Essay: Cosmological Argument

February 26, 2013.

Print Friendly, PDF & Email

Evaluate the Strengths and Weaknesses of the Cosmological Argument for Proving God Exists. (40)

This essay, of A grade standard, has been submitted by a student. PB

The Cosmological argument is an argument put forward by the Christian Philosopher St. Thomas Aquinas (1225-1274) in an attempt to prove God’s existence. However, it is important to take into account that Aquinas already had a strong belief in God when putting this theory forward in his Summa Theologiae, meaning that instead of trying to prove God’s existence, he was more trying to solidify his already established faith based on reason through looking at the cause of the Universe which Aquinas claims must be God.

In Summa Theolgiae, Aquinas attempts to logically prove God’s existence in his ‘Five Ways’ though the first three are the ones which are predominantly sited when referring to the argument. The first of these give ways make similar points based on the idea that infinite regression is not possible; there must have been one thing that started off everything that happened. Aquinas argues that this must be God. The first way is an argument for an ‘Unmoved Mover’. Here, Aquinas claims that everything in the world is in a constant state of change or ‘motion.’ He goes on to argue that something cannot be both potentially and actually the same thing; a cup of boiling tea could not be both hot and potentially hot, though it could be potentially cold and actually hot. By this logic, everything which is ‘in a state of movement’ must have been put into this state by a different object. Because of Aquinas’ rejection of the possibility of infinite regression, this means that there must have been a ‘first mover’ who is ‘put into motion by no other.’ This is, by Aquinas’ logic.

The second way makes a very similar point and is an argument for an ‘Uncaused Cause.’ Aquinas starts off by stating that nothing can be an efficient cause of itself; everything is caused by something else. The efficient causes of a thing follow in order meaning that there was a first cause which caused a second cause and so on and so forth. Once again, because Aquinas rejects the possibility of infinite regression, this means that ‘it is necessary to admit a first efficient cause to which everyone gives the name of God.’ Both of these two ways are heavily influenced by Aristotle’s idea of a prime mover. However, Aquinas does not mean to argue that God is merely the being that started off the chain of events which lead to cause the universe and everything in it. He is rather claiming that he must still exist; Coppleston used the example of winding up a pocket watch every night rather than knocking over the first domino in a chain.

These two ways leave Aquinas’ argument open for several criticisms, as well as showing some strength. One such strength is the way in which it is a satisfying argument for Humans to understand. It is true that, by human, a posteriori logic, things must indeed have a cause which exists outside its own essence or self. We as humans were caused by our parents and the universe was caused by the big bang. However, if the big bang required matter to take place, then that matter, logically, had to have been caused by something and put into the correct environment for the event to take place. Aquinas argues that this causer must have been God.

However, it is possible to severely weaken Aquinas’ argument if you argue that it is in fact possible to have an infinite chain of regression. Seeing as the argument is hinged upon the assumption that this is impossible, disregarding this assumption therefore dramatically reduces the strength of the argument. The philosopher David Hume questioned the very notion of cause and effect. He argued that we make assumptions about the relationship between Cause and Effect which are by no means necessarily true. While it is true that, according to human logic, infinite regression does not seem logical, in mathematics, it is possible to have an infinite series of regression; numbers can keep increasing or decreasing in size infinitely, thereby proving that infinite regression is entirely possible. Using a posteriori knowledge, it may seem apparent that every effect has a cause. However, if you use a priori knowledge, you could easily reason that, not everything which exists has a cause. It is impossible to claim that this is analytically true. Hume would argue that the universe is just a ‘brute fact’; it just is and has no cause. This completely undermines Aquinas’ first two ways.

The Fallacy of Composition is another weakness of Aquinas’ first two ways which David Hume outlines and uses to weaken the Cosmological Argument. While it may indeed be true that everything in the universe does have a cause, it does not necessarily mean that the universe itself has a cause; the fact that everything which humans can observe can be explained by a precedent cause, this doesn’t mean that the universe can be explained in the same way. The atheistic philosopher Betrand Russell agrees with this point and claims that while all humans have mothers, ‘Obviously, the human race hasn’t a mother, that’s a different logical sphere’ in his book Why I Am Not a Christian.

However, Aquinas’ argument can be re-strengthened through Anscombe’s criticism of Hume’s criticism in ‘Whatever Has a Beginning of Existence Must Have a Cause’: Hume’s Argument Exposed. In this work, Anscombe argues that while it possible to imagine something coming into existence without a cause, this does not mean that it is ‘possible to suppose “without contradiction of absurdity”’ that this is the case. For example, though it may be possible to imagine a magician pulling a rabbit out of a magician’s hat without having a cause of its existence, this does not mean that it is logical to think that it is possible. By this logic, while it is possible imagining the universe coming into existence without a cause, that does not mean that it is logical or reasonable to think so.

Aquinas goes on to attempt to further strengthen his Cosmological Argument in his Third Way: The Argument from Contingency. In this way, Aquinas argues that all things which exist in nature are contingent; they did not exist, in the future will cease to exist and, as well as this, it is possible for them never to have come into existence. Aquinas believed that, using this logic, the fact that everything used to not exist must mean that there was a time when nothing at all existed because there would be nothing to bring anything else into existence. Therefore, ‘there must exist something the existence of which is necessary.’ Aquinas goes on to state that, because he believes infinite regression to be impossible, there must be ‘some being having of itself its own necessity…causing in others their necessity’ which he argues is God. In other words, seeing as how there was once a time when nothing contingent existed, there must have been a non-contingent, necessary being which is necessary in itself to cause the existence of contingent things.

This third way could be argued to be either strong or weak. One strength which the argument holds is that, as with the first two ways, this argument appeals strongly to human reason and logic, leading it to be widely accepted by empiricists. In accordance with human logic, things in existence are indeed caused by other things; we are made by our parents, mountains are made by tectonic plate movement etc. Aquinas draws on this logic when putting forward his third way, meaning that it is a fairly satisfying argument.

However, there are also several strengths which are pointed out by philosophers including Immanuel Kant and J.L. Mackie. Kant’s criticism lies in his rejection of the concept of necessary existence. He entirely rejects the idea of the existence of a subject being necessary; existence could not possibly be a defining predicate of a sunject as it adds nothing to the definition of the subject. In other words, nothing can be necessary. However, this criticism could be weakened by arguing that Kant is just rehashing his criticism of the Ontological differences despite the obvious differences in the Ontological and Cosmological Arguments (Ontological Argument is a priori, Cosmological argument is a posteriori).

Another weakness of the Cosmological Argument is put forward by J.L. Mackie in his The Miracle of Theism. Mackie accepts the logic behind Aquinas’ third way up until the point when he claims that the cause of all contingent objects must be a necessary being. While Aquinas argues that all contingent things whose essence does not include existence must rely on a necessary being to exist, Mackie retorts by claiming that this is not necessarily true; contingent beings could be argued, by Aquinas’ logic, to have been ultimately caused by some necessary stock of matter which has always existed and always will. This severely undermines Aquinas’ third way by proving that Aquinas’ logic has not actually managed to prove the necessity of a Christian God, but rather just some necessary thing-a being, beings or otherwise.

The Cosmological Argument for proving God’s existence has a number of clear strengths and weaknesses. Personally, however, I would argue that the argument’s criticisms outweigh its strengths, thereby making it a weak argument for proving God’s existence. One clear strength of the arguments is its appeal to human logic and reason. As an a posteriori argument which is based on human experience, it satisfies human assumptions. It is illogical to humans to think of an infinite chain of regression in regards to anything, let alone to creation of the universe. However, this strength does not necessarily add to the arguments ability to prove the existence of God, but more to the accessibility of the argument to a wide range of people.

Conversely, perhaps the most severe and damaging criticism of this argument is the idea that an infinite chain of regression is in fact possible. When writing Summa Theologiae and outlining his Cosmological Argument, Aquinas makes the assumption that it is impossible to have an infinite chain of aggression; there must be an ‘uncaused causer’ or ‘unmoved mover.’ However, retrospectively, this assumption is by no means necessarily true. In terms of mathematics, infinite regress is entirely possible as it is always possible to increase or decrease a number. Therefore, it is definitely possible to infinitely regress. This hugely takes away from the strength of the argument as it is upon this assumption which Aquinas bases his entire premise.

On the other hand, a clear strength of the argument is outlined by Copleston in his radio debate with Russell in 1948, and that is that the argument does offer a sound reason as to why anything exists through developing on Aquinas’ Argument from Contingency. In this debate, Copleston claims that the universe is, in itself, not a physical thing, it is instead the aggregate (or sum of) all the objects which it contains. He goes on to argue that all the things which make up the universe are contingent and, as a result, do not contain their own reason for existence. Therefore, seeing as the universe is the aggregate of these contingent parts, the universe itself must also be contingent and therefore have a cause outside of itself; Copleston argues (and Aquinas would agree) that the only feasible cause of the universe is God. While this is an obvious strength, the degree to which it strengthens the argument could be brought into existence because, once again, it relies on the assumption that an infinite regression is not possible which, if untrue, would completely unbalance the entire argument.

Contrary to this, there is another very obvious weakness to the argument which contradicts this idea of God being the only feasible explanation for the creation of the contingent argument, and that is that, while Aquinas’ logic in building up to this conclusion is sound, his reasoning does nothing to prove that it is the omniscient, omnibenevolent, omnipresent, omnipotent God of Christianity which initially caused the creation of the Universe. This argument is put forward by J.L Mackie. He argues that, assuming that Aquinas is right in claiming there cannot be infinite regression, and assuming that the existence of everything contingent relies on the existence of some necessary thing, there is no proof that the initial cause of the universe is a necessary being. He claims that, by Aquinas’ logic, the cause could be a ‘permanent stock of matter whose essence did not involve existence from anything else.’ Equally, the creator of the universe could well be a necessary being, but not the Christian God; it could be Allah, or even the multiple Gods of Hinduism. This again is a strong criticism of Aquinas’ argument as it shows that, even if his logic in reaching his conclusion is accurate, his conclusion lacks evidence and therefore, does not prove the existence of a Christian God.

Yet another obvious weakness of the Cosmological argument was highlighted by Russell in the afore mentioned radio debate with Copleston and is supported by David Hume. While, according to human reason, all effects have a cause (a headache may be caused by banging your head, or you may put on weight from eating a lot of fatty foods), it is, Russell and Hume would argue, to assume that this is true in the case of the universe as it is ‘a different logical sphere.’ Russell used the example of humans; while all humans have mothers, ‘Obviously, the human race doesn’t have a mother.’ Therefore, not only is there the possibility of infinite regress when looking at the cause of the universe, but there is also the possibility that the universe is ‘just a brute fact’ which ‘just is’ and needs no further explanation; in laymen’s terms, it has always existed. Again, this is a clearly thought out criticism of the Cosmological Argument which takes away from its strength.

Overall, therefore, it is clear that, while not without its strengths, the Cosmological Argument is a weak argument for proving God’s existence as it lacks in both the proof given for the conclusion that the Christian God is the cause of the universe and in the logic behind the concept that the universe must have had a cause at all.

Study with us

Peped Online Religious Studies Courses

Practise Questions 2020

OCR Religious Studies Practise Questions front cover

Religious Studies Guides – 2020

Religious Studies Philosophy of Religion OCR Revision Complete Guide – New Edition (2020)

Check out our great books in the Shop

Thank you. Very concise and helpful

Leave a Reply Cancel

This site uses Akismet to reduce spam. Learn how your comment data is processed .

6.3 Cosmology and the Existence of God

Learning objectives.

By the end of this section, you will be able to:

  • Describe teleological and moral arguments for the existence of God.
  • Outline Hindu cosmology and arguments for and against the divine.
  • Explain Anselm’s ontological argument for the existence of God.
  • Articulate the distinction between the logical and evidential problems of evil.

Another major question in metaphysics relates to cosmology. Cosmology is the study of how reality is ordered. How can we account for the ordering, built upon many different elements such as causation, contingency, motion, and change, that we experience within our reality? The primary focus of cosmological arguments will be on proving a logically necessary first cause to explain the order observed. As discussed in earlier sections, for millennia, peoples have equated the idea of a first mover or cause with the divine that exists in another realm. This section will discuss a variety of arguments for the existence of God as well as how philosophers have reconciled God's existence with the presence of evil in the world.

Teleological Arguments for God

Teleological arguments examine the inherent design within reality and attempt to infer the existence of an entity responsible for the design observed. Teleological arguments consider the level of design found in living organisms, the order displayed on a cosmological scale, and even how the presence of order in general is significant.

