towards a natural place of rest. Thus the heavenly bodies themselves are moved by one or more immaterial agents. Since their motions are choreographed to maintain the natural cycles and teleological harmony here on earth, they must ultimately be directed by one single, overarching immaterial being who is directed by no other. This is God.

But if this is Aquinas’ argument, of what interest is it to us? We surely do not believe in the harmony of the spheres, celestial influences, or angels moving planets around. We also believe in inertia, indefinite “natural” motion that never comes to rest, and the conservation of energy as a physical law. Yet a closer look at modern physics reveals that Aquinas’ proof, in its essentials, is just as valid as ever. Allow me to explain.

Isaac Newton, the father of modern physics, stated that “The force of inertia is a passive principle by which bodies persist in their motion or rest. . . . By this principle alone there never could have been any motion in the world. Some other principle was necessary for putting bodies into motion; and now [that] they are in motion, some other principle is necessary for conserving the motion. . . . By reason of the tenacity of fluids, and attrition of their parts, and the weakness of elasticity in solids, motion is much more apt to be lost than got, and is always upon the decay.” No collision is perfectly elastic, and so bodies rebound with less total motion than they had before. Also, the viscosity of fluids and the friction of surfaces diminishes the motion of bodies moving through them or over them.

But the first law of thermodynamics states that no physical process can increase or decrease the amount of energy in a closed system. The energy of motion (“kinetic energy”) lost in collisions and through friction is converted to the energy of heat. Heat, in turn, can be converted back into kinetic energy by various processes, for example in steam locomotives. It might seem, then, that Aquinas’ argument cannot work: no immaterial mover is needed to sustain motion in the universe forever; the energy conservation law guarantees perpetual motion.

Yet scientists are agreed that a perpetual motion machine cannot exist. Why? Because the first law of thermodynamics is complemented by the second law of thermodynamics, according to which the energy in a closed system can only decrease in its availability, and never increase. “Entropy” is a measure of the degree to which energy has become unavailable to do work, and thus the second law is often expressed thus: the entropy of a closed system can only increase, and never decrease. As energy goes through various changes, it tends to find its way into the form of heat, and thermal energy can only be converted back into other forms of energy if it exists in two bodies at different temperatures. But heat tends to flow directly from a hotter body to a cooler body without doing any work, and thus more and more of a system’s energy tends to get locked away into the unusable form of thermal energy at equilibrium. In a sense Aquinas was right: all bodies head towards the natural condition of equilibrium, and once they reach it they never change again unless acted on by an external agent.

Rudolf Clausius, the pioneer of thermodynamics who coined the term entropy, applied the two laws of thermodynamics to the universe at large and expressed them thus: “1. The energy of the universe is constant. 2. The entropy of the universe tends to a maximum.” Modern physics and cosmology conclude that the universe must inevitably head towards the equilibrium condition known as the “heat death,” in which all energy is dissipated about the universe at even temperature, and nothing interesting ever happens again.

All bodies radiate their energy away in the form of photons; they shine at a frequency proportional to their temperature. This is why night-vision goggles work. But many of these photons inevitably escape into space where they undergo a cosmological redshift as the universe expands. As each photon’s frequency drops, the amount of energy it can deliver upon impact with a body diminishes. Thus energy is spread thinner and thinner throughout space, entropy rises, and the universe gets ever closer to the heat death.

If the universe had no beginning and had already existed for an infinite time, it would already have reached heat death. The fact that it has not indicates that the universe had a beginning a finite duration in the past. Alternatively—and this is where Aquinas’ argument differs from that the kalam argument—the universe has existed forever but God has either continuously or periodically converted some of its high-entropy, unavailable energy back into usable energy. In that way it could still exist out of equilibrium as it does today, without having a beginning in time.

Thus Aquinas’ unmoved mover argument is still valid, and serves as a welcome, more accessible addition to his other arguments, particularly to the argument based on the act of existence in De Ente et Essentia. The De Ente argument is a powerful one, but very subtle and not very accessible to the empirically minded. Aquinas’ more physical unmoved mover argument is a good first step for such people.

Infinite Regress and the Hume-Edwards-Ockham Objection

My article “Infinite Regress and the Hume-Edwards-Ockham Ojbection: A Thomistic Analysis” just appeared in the Proceedings of the American Catholic Philosophical Association. Here is the abstract:

One of the standard objections against the impossibility of infinite regress is associated with David Hume and Paul Edwards, but originates with William Ockham. They claim that in an infinite regress every member of the series is explained, and nothing is unexplained. Every member is explained by the one before it, and the series as a whole is nothing over and above its members, and so needs no cause of its own. Utilizing the well-known Thomistic distinction between essentially ordered and accidentally ordered causal series, I show that the Hume-Edwards-Ockham objection fails to touch Aquinas’s argument against the impossibility of infinite regress in an essentially ordered series. However, Aquinas also argues that accidentally ordered causal series can only regress infinitely if supported by an everlasting essential cause. The Hume-Edwards-Ockham objection does raise a question about this thesis, but I show how St. Thomas can reply to it convincingly.

Of note is that I respond in the endnotes to recent arguments in favor of infinite regress by Thomas Oberle.

Forthcoming Article Responding to Schmid on “Providential Collapse”

The International Journal for Philosophy of Religion just accepted my article “Simply Providential: A Thomistic Response to Schmid’s Providential Collapse Argument against Classical Theism.” In the article I advance an alternative to the standard “extrinsic predication” defense of classical theism from modal collapse. That defense works against modal collapse but leaves the doctrine of divine simplicity open to the charge of “providential collapse,” as Schmid has argued. On the extrinsic predication view the creation that obtains determines what it is that God can truly be said to will, rather than the other way around as ought to be the case. I argue that the extrinsic predication thesis is true of God’s “transeunt” actions such as creating this universe, but it is not true of God’s “immanent” actions such as willing or intending this universe to exist. The predication “God wills this creature to exist” is true in virtue of God’s relation to the creature, rather than in virtue of the creature itself. God, with no differences in Himself, can relate himself differently to possible creatures, thereby freely determining which will exist and which will not. Although Aquinas calls such relations of God to creatures “relations of reason,” the relations of willing and knowing these creatures is not, according to Aquinas, a relation invented by our reason reflecting on God and creatures. Rather, they are relations established by God’s reason. Such relations would exist even if no rational creatures existed.

Here is the abstract:

Classical theism is often said to suffer from the problem of modal collapse: if God is necessary and simple then all of his effects (creatures) are also necessary. Classical theists often turn to extrinsic predication in response: God’s simple and necessary act is compatible with any number of possible effects or no effects, and is only said to be an act of creating in virtue of the existence of the universe itself. Leftow and Schmid criticize this solution for leading to “providential collapse”: God would not have control over which creation obtains if all his intrinsic features are compatible with any possible universe. Thomistic classical theism avoids both modal collapse and providential collapse by utilizing the metaphysics of relations. With no differences within his simple essence/act, God can relate himself differently to his possible effects, willing some and not willing others. These relations determine whether or not the effects will exist. Thomistic classical theism’s version of divine simplicity is not incompatible with God having a multiplicity of relations, for the three divine persons are distinct relations within God. Divine simplicity is only incompatible with a multiplicity of absolute items within God. Furthermore, not all differences of relation are grounded in the different absolute characteristics of their relata. Rather, sometimes differences of relation themselves ground differences in the absolute characteristics of their relata. God’s divine act is thus said to be an act of willing this creation rather than that in virtue of his chosen relation to possible creatures, rather than in virtue of the creatures themselves.