Aquinas’s Design Argument

Thomas Aquinas’s Five Ways is known as a teleological argument for the existence of God from the presence of design in experience. Here is one possible formulation of Aquinas’s design argument:

  • Things that lack knowledge tend to act toward an end/goal.
  • It is obvious that it is not by chance.
  • Things that lack knowledge act toward an end by design.
  • If a thing is being directed toward an end, it requires direction by some being endowed with intelligence (e.g. the arrow being directed by the archer).
  • Therefore, some intelligent being exists that directs all natural things toward their end. This being is known as God.

Design Arguments in Biology

Though Aquinas died long ago, his arguments still live on in today’s discourse, exciting passionate argument. Such is the case with design arguments in biology. William Paley (1743–1805) proposed a teleological argument, sometimes called the design argument, that there exists so much intricate detail, design, and purpose in the world that we must suppose a creator. The sophistication and incredible detail we observe in nature could not have occurred by chance.

Paley employs an analogy between design as found within a watch and design as found within the universe to advance his position. Suppose you were walking down a beach and you happened to find a watch. Maybe you were feeling inquisitive, and you opened the watch (it was an old-fashioned pocket watch). You would see all the gears and coils and springs. Maybe you would wind up the watch and observe the design of the watch at work. Considering the way that all the mechanical parts worked together toward the end/goal of telling time, you would be reluctant to say that the watch was not created by a designer.

Now consider another object—say, the complexity of the inner workings of the human eye. If we can suppose a watchmaker for the watch (due to the design of the watch), we must be able to suppose a designer for the eye. For that matter, we must suppose a designer for all the things we observe in nature that exhibit order. Considering the complexity and grandeur of design found in the world around us, the designer must be a Divine designer. That is, there must be a God.

Often, the design argument is formulated as an induction:

  • In all things we have experienced that exhibit design, we have experienced a designer of that artifact.
  • The universe exhibits order and design.
  • Given #1, the universe must have a designer.
  • The designer of the universe is God.

Think Like a Philosopher

Read “ The Fine-Tuning Argument for the Existence of God ” by Thomas Metcalf.

Evaluate the arguments and counterarguments presented in this short article. Which are the most cogent, and why?

Moral Arguments for God

Another type of argument for the existence of God is built upon metaethics and normative ethics. Consider subjective and objective values. Subjective values are those beliefs that guide and drive behaviors deemed permissible as determined by either an individual or an individual’s culture. Objective values govern morally permissible and desired outcomes that apply to all moral agents. Moral arguments for the existence of God depend upon the existence of objective values.

If there are objective values, then the question of “Whence do these values come?” must be raised. One possible answer used to explain the presence of objective values is that the basis of the values is found in God. Here is one premise/conclusion form of the argument:

  • If objective values exist, there must be a source for their objective validity.
  • The source of all value (including the validity held by objective values) is God.
  • Objective values do exist.
  • Therefore, God exists.

This argument, however, raises questions. Does moral permissibility (i.e., right and wrong) depend upon God? Are ethics an expression of the divine, or are ethics better understood separate from divine authority? To explore this topic further, students will find a helpful overview and updated references in the Stanford Encyclopedia article, " Moral Arguments for the Existence of God ."

Write Like a Philosopher

Watch “ God & Morality: Part 2 ” by Steven Darwall.

Darwall’s argument for the autonomy of ethics may be restated as follows:

  • God knows morality best (1:44).
  • God knows what is best for us (2:12).
  • God has authority over us (2:48).

How does Darwall refute the conclusion? What is the evidence offered, and at what point within the argument is the evidence introduced? What does his approach suggest about refutational strategies? Can you refute Darwall’s argument?

As you write, begin by defining the conclusion. Remember that in philosophy, conclusions are not resting points but mere starting points. Next, present the evidence, both stated and unstated, and explain how it supports the conclusion.

The Ontological Argument for God

An ontological argument for God was proposed by the Italian philosopher, monk, and Archbishop of Canterbury Anselm (1033–1109). Anselm lived in a time where belief in a deity was often assumed. He, as a person and as a prior of an abbey, had experienced and witnessed doubt. To assuage this doubt, Anselm endeavored to prove the existence of God in such an irrefutable way that even the staunchest of nonbelievers would be forced, by reason, to admit the existence of a God.

Anselm’s proof is a priori and does not appeal to empirical or sense data as its basis. Much like a proof in geometry, Anselm is working from a set of “givens” to a set of demonstrable concepts. Anselm begins by defining the most central term in his argument—God. For the purpose of this argument, Anselm suggests, let “God” = “a being than which nothing greater can be conceived.” He makes two key points:

  • When we speak of God (whether we are asserting God is or God is not), we are contemplating an entity who can be defined as “a being than which nothing greater can be conceived.”
  • When we speak of God (either as believer or nonbeliever), we have an intramental understanding of that concept—in other words, the idea is within our understanding.

Anselm continues by examining the difference between that which exists in the mind and that which exists both in the mind and outside of the mind. The question is: Is it greater to exist in the mind alone or in the mind and in reality (or outside of the mind)? Anselm asks you to consider the painter—for example, define which is greater: the reality of a painting as it exists in the mind of an artist or that same painting existing in the mind of that same artist and as a physical piece of art. Anselm contends that the painting, existing both within the mind of the artist and as a real piece of art, is greater than the mere intramental conception of the work.

At this point, a third key point is established:

Have you figured out where Anselm is going with this argument?

  • If God is a being than which nothing greater can be conceived (established in #1 above);
  • And since it is greater to exist in the mind and in reality than in the mind alone (established in #3 above);
  • Then God must exist both in the mind (established in #2 above) and in reality;
  • In short, God must be. God is not merely an intramental concept but an extra-mental reality as well.

Hindu Cosmology

One of the primary arguments for the existence of God as found within Hindu traditions is based on cosmological conditions necessary to explain the reality of karma. As explained in the introduction to philosophy chapter and earlier in this chapter, karma may be thought of as the causal law that links causes to effects. Assuming the doctrine of interdependence, karma asserts that if we act in such a way to cause harm to others, we increase the amount of negativity in nature. We therefore hurt ourself by harming others. As the self moves through rebirth ( samsara ), the karmic debt incurred is retained. Note that positive actions also are retained. The goal is liberation of the soul from the cycle of rebirth.

Maintenance of the Law of Karma

While one can understand karmic causality without an appeal to divinity, how the causal karmic chain is so well-ordered and capable of realizing just results is not as easily explainable without an appeal to divinity. One possible presentation of the argument for the existence of God from karma could therefore read as follows:

  • If karma is, there must be some force/entity that accounts for the appropriateness (justice) of the karmic debt or karmic reward earned.
  • The source responsible for the appropriateness (justice) of the debt or reward earned must be a conscious agent capable of lending order to all karmic interactions (past, present, and future).
  • Karmic appropriateness (justice) does exist.
  • Therefore, a conscious agent capable of lending order to all karmic interactions (past, present, and future) must exist.

Physical World as Manifestation of Divine Consciousness

The cosmology built upon the religious doctrines allows for an argument within Hindu thought that joins a version of the moral argument and the design argument. Unless a divine designer were assumed, the moral and cosmological fabric assumed within the perspective could not be asserted.

Hindu Arguments Against the Existence of God

One of the primary arguments against the existence of God is found in the Mīmāmsā tradition. This ancient school suggests that the Vedas were eternal but without authors. The cosmological and teleological evidence as examined above was deemed inconclusive. The focus of this tradition and its several subtraditions was on living properly.

Problem of Evil

The problem of evil poses a philosophical challenge to the traditional arguments (in particular the design argument) because it implies that the design of the cosmos and the designer of the cosmos are flawed. How can we assert the existence of a caring and benevolent God when there exists so much evil in the world? The glib answer to this question is to say that human moral agents, not God, are the cause of evil. Some philosophers reframe the problem of evil as the problem of suffering to place the stress of the question on the reality of suffering versus moral agency.

The Logical Problem of Evil

David Hume raised arguments not only against the traditional arguments for the existence of God but against most of the foundational ideas of philosophy. Hume, the great skeptic, starts by proposing that if God knows about the suffering and would stop it but cannot stop it, God is not omnipotent. If God is able to stop the suffering and would want to but does not know about it, then God is not omniscient. If God knows about the suffering and is able to stop it but does not wish to assuage the pain, God is not omnibenevolent. At the very least, Hume argues, the existence of evil does not justify a belief in a caring Creator.

The Evidential Problem of Evil

The evidential problem considers the reality of suffering and the probability that if an omnibenevolent divine being existed, then the divine being would not allow such extreme suffering. One of the most formidable presentations of the argument was formulated by William Rowe :

There exist instances of intense suffering which an omnipotent, omniscient being could have prevented without thereby losing some greater good or permitting some evil equally bad or worse. An omniscient, wholly good being would prevent the occurrence of any intense suffering it could, unless it could not do so without thereby losing some greater good or permitting some evil equally bad or worse. (Therefore) there does not exist an omnipotent, omniscient, wholly good being. (Rowe 1979, 336)

Western Theistic Responses to the Problem of Evil

Many theists (those who assert the existence of god/s) have argued against both the logical and evidential formulations of the problem of evil. One of the earliest Christian defenses was authored by Saint Augustine. Based upon a highly Neo-Platonic methodology and ontology, Augustine argued that as God was omnibenevolent (all good), God would not introduce evil into our existence. Evil, observed Augustine, was not real. It was a privation or negation of the good. Evil therefore did not argue against the reality or being of God but was a reflection for the necessity of God. Here we see the application of a set of working principles and the stressing of a priori resulting in what could be labeled ( prima facie ) a counterintuitive result.

An African Perspective on the Problem of Evil

In the above sections, the problem of evil was centered in a conception of a god as all-powerful, all-loving, and all-knowing. Evil, from this perspective, reflects a god doing evil (we might say reflecting the moral agency of a god) and thus results in the aforementioned problem—how could a “good” god do evil or perhaps allow evil to happen? The rich diversity of African thought helps us examine evil and agency from different starting points. What if, for example, the lifting of the agency (the doing of evil) was removed entirely from the supernatural? In much of Western thought, God was understood as the creator. Given the philosophical role and responsibilities that follow from the assignment of “the entity that made all things,” reconciling evil and creation and God as good becomes a problem. But if we were to remove the concept of God from the creator role, the agency of evil (and reconciling evil with the creator) is no longer present.

Within the Yoruba-African perspective, the agency of evil is not put upon human agency, as might be expected in the West, but upon “spiritual beings other than God” (Dasaolu and Oyelakun 2015). These multiple spiritual beings, known as “Ajogun,” are “scattered around the cosmos” and have specific types of wrongdoing associated specifically with each being (Dasaolu and Oyelakun 2015). Moving the framework (or cosmology) upon which goodness and evil is understood results in a significant philosophical shift. The meaning of evil, instead of being packed with religious or supernatural connotations, has a more down-to-earth sense. Evil is not so much sin as a destruction of life. It is not an offense against an eternal Creator, but an action conducted by one human moral agent that harms another human moral agent.

Unlike Augustine’s attempt to explain evil as the negation of good (as not real), the Yoruban metaphysics asserts the necessity of evil. Our ability to contrast good and evil are required logically so that we can make sense of both concepts.

As an Amazon Associate we earn from qualifying purchases.

This book may not be used in the training of large language models or otherwise be ingested into large language models or generative AI offerings without OpenStax's permission.

Want to cite, share, or modify this book? This book uses the Creative Commons Attribution License and you must attribute OpenStax.

Access for free at https://openstax.org/books/introduction-philosophy/pages/1-introduction
  • Authors: Nathan Smith
  • Publisher/website: OpenStax
  • Book title: Introduction to Philosophy
  • Publication date: Jun 15, 2022
  • Location: Houston, Texas
  • Book URL: https://openstax.org/books/introduction-philosophy/pages/1-introduction
  • Section URL: https://openstax.org/books/introduction-philosophy/pages/6-3-cosmology-and-the-existence-of-god

© Dec 19, 2023 OpenStax. Textbook content produced by OpenStax is licensed under a Creative Commons Attribution License . The OpenStax name, OpenStax logo, OpenStax book covers, OpenStax CNX name, and OpenStax CNX logo are not subject to the Creative Commons license and may not be reproduced without the prior and express written consent of Rice University.

1000-Word Philosophy: An Introductory Anthology

1000-Word Philosophy: An Introductory Anthology

Philosophy, One Thousand Words at a Time

Cosmological Arguments for the Existence of God

Author: Thomas Metcalf Categories: Philosophy of Religion , Metaphysics , Historical Philosophy , Islamic Philosophy Word count: 1000

Why is there something rather than nothing?

Intuitively, it could have been that nothing existed at all. Yet you and I are here, plus a whole universe of other stuff. We had parents, and our parents had parents, who had parents, and on and on, back to prebiotic chemicals, stars, and the Big Bang. Yet where did that come from? Surely there has to be some sort of ultimate explanation, right?

Cosmological arguments for God’s existence propose that God is the ultimate explanation or cause of everything. Such arguments begin with an empirical observation of the world—that there is motion, or causes, or just ordinary things that exist—and conclude this observation is explained by God’s existence. [1]

This essay surveys three types of cosmological arguments.

1. First-Cause Cosmological Arguments

St. Thomas Aquinas (1225-1274 CE) argued that all the causation and motion we observe can be traced back to God, who is an uncaused cause or unmoved mover. [2] In summary:

  • Causes and motion exist.
  • All causation or motion requires some prior cause or motion; things can’t happen nor move for no reason.
  • But this chain of causation or motion cannot go back infinitely; this seems impossible.
  • Therefore, there must be a first, uncaused cause, or first, unmoved mover.
  • The most plausible example of an uncaused cause or unmoved mover would be God.
  • Therefore, God exists.

To challenge (2), one might argue that things can just happen for no reason. [3]

To challenge (3), one might question why exactly there can’t be an infinite regress of causes or motion: while this is puzzling, so are uncaused causes. [4] Some also argue that there’s a contradiction in the argument: if every cause or motion requires some prior cause, then how would God or anything else be an exception? [5]

To challenge the inference from (4) to (5), the existence of God, one might ask why the first cause or unmoved mover can’t be a mindless event: it wouldn’t have to be God or a god. [6]

2. The Kalām Cosmological Argument

While the First-Cause Argument holds that God is the ultimate originator of all causation or motion, the kalām cosmological argument is more limited: it’s only about God’s being the cause of the universe’s beginning to exist . [7] In summary:

  • If something begins to exist, its existence was caused by something.
  • The universe, including time and space, cannot go back infinitely far in time.
  • Therefore, the universe began to exist.
  • Therefore, the universe’s existence was caused by something.
  • The most-plausible example of a creator of time and space would be something like God.

The standard support for (1) is both intuitive and empirical. If things could begin to exist uncaused, why doesn’t that happen more often? In terms of our experience, that just doesn’t happen.

Following William Lane Craig, the leading contemporary proponent of this argument, we can distinguish both philosophical and scientific evidence for (2). [8]

The main philosophical evidence is that, allegedly, infinities cannot exist in the real physical world: e.g., a hotel with infinitely many rooms could be full, and yet accommodate infinitely many buses each containing infinitely many passengers without having to kick anyone out. [9]

The main scientific evidence comes from cosmology and thermodynamics, both of which suggest that the universe cannot have an infinite past.

Still, philosophers and scientists have challenged both of these alleged lines of evidence: e.g., there is some scientific evidence that the universe did not begin to exist after all. [10]

As for (5), again, why would the creator have to be God? Why couldn’t it have been some mindless force instead? Defenders of the argument sometimes reply as follows: [11]

All causes we know of are either mechanical or personal, where mechanical causes are just events that follow laws of nature, and personal causes are intentional—perhaps freely-willed—choices. Yet laws of nature seem to be a part of time, and the creation of the universe was allegedly the creation of time itself.

Therefore, defenders allege, the creation couldn’t have been a consequence of a simple law of nature.

3. Contingency Arguments

Cosmological arguments from contingency depend on the concepts of contingency and necessity. Contingent beings actually exist, but could have not existed. In contrast, a necessary being exists and could not possibly not have existed. The same terms describe truths or facts: A contingent fact is actually the case but could have not been the case, and a necessary truth is true, must be true and couldn’t have been false. [12]

Gottfried Leibniz (1646-1716 CE) developed the most famous contingency cosmological argument. [13] In summary:

  • Every contingent being needs an explanation for its existence, the cause or reason for its existence.
  • Contingent beings can’t be the cause or reason for their own existence.
  • Therefore, the explanation for the existence of contingent beings must be some non-contingent being; and since it’s non-contingent, that being must exist necessarily.
  • The most-plausible necessarily existing explanation for all the contingent beings would be God.

The support for (1) is a principle sometimes called the ‘Principle of Sufficient Reason,’ according to which things don’t just happen for no reason at all. [14]

The support for (2) is intuition or experience: it seems strange for a contingent being to somehow explain its own existence, and that certainly doesn’t happen in our experience.

Some critics challenge the Principle of Sufficient Reason. A few others argue that a being could potentially explain itself. [15]

One might also question (4). Perhaps this necessary being could be some sort of mindless entity or law of nature.

4. Conclusion

Cosmological arguments enlist both scientific and philosophical evidence to make their cases. The only general objection to all of them challenges the last step: to conclude that the ultimate explanation is God, and not something else.

To claim that the universe began to exist is counter-intuitive, but so is claiming that it never began to exist. It’s mysterious to think that God caused it and mysterious to imagine that there’s no first cause. Cosmological arguments attempt to make progress against these mysteries.

[1] Cosmological arguments are a posteriori arguments, meaning that the justification for at least one premise of the argument is empirical or sensory-based, in contrast to a priori arguments, which depend on thought and reflections alone for the justification for their premises. An example of an a priori argument for God’s existence is the ontological arguments; further examples of a posteriori arguments for God’s existence include design arguments. See The Ontological Argument for the Existence of God by Andrew Chapman, Design Arguments for the Existence of God by Tom Metcalf and Epistemology, or Theory of Knowledge by Tom Metcalf for further discussion of the a priori and a posteriori distinction. Note that philosophers generally regard cosmological arguments to be the most-persuasive arguments for the existence of God (PhilPapers n.d.).

[2] Aquinas (2017 [1920], I, q.2, a3).

[3] See Smith (1993: 121-23) on quantum mechanics, plus Leibniz’s Principle of Sufficient Reason by Marc Bobro.

[4] See Cameron (2021) on infinite regress arguments and see below at n. 9 for some discussion of whether actual infinities are possible.

[5] A version of this objection is sometimes stated in the form of a question: “What caused God?” Defenders of cosmological arguments typically argue that God, unlike the events in the universe, does not need a cause after all, perhaps because he is a necessary being, or did not begin to exist. See the following two types of cosmological arguments in this entry.

[6] See e.g. O’Connor (2004) and Rasmussen (2009) for attempts to bridge this gap.

[7] Craig (1979: 42); Richardson (2021). The argument goes back at least as far as al-Ghazālī (c. 1058-1111 CE). For more on Al-Ghazālī see also Al-Ghazālī’s Dream Argument by John Ramsey.

[8] Craig (1979: 65); see Reichenbach (2021: Section 7) for more-detailed discussion of these steps.

[9] See Oppy et al. (2021, Section 5.1). Here’s the example in more detail: If infinitely many people arrive and want rooms at the hotel, the hotel management can move all the existing guests into the room that’s double their current room number, thereby freeing up infinitely many new rooms without having to kick anyone out. As for infinitely many infinitely-large buses, imagine that each room at the hotel is numbered by a positive integer: 1, 2, 3, 4, and so on. Imagine that each of these infinitely large buses is numbered by a prime number greater than 2 as in 3, 5, 7, 11, and so on. And each passenger in each of the buses is numbered by a positive integer, as in 1, 2, 3, 4, and so on. Now when the infinitely many buses arrive, move the hotel guest in room 1 into room 2 1 = 2; the hotel guest in room 2 into room 2 2 = 4; the hotel guest in room 3 into room 2 3 = 8, and so on. Move the first passenger in bus #3 into room 3 1 = 3; the second passenger into room 3 2 = 9; the third passenger into room 3 3 = 27; and so on. Move the first passenger in bus #5 into room 5 1 = 5; the second passenger into room 5 2 = 25; the third passenger into room 5 3 = 125; and so on. Move the first passenger in bus #7 into room 7 1 = 7; the second passenger into room 7 2 = 49; and so on. The infinitely many infinitely-large buses’ passengers can all get rooms at the hotel without kicking anyone out. See Pires (2016) for more discussion.

[10] It may go back infinitely in time after all, or it may be wrong to consider it an event (Grünbaum 1991).

[11] See e.g. Craig (1994: 117) and compare Morriston (2000).

[12] See Possibility and Necessity: An Introduction to Modality by Andre Leo Rusavuk for more explanation of these concepts.

[13] Aquinas (2017 [1920], I, q.2, a3) offered a sort of contingency-based argument too. See also Gale and Pruss (1999).

[14] See Leibniz’s Principle of Sufficient Reason by Marc Bobro for an overview.

[15] Maybe there could be closed loops in time. See e.g. Lewis (1976) and Time Travel by Taylor Cyr, § 3.

Aquinas, T. (2017 [1920]). The Summa Theologiae of St. Thomas Aquinas (2 ed.). Ed. Kevin Knight. Tr. Fathers of the English Dominican Province.

Cameron, R. (2021). Infinite Regress Arguments. In E. N. Zalta (ed.), The Stanford Encyclopedia of Philosophy, Fall 2021 edition.

Craig, W. L. (1979). The Kalām Cosmological Argument . Eugene, OR: Wipf and Stock.

Craig, W. L. (1994). Reasonable Faith. Wheaton, IL: Crossway Books.

Gale, R. M. and Pruss, A. R. (1999). A new cosmological argument. Religious Studies, 35(4), 461-76.

Grünbaum, A. (1991). Creation as a Pseudo-Explanation in Current Physical Cosmology. Erkenntnis , 35(1), 233-254.

Lewis, D. (1976). The paradoxes of time travel. American Philosophical Quarterly , 13(2), 145-152.

Morriston, W. (2000). Must the Beginning of the Universe Have a Personal Cause? A Critical Examination of the Kalam Cosmological Argument. Faith and Philosophy , 17(2), 149-169.

O’Connor, T. (2004). “And This All Men Call God.” Faith and Philosophy , 21(4), 417-35.

Oppy, G.; Hájek, A.; Easwaran, K.; and Mancosu, P. (2021). Infinity. In E. N. Zalta (ed.), The Stanford Encyclopedia of Philosophy, Fall 2021 edition.

PhilPapers. (N.d.). Survey results. 2020 PhilPapers Survey .

Pires, A. (2016). Hospitality at the Grand Hilbert Hotel. The Institute Letter , Spring 2016.

Rasmussen, J. (2009). From a Necessary Being to God. International Journal for Philosophy of Religion , 66(1), 1-13.

Richardson, K. (2021). Causation in Arabic and Islamic Thought. In E. N. Zalta (ed.), The Stanford Encyclopedia of Philosophy, Fall 2021 edition.

Reichenbach, B. (2021). Cosmological Arguments. In E. N. Zalta (ed.), The Stanford Encyclopedia of Philosophy, Fall 2021 edition.

Smith, Q. (1993). The Uncaused Beginning of the Universe . In W. L. Craig and Q. Smith, Theism, Atheism, and Big Bang Cosmology (Oxford: Clarendon Press), 108-140.

Related Essays

The Concept of God: Divine Attributes by Bailie Peterson

Leibniz’s Principle of Sufficient Reason by Mark Bobro

The Ontological Argument for the Existence of God by Andrew Chapman

Design Arguments for the Existence of God by Tom Metcalf

Epistemology, or Theory of Knowledge by Tom Metcalf

Al-Ghazālī’s Dream Argument for Skepticism by John Ramsey

Possibility and Necessity: An Introduction to Modality by Andre Leo Rusavuk

Time Travel by Taylor W. Cyr

About the Author

Tom Metcalf is an associate professor at Spring Hill College in Mobile, AL. He received his PhD in philosophy from the University of Colorado, Boulder. He specializes in ethics, metaethics, epistemology, and the philosophy of religion. Tom has two cats whose names are Hesperus and Phosphorus. http://shc.academia.edu/ThomasMetcalf

Follow 1000-Word Philosophy on Facebook , Twitter and subscribe to receive email notifications of new essays at 1000WordPhilosophy.com

Share this:, 4 thoughts on “ cosmological arguments for the existence of god ”.

  • Pingback: Philosophy of Mysticism: Do Mystical Experiences Justify Religious Beliefs? – 1000-Word Philosophy: An Introductory Anthology
  • Pingback: “Properly Basic” Belief in God: Believing in God without an Argument – 1000-Word Philosophy: An Introductory Anthology
  • Pingback: Online Philosophy Resources Weekly Update | Daily Nous
  • Pingback: Leibniz’s Principle of Sufficient Reason – 1000-Word Philosophy: An Introductory Anthology

Comments are closed.

  • Search Menu

Sign in through your institution

  • Browse content in Arts and Humanities
  • Browse content in Archaeology
  • Anglo-Saxon and Medieval Archaeology
  • Archaeological Methodology and Techniques
  • Archaeology by Region
  • Archaeology of Religion
  • Archaeology of Trade and Exchange
  • Biblical Archaeology
  • Contemporary and Public Archaeology
  • Environmental Archaeology
  • Historical Archaeology
  • History and Theory of Archaeology
  • Industrial Archaeology
  • Landscape Archaeology
  • Mortuary Archaeology
  • Prehistoric Archaeology
  • Underwater Archaeology
  • Zooarchaeology
  • Browse content in Architecture
  • Architectural Structure and Design
  • History of Architecture
  • Residential and Domestic Buildings
  • Theory of Architecture
  • Browse content in Art
  • Art Subjects and Themes
  • History of Art
  • Industrial and Commercial Art
  • Theory of Art
  • Biographical Studies
  • Byzantine Studies
  • Browse content in Classical Studies
  • Classical Literature
  • Classical Reception
  • Classical History
  • Classical Philosophy
  • Classical Mythology
  • Classical Art and Architecture
  • Classical Oratory and Rhetoric
  • Greek and Roman Papyrology
  • Greek and Roman Archaeology
  • Greek and Roman Epigraphy
  • Greek and Roman Law
  • Late Antiquity
  • Religion in the Ancient World
  • Digital Humanities
  • Browse content in History
  • Colonialism and Imperialism
  • Diplomatic History
  • Environmental History
  • Genealogy, Heraldry, Names, and Honours
  • Genocide and Ethnic Cleansing
  • Historical Geography
  • History by Period
  • History of Emotions
  • History of Agriculture
  • History of Education
  • History of Gender and Sexuality
  • Industrial History
  • Intellectual History
  • International History
  • Labour History
  • Legal and Constitutional History
  • Local and Family History
  • Maritime History
  • Military History
  • National Liberation and Post-Colonialism
  • Oral History
  • Political History
  • Public History
  • Regional and National History
  • Revolutions and Rebellions
  • Slavery and Abolition of Slavery
  • Social and Cultural History
  • Theory, Methods, and Historiography
  • Urban History
  • World History
  • Browse content in Language Teaching and Learning
  • Language Learning (Specific Skills)
  • Language Teaching Theory and Methods
  • Browse content in Linguistics
  • Applied Linguistics
  • Cognitive Linguistics
  • Computational Linguistics
  • Forensic Linguistics
  • Grammar, Syntax and Morphology
  • Historical and Diachronic Linguistics
  • History of English
  • Language Evolution
  • Language Reference
  • Language Variation
  • Language Families
  • Language Acquisition
  • Lexicography
  • Linguistic Anthropology
  • Linguistic Theories
  • Linguistic Typology
  • Phonetics and Phonology
  • Psycholinguistics
  • Sociolinguistics
  • Translation and Interpretation
  • Writing Systems
  • Browse content in Literature
  • Bibliography
  • Children's Literature Studies
  • Literary Studies (Romanticism)
  • Literary Studies (American)
  • Literary Studies (Modernism)
  • Literary Studies (Asian)
  • Literary Studies (European)
  • Literary Studies (Eco-criticism)
  • Literary Studies - World
  • Literary Studies (1500 to 1800)
  • Literary Studies (19th Century)
  • Literary Studies (20th Century onwards)
  • Literary Studies (African American Literature)
  • Literary Studies (British and Irish)
  • Literary Studies (Early and Medieval)
  • Literary Studies (Fiction, Novelists, and Prose Writers)
  • Literary Studies (Gender Studies)
  • Literary Studies (Graphic Novels)
  • Literary Studies (History of the Book)
  • Literary Studies (Plays and Playwrights)
  • Literary Studies (Poetry and Poets)
  • Literary Studies (Postcolonial Literature)
  • Literary Studies (Queer Studies)
  • Literary Studies (Science Fiction)
  • Literary Studies (Travel Literature)
  • Literary Studies (War Literature)
  • Literary Studies (Women's Writing)
  • Literary Theory and Cultural Studies
  • Mythology and Folklore
  • Shakespeare Studies and Criticism
  • Browse content in Media Studies
  • Browse content in Music
  • Applied Music
  • Dance and Music
  • Ethics in Music
  • Ethnomusicology
  • Gender and Sexuality in Music
  • Medicine and Music
  • Music Cultures
  • Music and Media
  • Music and Culture
  • Music and Religion
  • Music Education and Pedagogy
  • Music Theory and Analysis
  • Musical Scores, Lyrics, and Libretti
  • Musical Structures, Styles, and Techniques
  • Musicology and Music History
  • Performance Practice and Studies
  • Race and Ethnicity in Music
  • Sound Studies
  • Browse content in Performing Arts
  • Browse content in Philosophy
  • Aesthetics and Philosophy of Art
  • Epistemology
  • Feminist Philosophy
  • History of Western Philosophy
  • Metaphysics
  • Moral Philosophy
  • Non-Western Philosophy
  • Philosophy of Language
  • Philosophy of Mind
  • Philosophy of Perception
  • Philosophy of Action
  • Philosophy of Law
  • Philosophy of Religion
  • Philosophy of Science
  • Philosophy of Mathematics and Logic
  • Practical Ethics
  • Social and Political Philosophy
  • Browse content in Religion
  • Biblical Studies
  • Christianity
  • East Asian Religions
  • History of Religion
  • Judaism and Jewish Studies
  • Qumran Studies
  • Religion and Education
  • Religion and Health
  • Religion and Politics
  • Religion and Science
  • Religion and Law
  • Religion and Art, Literature, and Music
  • Religious Studies
  • Browse content in Society and Culture
  • Cookery, Food, and Drink
  • Cultural Studies
  • Customs and Traditions
  • Ethical Issues and Debates
  • Hobbies, Games, Arts and Crafts
  • Natural world, Country Life, and Pets
  • Popular Beliefs and Controversial Knowledge
  • Sports and Outdoor Recreation
  • Technology and Society
  • Travel and Holiday
  • Visual Culture
  • Browse content in Law
  • Arbitration
  • Browse content in Company and Commercial Law
  • Commercial Law
  • Company Law
  • Browse content in Comparative Law
  • Systems of Law
  • Competition Law
  • Browse content in Constitutional and Administrative Law
  • Government Powers
  • Judicial Review
  • Local Government Law
  • Military and Defence Law
  • Parliamentary and Legislative Practice
  • Construction Law
  • Contract Law
  • Browse content in Criminal Law
  • Criminal Procedure
  • Criminal Evidence Law
  • Sentencing and Punishment
  • Employment and Labour Law
  • Environment and Energy Law
  • Browse content in Financial Law
  • Banking Law
  • Insolvency Law
  • History of Law
  • Human Rights and Immigration
  • Intellectual Property Law
  • Browse content in International Law
  • Private International Law and Conflict of Laws
  • Public International Law
  • IT and Communications Law
  • Jurisprudence and Philosophy of Law
  • Law and Society
  • Law and Politics
  • Browse content in Legal System and Practice
  • Courts and Procedure
  • Legal Skills and Practice
  • Primary Sources of Law
  • Regulation of Legal Profession
  • Medical and Healthcare Law
  • Browse content in Policing
  • Criminal Investigation and Detection
  • Police and Security Services
  • Police Procedure and Law
  • Police Regional Planning
  • Browse content in Property Law
  • Personal Property Law
  • Study and Revision
  • Terrorism and National Security Law
  • Browse content in Trusts Law
  • Wills and Probate or Succession
  • Browse content in Medicine and Health
  • Browse content in Allied Health Professions
  • Arts Therapies
  • Clinical Science
  • Dietetics and Nutrition
  • Occupational Therapy
  • Operating Department Practice
  • Physiotherapy
  • Radiography
  • Speech and Language Therapy
  • Browse content in Anaesthetics
  • General Anaesthesia
  • Neuroanaesthesia
  • Clinical Neuroscience
  • Browse content in Clinical Medicine
  • Acute Medicine
  • Cardiovascular Medicine
  • Clinical Genetics
  • Clinical Pharmacology and Therapeutics
  • Dermatology
  • Endocrinology and Diabetes
  • Gastroenterology
  • Genito-urinary Medicine
  • Geriatric Medicine
  • Infectious Diseases
  • Medical Toxicology
  • Medical Oncology
  • Pain Medicine
  • Palliative Medicine
  • Rehabilitation Medicine
  • Respiratory Medicine and Pulmonology
  • Rheumatology
  • Sleep Medicine
  • Sports and Exercise Medicine
  • Community Medical Services
  • Critical Care
  • Emergency Medicine
  • Forensic Medicine
  • Haematology
  • History of Medicine
  • Browse content in Medical Skills
  • Clinical Skills
  • Communication Skills
  • Nursing Skills
  • Surgical Skills
  • Medical Ethics
  • Browse content in Medical Dentistry
  • Oral and Maxillofacial Surgery
  • Paediatric Dentistry
  • Restorative Dentistry and Orthodontics
  • Surgical Dentistry
  • Medical Statistics and Methodology
  • Browse content in Neurology
  • Clinical Neurophysiology
  • Neuropathology
  • Nursing Studies
  • Browse content in Obstetrics and Gynaecology
  • Gynaecology
  • Occupational Medicine
  • Ophthalmology
  • Otolaryngology (ENT)
  • Browse content in Paediatrics
  • Neonatology
  • Browse content in Pathology
  • Chemical Pathology
  • Clinical Cytogenetics and Molecular Genetics
  • Histopathology
  • Medical Microbiology and Virology
  • Patient Education and Information
  • Browse content in Pharmacology
  • Psychopharmacology
  • Browse content in Popular Health
  • Caring for Others
  • Complementary and Alternative Medicine
  • Self-help and Personal Development
  • Browse content in Preclinical Medicine
  • Cell Biology
  • Molecular Biology and Genetics
  • Reproduction, Growth and Development
  • Primary Care
  • Professional Development in Medicine
  • Browse content in Psychiatry
  • Addiction Medicine
  • Child and Adolescent Psychiatry
  • Forensic Psychiatry
  • Learning Disabilities
  • Old Age Psychiatry
  • Psychotherapy
  • Browse content in Public Health and Epidemiology
  • Epidemiology
  • Public Health
  • Browse content in Radiology
  • Clinical Radiology
  • Interventional Radiology
  • Nuclear Medicine
  • Radiation Oncology
  • Reproductive Medicine
  • Browse content in Surgery
  • Cardiothoracic Surgery
  • Gastro-intestinal and Colorectal Surgery
  • General Surgery
  • Neurosurgery
  • Paediatric Surgery
  • Peri-operative Care
  • Plastic and Reconstructive Surgery
  • Surgical Oncology
  • Transplant Surgery
  • Trauma and Orthopaedic Surgery
  • Vascular Surgery
  • Browse content in Science and Mathematics
  • Browse content in Biological Sciences
  • Aquatic Biology
  • Biochemistry
  • Bioinformatics and Computational Biology
  • Developmental Biology
  • Ecology and Conservation
  • Evolutionary Biology
  • Genetics and Genomics
  • Microbiology
  • Molecular and Cell Biology
  • Natural History
  • Plant Sciences and Forestry
  • Research Methods in Life Sciences
  • Structural Biology
  • Systems Biology
  • Zoology and Animal Sciences
  • Browse content in Chemistry
  • Analytical Chemistry
  • Computational Chemistry
  • Crystallography
  • Environmental Chemistry
  • Industrial Chemistry
  • Inorganic Chemistry
  • Materials Chemistry
  • Medicinal Chemistry
  • Mineralogy and Gems
  • Organic Chemistry
  • Physical Chemistry
  • Polymer Chemistry
  • Study and Communication Skills in Chemistry
  • Theoretical Chemistry
  • Browse content in Computer Science
  • Artificial Intelligence
  • Computer Architecture and Logic Design
  • Game Studies
  • Human-Computer Interaction
  • Mathematical Theory of Computation
  • Programming Languages
  • Software Engineering
  • Systems Analysis and Design
  • Virtual Reality
  • Browse content in Computing
  • Business Applications
  • Computer Games
  • Computer Security
  • Computer Networking and Communications
  • Digital Lifestyle
  • Graphical and Digital Media Applications
  • Operating Systems
  • Browse content in Earth Sciences and Geography
  • Atmospheric Sciences
  • Environmental Geography
  • Geology and the Lithosphere
  • Maps and Map-making
  • Meteorology and Climatology
  • Oceanography and Hydrology
  • Palaeontology
  • Physical Geography and Topography
  • Regional Geography
  • Soil Science
  • Urban Geography
  • Browse content in Engineering and Technology
  • Agriculture and Farming
  • Biological Engineering
  • Civil Engineering, Surveying, and Building
  • Electronics and Communications Engineering
  • Energy Technology
  • Engineering (General)
  • Environmental Science, Engineering, and Technology
  • History of Engineering and Technology
  • Mechanical Engineering and Materials
  • Technology of Industrial Chemistry
  • Transport Technology and Trades
  • Browse content in Environmental Science
  • Applied Ecology (Environmental Science)
  • Conservation of the Environment (Environmental Science)
  • Environmental Sustainability
  • Environmentalist Thought and Ideology (Environmental Science)
  • Management of Land and Natural Resources (Environmental Science)
  • Natural Disasters (Environmental Science)
  • Nuclear Issues (Environmental Science)
  • Pollution and Threats to the Environment (Environmental Science)
  • Social Impact of Environmental Issues (Environmental Science)
  • History of Science and Technology
  • Browse content in Materials Science
  • Ceramics and Glasses
  • Composite Materials
  • Metals, Alloying, and Corrosion
  • Nanotechnology
  • Browse content in Mathematics
  • Applied Mathematics
  • Biomathematics and Statistics
  • History of Mathematics
  • Mathematical Education
  • Mathematical Finance
  • Mathematical Analysis
  • Numerical and Computational Mathematics
  • Probability and Statistics
  • Pure Mathematics
  • Browse content in Neuroscience
  • Cognition and Behavioural Neuroscience
  • Development of the Nervous System
  • Disorders of the Nervous System
  • History of Neuroscience
  • Invertebrate Neurobiology
  • Molecular and Cellular Systems
  • Neuroendocrinology and Autonomic Nervous System
  • Neuroscientific Techniques
  • Sensory and Motor Systems
  • Browse content in Physics
  • Astronomy and Astrophysics
  • Atomic, Molecular, and Optical Physics
  • Biological and Medical Physics
  • Classical Mechanics
  • Computational Physics
  • Condensed Matter Physics
  • Electromagnetism, Optics, and Acoustics
  • History of Physics
  • Mathematical and Statistical Physics
  • Measurement Science
  • Nuclear Physics
  • Particles and Fields
  • Plasma Physics
  • Quantum Physics
  • Relativity and Gravitation
  • Semiconductor and Mesoscopic Physics
  • Browse content in Psychology
  • Affective Sciences
  • Clinical Psychology
  • Cognitive Psychology
  • Cognitive Neuroscience
  • Criminal and Forensic Psychology
  • Developmental Psychology
  • Educational Psychology
  • Evolutionary Psychology
  • Health Psychology
  • History and Systems in Psychology
  • Music Psychology
  • Neuropsychology
  • Organizational Psychology
  • Psychological Assessment and Testing
  • Psychology of Human-Technology Interaction
  • Psychology Professional Development and Training
  • Research Methods in Psychology
  • Social Psychology
  • Browse content in Social Sciences
  • Browse content in Anthropology
  • Anthropology of Religion
  • Human Evolution
  • Medical Anthropology
  • Physical Anthropology
  • Regional Anthropology
  • Social and Cultural Anthropology
  • Theory and Practice of Anthropology
  • Browse content in Business and Management
  • Business Ethics
  • Business History
  • Business Strategy
  • Business and Technology
  • Business and Government
  • Business and the Environment
  • Comparative Management
  • Corporate Governance
  • Corporate Social Responsibility
  • Entrepreneurship
  • Health Management
  • Human Resource Management
  • Industrial and Employment Relations
  • Industry Studies
  • Information and Communication Technologies
  • International Business
  • Knowledge Management
  • Management and Management Techniques
  • Operations Management
  • Organizational Theory and Behaviour
  • Pensions and Pension Management
  • Public and Nonprofit Management
  • Strategic Management
  • Supply Chain Management
  • Browse content in Criminology and Criminal Justice
  • Criminal Justice
  • Criminology
  • Forms of Crime
  • International and Comparative Criminology
  • Youth Violence and Juvenile Justice
  • Development Studies
  • Browse content in Economics
  • Agricultural, Environmental, and Natural Resource Economics
  • Asian Economics
  • Behavioural Finance
  • Behavioural Economics and Neuroeconomics
  • Econometrics and Mathematical Economics
  • Economic History
  • Economic Methodology
  • Economic Systems
  • Economic Development and Growth
  • Financial Markets
  • Financial Institutions and Services
  • General Economics and Teaching
  • Health, Education, and Welfare
  • History of Economic Thought
  • International Economics
  • Labour and Demographic Economics
  • Law and Economics
  • Macroeconomics and Monetary Economics
  • Microeconomics
  • Public Economics
  • Urban, Rural, and Regional Economics
  • Welfare Economics
  • Browse content in Education
  • Adult Education and Continuous Learning
  • Care and Counselling of Students
  • Early Childhood and Elementary Education
  • Educational Equipment and Technology
  • Educational Strategies and Policy
  • Higher and Further Education
  • Organization and Management of Education
  • Philosophy and Theory of Education
  • Schools Studies
  • Secondary Education
  • Teaching of a Specific Subject
  • Teaching of Specific Groups and Special Educational Needs
  • Teaching Skills and Techniques
  • Browse content in Environment
  • Applied Ecology (Social Science)
  • Climate Change
  • Conservation of the Environment (Social Science)
  • Environmentalist Thought and Ideology (Social Science)
  • Natural Disasters (Environment)
  • Social Impact of Environmental Issues (Social Science)
  • Browse content in Human Geography
  • Cultural Geography
  • Economic Geography
  • Political Geography
  • Browse content in Interdisciplinary Studies
  • Communication Studies
  • Museums, Libraries, and Information Sciences
  • Browse content in Politics
  • African Politics
  • Asian Politics
  • Chinese Politics
  • Comparative Politics
  • Conflict Politics
  • Elections and Electoral Studies
  • Environmental Politics
  • Ethnic Politics
  • European Union
  • Foreign Policy
  • Gender and Politics
  • Human Rights and Politics
  • Indian Politics
  • International Relations
  • International Organization (Politics)
  • International Political Economy
  • Irish Politics
  • Latin American Politics
  • Middle Eastern Politics
  • Political Behaviour
  • Political Economy
  • Political Institutions
  • Political Theory
  • Political Methodology
  • Political Communication
  • Political Philosophy
  • Political Sociology
  • Politics and Law
  • Politics of Development
  • Public Policy
  • Public Administration
  • Quantitative Political Methodology
  • Regional Political Studies
  • Russian Politics
  • Security Studies
  • State and Local Government
  • UK Politics
  • US Politics
  • Browse content in Regional and Area Studies
  • African Studies
  • Asian Studies
  • East Asian Studies
  • Japanese Studies
  • Latin American Studies
  • Middle Eastern Studies
  • Native American Studies
  • Scottish Studies
  • Browse content in Research and Information
  • Research Methods
  • Browse content in Social Work
  • Addictions and Substance Misuse
  • Adoption and Fostering
  • Care of the Elderly
  • Child and Adolescent Social Work
  • Couple and Family Social Work
  • Direct Practice and Clinical Social Work
  • Emergency Services
  • Human Behaviour and the Social Environment
  • International and Global Issues in Social Work
  • Mental and Behavioural Health
  • Social Justice and Human Rights
  • Social Policy and Advocacy
  • Social Work and Crime and Justice
  • Social Work Macro Practice
  • Social Work Practice Settings
  • Social Work Research and Evidence-based Practice
  • Welfare and Benefit Systems
  • Browse content in Sociology
  • Childhood Studies
  • Community Development
  • Comparative and Historical Sociology
  • Economic Sociology
  • Gender and Sexuality
  • Gerontology and Ageing
  • Health, Illness, and Medicine
  • Marriage and the Family
  • Migration Studies
  • Occupations, Professions, and Work
  • Organizations
  • Population and Demography
  • Race and Ethnicity
  • Social Theory
  • Social Movements and Social Change
  • Social Research and Statistics
  • Social Stratification, Inequality, and Mobility
  • Sociology of Religion
  • Sociology of Education
  • Sport and Leisure
  • Urban and Rural Studies
  • Browse content in Warfare and Defence
  • Defence Strategy, Planning, and Research
  • Land Forces and Warfare
  • Military Administration
  • Military Life and Institutions
  • Naval Forces and Warfare
  • Other Warfare and Defence Issues
  • Peace Studies and Conflict Resolution
  • Weapons and Equipment

The Existence of God (1st edn)

A newer edition of this book is available.

  • < Previous chapter
  • Next chapter >

7 The Cosmological Argument

Author Webpage

  • Published: March 1991
  • Cite Icon Cite
  • Permissions Icon Permissions

The cosmological argument is the argument to God from the existence of a complex physical universe. God can create such a universe (by keeping it in existence as long as it exists ‐ whether for a finite or infinite time) and has some reason to do so because it is a theatre for finite agents which they can shape and in which they can develop. But it is a far less simple beginning of things than is God, and so a priori not to be expected but for the action of God.

Signed in as

Institutional accounts.

  • Google Scholar Indexing
  • GoogleCrawler [DO NOT DELETE]

Personal account

  • Sign in with email/username & password
  • Get email alerts
  • Save searches
  • Purchase content
  • Activate your purchase/trial code
  • Add your ORCID iD

Institutional access

Sign in with a library card.

  • Sign in with username/password
  • Recommend to your librarian
  • Institutional account management
  • Get help with access

Access to content on Oxford Academic is often provided through institutional subscriptions and purchases. If you are a member of an institution with an active account, you may be able to access content in one of the following ways:

IP based access

Typically, access is provided across an institutional network to a range of IP addresses. This authentication occurs automatically, and it is not possible to sign out of an IP authenticated account.

Choose this option to get remote access when outside your institution. Shibboleth/Open Athens technology is used to provide single sign-on between your institution’s website and Oxford Academic.

  • Click Sign in through your institution.
  • Select your institution from the list provided, which will take you to your institution's website to sign in.
  • When on the institution site, please use the credentials provided by your institution. Do not use an Oxford Academic personal account.
  • Following successful sign in, you will be returned to Oxford Academic.

If your institution is not listed or you cannot sign in to your institution’s website, please contact your librarian or administrator.

Enter your library card number to sign in. If you cannot sign in, please contact your librarian.

Society Members

Society member access to a journal is achieved in one of the following ways:

Sign in through society site

Many societies offer single sign-on between the society website and Oxford Academic. If you see ‘Sign in through society site’ in the sign in pane within a journal:

  • Click Sign in through society site.
  • When on the society site, please use the credentials provided by that society. Do not use an Oxford Academic personal account.

If you do not have a society account or have forgotten your username or password, please contact your society.

Sign in using a personal account

Some societies use Oxford Academic personal accounts to provide access to their members. See below.

A personal account can be used to get email alerts, save searches, purchase content, and activate subscriptions.

Some societies use Oxford Academic personal accounts to provide access to their members.

Viewing your signed in accounts

Click the account icon in the top right to:

  • View your signed in personal account and access account management features.
  • View the institutional accounts that are providing access.

Signed in but can't access content

Oxford Academic is home to a wide variety of products. The institutional subscription may not cover the content that you are trying to access. If you believe you should have access to that content, please contact your librarian.

For librarians and administrators, your personal account also provides access to institutional account management. Here you will find options to view and activate subscriptions, manage institutional settings and access options, access usage statistics, and more.

Our books are available by subscription or purchase to libraries and institutions.

  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Rights and permissions
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

The Cosmological Argument Exploratory Essay

  • To find inspiration for your paper and overcome writer’s block
  • As a source of information (ensure proper referencing)
  • As a template for you assignment

The Cosmological argument is defined as the argument which proves that the universe was created by God. This argument clearly claims that the world is in existence because it was created by God. The first arguments are said to have been generated by Aristotle around three hundred years before Christ.

In these arguments, it is well explained that everything that is in existence has a cause. This means that if one was to decide to go back far enough, a cause would be determined. It is also said that the first initial cause was God. The cosmological argument was initially developed by philosophers of Muslim decent in the middle ages. On a day to day basis, it seems like everything comes in and out of existence, depending on their coming to existence and also the fact that they might have not existed at all.

Seeking an explanation to these kinds of things is not in any way improper. God is defined as the Supreme Being (Oddie para 2). Unlike other beings, God is not one to come into existence and also fade out of existence. God is termed as being an eternal being. God is not dependent on any other beings for existence thus he is termed as an Independent Being. God’s existence is not a matter of luck or chancing unlike every other being that are a matter of chance.

Hence God is a necessary Being. In addition, God can be said to be a necessary, independent, and eternal Being who the rest of the universe or cosmos depends on for their well being. God is regarded as a creator and sustainer of lives and the universe. Among the many arguments surrounding this topic is that there cannot be a true infinite. This is because even a small part of what is referred to as infinite is itself seen as an infinite since it has been derived from the entire infinite.

For example, an infinite collection of the blue and white marbles would be best suited to explain this theory. The number of blue marbles in this pack is directly proportional to the total number of all marbles in the pack, because both are infinite. The same holds for the number of marbles in the pack.

Thus, the number of blue marbles equals the number of white marbles which equals the sum of all blue and white marbles. It would be incongruous to hold to the claims of existence of a true infinite. Similarly, thinking that infinite happenings can occur prior to a given time is awkward.

The second argument talks about the formation of actual infinite. It states that it cannot be formed. It is fascinating that philosopher Craig also states that the cause of the universe is as a result of the Personal Creator. He goes on to state that “the only way to have an eternal cause but a temporal effect would seem to be if the cause is a personal agent who freely chooses to create an effect in time” (Craig p. 88)

The creation theory has been defined numerous times and there are a number of different stories telling on how the earth was formed. The evolution theory is somewhat believable, probably by pagans and evolutionists. Most communities have come up with theories in a bid to explain their existence.

This is most common in African cultures where they have learned to coin theories explaining their existence. Christians believe in the existence of a God who is considered to have willed the existence of all that there is. This God is not a physical being that is part of the universe. God’s mighty power rules over the entire universe.

The Bible clearly shows us that God is indeed the uncaused First Cause that created the universe by willing it into existence. Another cosmological argument that is consistent with the Bible is the Kalam Cosmological Argument. Aristotle’s theory was of great importance in the cosmological arguments about the existence of God.

He argued that everything that takes place is as a result of something else. For example, if one was to slip on a banana peel and fall, there would be the question of who left the peel there? Why was the peel left there? Where did it come from? This can be traced back as far as one would want. Aristotle thought that it would have been easier by reaching only the first cause which is causing but itself uncaused. This theory is now known as the prime Mover.

Creation is the only widely accepted theory on the existence of the universe. It may vary depending on one’s religion but it all leads to the same, one natural and Supreme Being. The Christians believe in God and his son Jesus Christ who came to earth in humanly form and ascended into heaven.

Their holy book is known as The Bible. The Muslims believe in Allah and Mohamed his Prophet, their holy book is known as The Koran. The Hindus believe in a number of gods but there is a supreme being they believe in and they refer to their holy book as the Bahagdvita. These are just the main religions in the world and a little of their beliefs. It is evident that God is a super natural Being.

Works Cited

Craig, William Lane. “The existence of God and the beginning of the universe.” Truth: A Journal of Modern Thought 3 (1991): 85-96.

Oddie Graham. The cosmological argument: Why is there something rather than nothing? 2011. Web.

  • Sir Robert Peel: The Impact on American Policing
  • The God's Existence: Cosmological Proof
  • The Ontological Argument for the Existence of God
  • Post Structuralism in Modern Day Society
  • Philosophical Analysis on the Death of Osama Bin Laden
  • Procrastination Issues: Cause and Effect
  • Integrative Philosophical Principles
  • Human Nature as a Power to Make Choices
  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2018, October 17). The Cosmological Argument. https://ivypanda.com/essays/the-cosmological-argument/

"The Cosmological Argument." IvyPanda , 17 Oct. 2018, ivypanda.com/essays/the-cosmological-argument/.

IvyPanda . (2018) 'The Cosmological Argument'. 17 October.

IvyPanda . 2018. "The Cosmological Argument." October 17, 2018. https://ivypanda.com/essays/the-cosmological-argument/.

1. IvyPanda . "The Cosmological Argument." October 17, 2018. https://ivypanda.com/essays/the-cosmological-argument/.

Bibliography

IvyPanda . "The Cosmological Argument." October 17, 2018. https://ivypanda.com/essays/the-cosmological-argument/.

A Level Philosophy & Religious Studies

The Cosmological argument summary notes

OCR Philosophy

This page contains summary revision notes for the Cosmological argument topic. There are two versions of these notes. Click on the A*-A grade tab, or the B-C grade tab, depending on the grade you are trying to get.

Find the full revision page  here.

  • Cosmological arguments attempt to show that God exists as the required explanation of the existence of the universe.
  • Aquinas’ 3 ways are 3 varieties of the cosmological argument.
  • A posteriori & inductive argument.

Aquinas’ 1st way (from motion): 

  • We observe that there is motion
  • Nothing moves itself – something only moves when moved by something else
  • There can’t be an infinite regress of movers
  • So, there must have been a first mover that was itself unmoved – that thing we call God.

Aquinas’ 2nd way (from causation):

  • We observe that there is cause and effect in the world
  • Nothing can cause itself – something is only caused when caused by something else.
  • There can’t be an infinite regress of causation.
  • So, there must have been a first cause that itself uncaused – that thing we call God.

Hume’s objection to the causal principle  

  • This especially attacks Aquinas’ 1st and 2nd way. 
  • The causal principle is the claim that every event has a cause. 
  • Aquinas’ 1st and 2nd way clearly assume that this causal principle is true – they assume that everything has a mover or cause.
  • However, Hume casts doubt on that. Firstly he says it cannot be an analytic truth (true by definition).
  • If you deny an analytic truth – you contradict yourself. If you say a triangle has four sides you contradict yourself because the idea of ‘triangle’ is contradicted by the idea of ‘four sides’. So, ‘a triangle has three sides’ is an analytic truth because it cannot be denied without contradiction.
  • But – if you say there could be an event without a cause – this doesn’t seem self-contradictory. The idea of ‘event’ doesn’t seem contradicted by the idea of ‘no cause’. E.g. you could imagine nothingness and then – pop – suddenly something is there without a cause. 
  • So, we cannot know that the causal principle is analytically true – we can’t know it is absolutely true in all cases. It could simply be that every event we have ever observed has a cause – but this doesn’t mean that every event, including the creation of the universe, has a cause.
  • If the universe has no cause – then God can’t be argued for as the required explanation of the causation (or motion) of the universe.
  • Adding modern science to Hume’s point: quantum mechanics

Evaluation : 

  • What if we accepted Hume’s claim that the causal principle is not an analytic truth, but instead argued that it was a synthetic truth – 
  • The causal principle is an empirical (scientific) hypothesis which has so far been justified by all the evidence we have ever observed. 
  • This makes the causal principle more reasonable to believe in than to not believe in. 
  • So, the cosmological argument would be more reasonable to accept than to not accept. 
  • Maybe we can’t prove the causal principle for absolutely certain as a matter of logic – but so far all the evidence supports it (every time we observe an event, it has a cause) so, arguably we are justified in accepting it.
  • Just accepting that the causal principle is more reasonable to believe that not is enough for the cosmological argument to be convincing, because it is an inductive argument attempting to provide evidence for God, not a deductive argument attempting to provide a logical proof of God.

Optional further evaluation:

  • The fact that every event we observe in the universe has a cause is not even evidence for the universe having a cause.
  • We have no basis on which to think that the conditions within the universe are like the conditions under which the universe came into existence.
  • So, empirical evidence of cause and effect within the universe does not make it reasonable to believe the causal principle.

Aquinas’ 3rd way (from contingency): 

  • The third way involve the terms contingent and necessary
  • A contingent being depends on something else for its existence – it can either exist or not exist.
  • A necessary being does not depend on anything else for its existence – it must exist.
  • There cannot be an infinite regress of contingent beings
  • If everything we see in the universe forms a chain of contingent beings – then ‘before’ this series, there would have been nothing. If everything is on that chain of contingent beings – then before it would have to be nothing.
  • But – how could this chain of contingent beings have come from nothing..? That’s impossible.
  • So, there must be a necessary being which began this series of contingent beings – that thing we call God.

Evaluation:

  • Hume’s critique of the idea of a necessary being:
  • A ‘necessary being’ is actually a meaningless concept. 
  • A necessary being is one which must exist. 
  • In that case, we shouldn’t even be able to imagine it not existing. 
  • However, Hume claims that ‘whatever we can imagine existing, we can imagine not existing’. Hume claims we can imagine God not existing – so it must therefore be possible for God to not exist, and this means it cannot be the case that God must exist. 
  • So, the very idea of a necessary being is absurd and meaningless – because there is no being which we are unable to imagine not existing.

Further evaluation

  • The masked man fallacy. 
  • Hume assumes that we cannot imagine impossible things. 
  • The masked man fallacy shows that we can. A person could hear of a bank robbery by a masked man – and they could imagine that it is not their father. However, if it was their father – then it’s impossible for it to not be their father. Yet, that is what they conceived – so, we can conceive of the impossible.
  • In that case, when Hume conceived of God not existing and claimed this meant God’s non-existence was ‘possible’ – Hume was wrong. 
  • God’s non-existence could be an example of something that is conceivable but impossible – as the masked man fallacy illustrates.

The issue of the infinite regress

  • All cosmological arguments rely on the premise that an infinite regress is impossible. 
  • This is because if the universe has always existed (in some form) then God’s existence cannot be argued for as the required explanation of the creation of the universe, since then it would not have been created – it would have just always existed.
  • Perhaps the universe has just always been here, in which case God didn’t cause it.
  • Hume argues that an infinite regress is possible. For something to be impossible, it has to be self-contradictory. The idea of events/things doesn’t seem contradicted by the idea that they go back in time forever. So, it seems possible for things to go back in time forever. 
  • Adding modern science to Hume’s point: E.g. Maybe the universe has been expanding and then crunching forever.
  • So, the cosmological argument rests on an assumption.

Evaluation: 

  • W. L. Craig defended the cosmological argument from this criticism, using ‘Hilbert’s Hotel’ – a hotel with an infinite number of rooms that are all full. A person shows up wanting a room, so reception puts everyone into their room number plus one. This makes room number 1 free. Then an infinite number of people show up to the hotel – but this is also no problem, since the receptionist can just put everyone into the room twice the number of their current room. This makes an infinite number of odd-numbered rooms free.
  • Craigs point about this – is that surely such a Hotel could never actually exist in reality. It’s absurd to think that infinities could actually exist. If a hotel is full, it can’t make more rooms free – but if it were an infinite hotel – then it could in fact make an infinite number of room free. This just defies all logic regarding how physical reality could actually work. Infinities cannot be possible in reality.

Further evaluation:

  • However, Craig’s argument focuses on the impossibility of an infinity of physical things, but it fails to target the idea of a temporal infinite – an infinite amount of time, which is what the infinite regress requires.
  • Aquinas has a better argument than Craig.
  • If there were an infinite regress, there would be an infinite amount of time before the present moment.
  • In that case, to get to the present moment, an infinite amount of time would have had to have passed.
  • However, an infinite amount of time cannot pass – no matter how long you wait, it’s not possible to wait for an infinite amount of time.
  • You cannot traverse an infinite through successive addition.
  • So, there cannot be an infinite amount of time before the present moment, and so an infinite regress is not possible.

Aquinas’ 1st and 2nd ways

  • Aquinas created three versions of the cosmological argument.
  • 1st way: motion
  • Everything is in motion
  • There can’t be an infinite regress of motion. It cannot be that there is just an infinite chain of movers going back in time forever. There has to have been a first mover – a start to the motion we observe.
  • E.g. if you see dominoes falling, there must have been a first one that was pushed – there couldn’t have just been dominoes falling forever.
  • So, there must have been a first mover that was unmoved – that is God.
  • 2nd way: causation
  • Everything has a cause
  • There can’t be an infinite regress of causation – there can’t be an infinite chain of cause and effect going back in time forever. 
  • There has to be a first cause which is uncaused – that is God.

Hume’s critique of the cosmological argument: objection to the causal principle

  • The causal principle is the claim that everything has a cause.
  • The cosmological argument assumes that everything has a cause – it assumes the universe must have a cause.
  • But Hume claims it’s possible the universe had no cause.

Evaluation: criticism of Hume’s objection to the causal principles

  • However – it seems scientifically justified to think everything has a cause `
  • Everything we’ve ever seen has a cause. 
  • So, it’s empirically more reasonable to think the universe also has a cause than to think it doesn’t.
  • The cosmological argument is therefore still convincing.

Aquinas’ 3rd way: contingency

  • A contingent being is one which depends on something else for its existence.
  • Everything we see is contingent.
  • There can’t have been an infinite regress of contingent beings – one creating the next, going back in time forever.
  • So, there must have been a first contingent being, but that can’t come from nothing, so there must have been a necessary being which created it.
  • That necessary being is God. (A necessary being does not depend on anything else for its existence).

Hume & Russell’s critique of the cosmological argument: fallacy of composition

  • The fallacy of composition states that just because something is true of the parts, doesn’t mean it is true of the whole.
  • E.g. just because every human has a mother, it doesn’t mean the whole human race itself has a mother.
  • Similarly – just because all the parts of the universe have a mover/cause/contingency – that doesn’t mean the whole universe itself has a mover/cause/is-contingent. 
  • If the universe has no cause/mover/contingency, then there is no need for a God to explain its existence.

Evaluation: criticism of the fallacy of composition

  • Leibniz points out that It’s illogical for something to happen without a reason/cause.
  • The universe must have a cause – nothing comes from nothing. 
  • Something only comes from something.
  • So, the universe must have come from some cause – God.

Hume’s critique of the cosmological argument: the possibility of an infinite series

  • Maybe an infinite regress actually is possible.
  • For something to be impossible it has to be logically self-contradictory.
  • But there doesn’t appear to be anything illogical or nonsensical about things going back in time forever.

Evaluation: criticism of the possibility of an infinite series

  • Dominoes example – shows things must have a beginning.
  • If you see dominoes falling – there must have been a first one which was pushed.
  • Similarly, if you see things in the universe changing, being caused or being contingent – there must have been a first one which started the process off but was not caused, moved or contingent.

Home — Essay Samples — Religion — Faith — Cosmological Argument: St. Thomas Aquinas

test_template

Cosmological Argument: St. Thomas Aquinas

  • Categories: Faith God

About this sample

close

Words: 627 |

Published: Feb 12, 2019

Words: 627 | Page: 1 | 4 min read

Image of Dr. Charlotte Jacobson

Cite this Essay

Let us write you an essay from scratch

  • 450+ experts on 30 subjects ready to help
  • Custom essay delivered in as few as 3 hours

Get high-quality help

author

Dr. Karlyna PhD

Verified writer

  • Expert in: Religion

writer

+ 120 experts online

By clicking “Check Writers’ Offers”, you agree to our terms of service and privacy policy . We’ll occasionally send you promo and account related email

No need to pay just yet!

Related Essays

2 pages / 685 words

2 pages / 908 words

3 pages / 1255 words

4 pages / 2046 words

Remember! This is just a sample.

You can get your custom paper by one of our expert writers.

121 writers online

Cosmological Argument: St. Thomas Aquinas Essay

Still can’t find what you need?

Browse our vast selection of original essay samples, each expertly formatted and styled

Related Essays on Faith

The Bible. New International Version. Bible Gateway, www.biblegateway.com.

We all encounter different situations where we build up negative issues and feel the need to clear the clutter out in order to grow. For that, we all need some help from the people who surround us. This clearing process is what [...]

Keller, H. (n.d.). Helen Keller Quotes. BrainyQuote. Retrieved from https://www.merriam-webster.com/dictionary/faith

Night Faith Quotes: Exploring the Depths of Human SpiritualityIn Elie Wiesel's haunting memoir, Night, faith emerges as both a source of solace and a catalyst for profound existential questioning amidst the horrors of the [...]

Drawing reference from indigenous traditions and stories, there are many aspects of life that one can learn about regarding religion and theology. As much as some cases such as the Eagle boy’s case may be fictional, they open up [...]

The work of Thomas Aquinas, though somewhat insignificant in his own day, is arguably some of the most studied, discussed, and revered to emerge from the medieval period. As Plantinga, Thompson and Lundberg maintain, 'of all the [...]

Related Topics

By clicking “Send”, you agree to our Terms of service and Privacy statement . We will occasionally send you account related emails.

Where do you want us to send this sample?

By clicking “Continue”, you agree to our terms of service and privacy policy.

Be careful. This essay is not unique

This essay was donated by a student and is likely to have been used and submitted before

Download this Sample

Free samples may contain mistakes and not unique parts

Sorry, we could not paraphrase this essay. Our professional writers can rewrite it and get you a unique paper.

Please check your inbox.

We can write you a custom essay that will follow your exact instructions and meet the deadlines. Let's fix your grades together!

Get Your Personalized Essay in 3 Hours or Less!

We use cookies to personalyze your web-site experience. By continuing we’ll assume you board with our cookie policy .

  • Instructions Followed To The Letter
  • Deadlines Met At Every Stage
  • Unique And Plagiarism Free

the cosmological argument essay

Reasonable Faith Logo

  • Equip Project
  • RF Chapters
  • Translations

Get Dr. Craig's newsletter and keep up with RF news and events.

the cosmological argument essay

The Kalam Cosmological Argument

This article is the text of Dr. Craig's 2015 lecture at the University of Birmingham, where he did his doctoral studies which led to the revival of the kalam cosmological argument in our day.

As a boy I wondered at the existence of the universe. I wondered where it came from. Did it have a beginning? I remember lying in bed at night trying to think of a beginningless universe. Every event would be preceded by another event, back and back into the past, with no stopping point—or, more accurately, no starting point! An infinite past, with no beginning! My mind reeled at the prospect. It just seemed inconceivable to me. There must have been a beginning at some point, I thought, in order for everything to get started.

Little did I suspect that for centuries—millennia, really—men had grappled with the idea of an infinite past and the question of whether there was a beginning of the universe. Ancient Greek philosophers believed that matter was necessary and uncreated and therefore eternal. God may be responsible for introducing order into the cosmos, but He did not create the universe itself.

This Greek view was in contrast to even more ancient Jewish thought about the subject. Hebrew writers held that the universe has not always existed but was created by God at some point in the past. As the first verse of the Hebrew holy scriptures states: “In the beginning God created the heavens and the earth” (Genesis 1:1).

Eventually these two competing traditions began to interact. There arose within Western philosophy an ongoing debate that lasted for well over a thousand years about whether or not the universe had a beginning. This debate played itself out among Jews and Muslims as well as Christians, both Catholic and Protestant. It finally sputtered to something of an inconclusive end in the thought of the great eighteenth century German philosopher Immanuel Kant. He held, ironically, that there are rationally compelling arguments for both sides, thereby exposing the bankruptcy of reason itself!

I first became aware of this debate only after graduating from university. Wanting to come to terms with this question, I decided upon completion of my Master’s degree work in philosophy to find someone who would be willing to supervise a doctoral thesis on this question. The person who stood out above all others was Prof. John Hick at the Universty of Birmingham. We did come to Birmingham, and I did write on the cosmological argument under Prof. Hick’s direction, and eventually three books flowed out of that doctoral thesis. I was able to explore the historical roots of the argument, as well as deepen and advance the analysis of the argument. I also discovered quite amazing connections to contemporary astronomy and cosmology.

Because of its historic roots in medieval Islamic theology, I christened the argument “the kalam cosmological argument” (“ kalam ” is the Arabic word for medieval theology). Today this argument, largely forgotten since the time of Kant, is once again back at center stage. The Cambridge Companion to Atheism (2007) reports, “A count of the articles in the philosophy journals shows that more articles have been published about . . . the Kalam argument than have been published about any other . . . contemporary formulation of an argument for God’s existence. . . . theists and atheists alike ‘cannot leave [the] Kalam argument alone’” (p. 183).

What is the argument which has stirred such interest? Let’s allow one of the greatest medieval protagonists in this debate to speak for himself. Al-Ghazali was a twelfth century Muslim theologian from Persia, or modern day Iran. He was concerned that Muslim philosophers of his day were being influenced by ancient Greek philosophy to deny God’s creation of the universe. After thoroughly studying the teachings of these philosophers, Ghazali wrote a withering critique of their views entitled The Incoherence of the Philosophers. In this fascinating book, he argues that the idea of a beginningless universe is absurd. The universe must have a beginning, and since nothing begins to exist without a cause, there must be a transcendent Creator of the universe.

Ghazali formulates his argument very simply: “Every being which begins has a cause for its beginning; now the world is a being which begins; therefore, it possesses a cause for its beginning.”  [1]

Ghazali’s reasoning involves three simple steps:

1. Whatever begins to exist has a cause of its beginning.

2. The universe began to exist.

3. Therefore, the universe has a cause of its beginning.

Let’s look at each step of this argument.

Notice that Ghazali does not need a premise so strong as (1) in order for his argument to succeed. The first premise can be more modestly stated.

1'. If the universe began to exist, then the universe has a cause of its beginning.

This more modest version of the first premise will enable us to avoid distractions about whether subatomic particles which are the result of quantum decay processes come into being without a cause. This alleged exception to (1) is irrelevant to (1'). For the universe comprises all contiguous spacetime reality. Therefore, for the whole universe to come into being without a cause is to come into being from nothing, which is absurd. In quantum decay events, the particles do not come into being from nothing. As Christopher Isham, Britain’s premier quantum cosmologist, cautions,

Care is needed when using the word ‘creation’ in a physical context. One familiar example is the creation of elementary particles in an accelerator. However, what occurs in this situation is the conversion of one type of matter into another, with the total amount of energy being preserved in the process.  [2]

Thus, this alleged exception to (1) is not an exception to (1').

Let me give three reasons in support of premise (1'):

1. Something cannot come from nothing. To claim that something can come into being from nothing is worse than magic. When a magician pulls a rabbit out of a hat, at least you’ve got the magician, not to mention the hat! But if you deny premise (1'), you’ve got to think that the whole universe just appeared at some point in the past for no reason whatsoever. But nobody sincerely believes that things, say, a horse or an Eskimo village, can just pop into being without a cause.

2. If something can come into being from nothing, then it becomes inexplicable why just anything or everything doesn’t come into being from nothing. Think about it: why don’t bicycles and Beethoven and root beer just pop into being from nothing? Why is it only universes that can come into being from nothing? What makes nothingness so discriminatory? There can’t be anything about nothingness that favors universes, for nothingness doesn’t have any properties. Nor can anything constrain nothingness, for there isn’t anything to be constrained!

3. Common experience and scientific evidence confirm the truth of premise 1'. The science of cosmogeny is based on the assumption that there are causal conditions for the origin of the universe. So it’s hard to understand how anyone committed to modern science could deny that (1') is more plausibly true than false.

So I think that the first premise of the kalam cosmological argument is surely true.

The more controversial premise in the argument is premise 2, that the universe began to exist. This is by no means obvious. Let’s examine both philosophical arguments and scientific evidence in support of premise 2.

First Philosophical Argument

Ghazali argued that if the universe never began to exist, then there has been an infinite number of past events prior to today. But, he argued, an infinite number of things cannot exist. Ghazali recognized that a potentially infinite number of things could exist, but he denied that an actually infinite number of things could exist.

When we say that something is potentially infinite, infinity serves merely as an ideal limit which is never reached. For example, you could divide any finite distance in half, and then into fourths, and then into eighths, and then into sixteenths, and so on to infinity. The number of divisions is potentially infinite, in the sense that you could go on dividing endlessly. But you would never arrive at an “infinitieth” division. You would never have an actually infinite number of parts or divisions.

Now Ghazali has no problem with the existence of merely potential infinites, for these are just ideal limits. But he argued that if an actually infinite number of things could exist, then various absurdities would result. If we’re to avoid these absurdities, then we must deny that an actually infinite number of things exist. That implies that the number of past events cannot be actually infinite. Therefore, the universe cannot be beginningless; rather the universe began to exist.

It’s very frequently alleged that this kind of argument has been invalidated by developments in modern mathematics. In modem set theory the use of actually infinite sets is commonplace. For example, the set of the natural numbers {0, 1, 2, . . .} has an actually infinite number of members in it. The number of members in this set is not merely potentially infinite, according to modern set theory; rather the number of members is actually infinite. Many people have inferred that these developments undermine Ghazali’s argument.

But is that really the case? Modern set theory shows that if you adopt certain axioms and rules, then you can talk about actually infinite collections in a consistent way, without contradicting yourself. All this accomplishes is showing how to set up a certain universe of discourse for talking consistently about actual infinites. But it does absolutely nothing to show that such mathematical entities really exist or that an actually infinite number of things can really exist. If Ghazali is right, then this universe of discourse may be regarded as just a fictional realm, like the world of Sherlock Holmes, or something that exists only in your mind.

The way in which Ghazali brings out the real impossibility of an actually infinite number of things is by imagining what it would be like if such a collection could exist and then drawing out the absurd consequences. Let me share one of my favorite illustrations called “Hilbert’s Hotel,” the brainchild of the great German mathematician David Hilbert.

Hilbert first invites us to imagine an ordinary hotel with a finite number of rooms. Suppose, furthermore, that all the rooms are full. If a new guest shows up at the desk asking for a room, the manager says, “Sorry, all the rooms are full,” and that’s the end of the story.

But now, says Hilbert, let’s imagine a hotel with an infinite number of rooms, and let’s suppose once again that all the rooms are full. This fact must be clearly appreciated. There is not a single vacancy throughout the entire infinite hotel; every room already has a flesh-and-blood person in it. Now suppose a new guest shows up at the front desk, asking for a room. “No problem,” says the manager. He moves the person who was staying in room #1 into room #2, the person who was staying in room #2 into room #3, the person who was staying in room #3 into room #4, and so on to infinity. As a result of these room changes, room #1 now becomes vacant, and the new guest gratefully checks in. But before he arrived, all the rooms were already full!

It gets worse! Let’s now suppose, Hilbert says, that an infinity of new guests shows up at the front desk, asking for rooms. “No problem, no problem!” says the manager. He moves the person who was staying in room #1 into room #2, the person who was staying in room #2 into room #4, the person who was staying in room #3 into room #6, each time moving the person into the room number twice his own. Since any number multiplied by two is an even number, all the guests wind up in even-numbered rooms. As a result, all the odd-numbered rooms become vacant, and the infinity of new guests is easily accommodated. In fact, the manager could do this an infinite number of times and always accommodate infinitely more guests. And yet, before they arrived, all the rooms were already full!

As a student once remarked to me, Hilbert’s Hotel, if it could exist, would have to have a sign posted outside: “No Vacancy (Guests Welcome).” Can such a hotel exist in reality?

Hilbert’s Hotel is absurd. Since nothing hangs on the illustration’s involving a hotel, the argument can be generalized to show that the existence of an actually infinite number of things is absurd.

Sometimes people react to Hilbert’s Hotel by saying that these absurdities result because the concept of infinity is beyond us and we can’t understand it. But this reaction is mistaken and naïve. As I said, infinite set theory is a highly developed and well-understood branch of modern mathematics. The absurdities result because we do understand the nature of the actual infinite. Hilbert was a smart guy, and he knew well how to illustrate the bizarre consequences of the existence of an actually infinite number of things.

Really, the only thing the critic can do at this point is to just bite the bullet and say that a Hilbert’s Hotel is not absurd. Sometimes critics will try to justify this move by saying that if an actual infinite could exist, then such situations are exactly what we should expect. But this response is inadequate. Hilbert would, of course, agree that if an actual infinite could exist, the situation with his imaginary hotel is what we would expect. Otherwise, it wouldn’t be a good illustration! But the question is whether such a hotel is really possible.

So I think Ghazali’s first argument is a good one. It shows that the number of past events must be finite. Therefore, the universe must have had a beginning. We can summarize Ghazali’s argument as follows:

1. An actual infinite cannot exist.

2. An infinite temporal regress of events is an actual infinite.

3. Therefore an infinite temporal regress of events cannot exist.

Second Philosophical Argument

Ghazali has a second, independent argument for the beginning of the universe. The series of past events, Ghazali observes, has been formed by adding one event after another. The series of past events is like a sequence of dominoes falling one after another until the last domino, today, is reached. But, he argues, no series which is formed by adding one member after another can be actually infinite. For you cannot pass through an infinite number of elements one at a time.

This is easy to see in the case of trying to count to infinity. No matter how high you count, there is always an infinity of numbers left to count.

But if you can’t count to infinity, how could you count down from infinity? This would be like someone’s claiming to have counted down all the negative numbers, ending at zero: . . ., -3, -2, -1, 0. This seems crazy. For before he could count 0, he would have to count -1, and before he could count -1, he would have to count -2, and so on, back to infinity. Before any number could be counted an infinity of numbers will have to have been counted first. You just get driven back and back into the past, so that no number could ever be counted.

But then the final domino could never fall if an infinite number of dominoes had to fall first. So today could never be reached. But obviously here we are! This shows that the series of past events must be finite and have a beginning.

Ghazali sought to heighten the impossibility of forming an infinite past by giving illustrations of the absurdities that would result if it could be done. For example, suppose that for every one orbit that Saturn completes around the sun Jupiter completes two. The longer they orbit, the further Saturn falls behind. If they continue to orbit forever, they will approach a limit at which Saturn is infinitely far behind Jupiter. Of course, they will never actually arrive at this limit.

But now turn the story around: suppose Jupiter and Saturn have been orbiting the sun from eternity past. Which will have completed the most orbits? The answer is that the number of their orbits is exactly the same: infinity! (We can’t slip out of this argument by saying that infinity is not a number. In modern mathematics it is a number, the number of elements in the set {0, 1, 2, 3, . . . }.) But that seems absurd, for the longer they orbit, the greater the disparity grows. So how does the number of orbits magically become equal by making them orbit from eternity past?

Another illustration: suppose we meet someone who claims to have been counting down from eternity past and is now finishing: . . . -3, -2, -1, 0! Whew! Why, we may ask, is he just finishing his countdown today? Why didn’t he finish yesterday or the day before? After all, by then an infinite amount of time had already elapsed. So if the man were counting at a rate of one number per second, he’s already had an infinite number of seconds to finish his countdown. He should already be done! In fact, at any point in the past, he has already had infinite time and so should already have finished. But then at no point in the past can we find the man finishing his countdown, which contradicts the hypothesis that he has been counting from eternity.

Alexander Pruss and Robert Koons have recently defended an engaging contemporary version of Ghazali’s argument called the Grim Reaper Paradox. There are infinitely many Grim Reapers (whom we may identify as gods, so as to forestall any physical objections). You are alive at midnight. Grim Reaper 1 will strike you dead at 1:00 a.m. if you are still alive at that time. Grim Reaper 2 will strike you dead at 12:30 a.m. if you are still alive then. Grim Reaper 3 will strike you dead at 12:15 a.m., and so on. Such a situation seems clearly conceivable—given the possibility of an actually infinite number of things—but leads to an impossibility: you cannot survive past midnight, and yet you cannot be killed by any Grim Reaper at any time. Pruss and Koons show how to re-formulate the paradox so that the Grim Reapers are spread out over infinite time rather than over a single hour, for example, by having each Grim Reaper swing his scythe on January 1 of each past year if you have managed to live that long.

These illustrations only strengthen Ghazali’s claim that no series which is formed by adding one member after another can be actually infinite. Since the series of past events has been formed by adding one event after another, it can’t be actually infinite. It must have had a beginning. So we have a second good argument for premise 2, that the universe began to exist. We can summarize this argument as follows:

1. A collection formed by successive addition cannot be an actual infinite.

2. The temporal series of events is a collection formed by successive addition.

3. Therefore, the temporal series of events cannot be an actual infinite.

First Scientific Confirmation

One of the most astonishing developments of modern astronomy, which Ghazali would never have anticipated, is that we now have strong scientific evidence for the beginning of the universe. The first scientific confirmation of the universe’s beginning comes from the expansion of the universe.

All throughout history men have assumed that the universe as a whole was unchanging. Of course, things in the universe were moving about and changing, but the universe itself was just there, so to speak. This was also Albert Einstein’s assumption when he first began to apply his new theory of gravity, called the General Theory of Relativity, to the universe in 1917.

But Einstein found there was something terribly amiss. His equations described a universe which was either blowing up like a balloon or else collapsing in upon itself. During the 1920s the Russian mathematician Alexander Friedman and the Belgian astronomer Georges LeMaître decided to take Einstein’s equations at face value, and as a result they came up independently with models of an expanding universe. In 1929 the American astronomer Edwin Hubble, through tireless observations at Mt. Wilson Observatory, made a startling discovery which verified Friedman and LeMaître’s theory. He found that the light from distant galaxies appeared to be redder than expected. This “red shift” in the light was most plausibly due to the stretching of the light waves as the galaxies are moving away from us. Wherever Hubble trained his telescope in the night sky, he observed this same red-shift in the light from the galaxies. It appeared that we are at the center of a cosmic explosion, and all of the other galaxies are flying away from us at fantastic speeds!

Now according to the Friedman-LeMaître model, we are not really at the center of the universe. Rather an observer in any galaxy will look out and see the other galaxies moving away from him. This is because, according to the theory, it is really space itself which is expanding. The galaxies are actually at rest in space, but they recede from one another as space itself expands.

The Friedman-LeMaître model eventually came to be known as the Big Bang theory. But that name can be misleading. Thinking of the expansion of the universe as a sort of explosion could mislead us into thinking that the galaxies are moving out into a pre-existing, empty space from a central point. That would be a complete misunderstanding of the model. The theory is much more radical than that.

As you trace the expansion of the universe back in time, everything gets closer and closer together. Eventually the distance between any two points in space becomes zero. You can’t get any closer than that! So at that point you’ve reached the boundary of space and time. Space and time cannot be extended any further back than that. It is literally the beginning of space and time.

To get a picture of this we can portray our three-dimensional space as a two-dimensional disk which shrinks as you go back in time (Fig. 2).

Fig. 2. Geometrical representation of space-time. The two-dimensional disc represents our three-dimensional space. The vertical dimension represents time. As one goes back in time, space shrinks until the distance between any two points is zero. Space-time thus has the geometry of a cone. The point of the cone is the boundary of space and time.

Eventually, the distance between any two points in space becomes zero. So space-time can be represented geometrically as a cone. What’s significant about this is that while a cone can be extended indefinitely in one direction, it has a boundary point in the other direction. Because this direction represents time and the boundary point lies in the past, the model implies that past time is finite and had a beginning.

Because space-time is the arena in which all matter and energy exist, the beginning of space-time is also the beginning of all matter and energy. It’s the beginning of the universe.

Notice that there’s simply nothing prior to the initial boundary of space-time. Let’s not be misled by words. When cosmologists say, “There is nothing prior to the initial boundary,” they do not mean that there is some state of affairs prior to it, and that is a state of nothingness. That would be to treat nothing as though it were something! Rather they mean that at the boundary point, it is false that “There is something prior to this point.”

The standard Big Bang model thus predicts an absolute beginning of the universe. If this model is correct, then we have amazing scientific confirmation of the second premise of the kalam cosmological argument.

So is the model correct, or, more importantly, is it correct in predicting a beginning of the universe? Despite its empirical confimation, the standard Big Bang model will need to be modified in various ways. The model is based, as we’ve seen, on Einstein’s General Theory of Relativity. But Einstein’s theory breaks down when space is shrunk down to sub-atomic proportions. We’ll need to introduce sub-atomic physics at that point, and no one is sure how this is to be done. Moreover, the expansion of the universe is probably not constant, as in the standard model. It’s probably accelerating and may have had a brief moment of super-rapid expansion in the past.

But none of these adjustments need affect the fundamental prediction of the absolute beginning of the universe. Indeed, physicists have proposed scores of alternative models over the decades since Friedman and LeMaître’s work, and those that do not have an absolute beginning have been repeatedly shown to be unworkable. Put more positively, the only viable non-standard models have been those that involve an absolute beginning to the universe. That beginning may or may not involve a beginning point. But on theories (such as Stephen Hawking’s “no boundary” proposal) that do not have a point-like beginning, the past is still finite, not infinite. The universe has not existed forever according to such theories but came into existence, even if it didn’t do so at a sharply defined point.

In a sense, the history of twentieth century cosmology can be seen as a series of one failed attempt after another to avoid the absolute beginning predicted by the standard Big Bang model. That prediction has now stood for nearly 100 years, during a period of enormous advances in observational astronomy and creative theoretical work in astrophysics.

Meanwhile, a series of remarkable singularity theorems has increasingly tightened the loop around empirically tenable models by showing that under more and more generalized conditions, a beginning is inevitable. In 2003 Arvind Borde, Alan Guth, and Alexander Vilenkin were able to show that any universe which is, on average, in a state of cosmic expansion throughout each history cannot be infinite in the past but must have a beginning. That goes for multiverse scenarios, too. In 2012 Vilenkin showed that models which do not meet this one condition still fail for other reasons to avert the beginning of the universe. Vilenkin concluded, “None of these scenarios can actually be past eternal.”  [3]  “All the evidence we have says that the universe had a beginning.”  [4]

The Borde-Guth-Vilenkin theorem proves that classical space-time, under a single, very general condition, cannot be extended to past infinity but must reach a boundary at some time in the finite past. Now either there was something on the other side of that boundary or not. If not, then that boundary just is the beginning of the universe. If there was something on the other side, then it will be a region described by the yet-to-be discovered theory of quantum gravity. In that case, Vilenkin says, it will be the beginning of the universe. Either way, the universe began to exist.

Of course, scientific results are always provisional. We can fully expect that new theories will be proposed, attempting to avoid the universe’s beginning. Such proposals are to be welcomed and tested. Nevertheless, it’s pretty clear which way the evidence points. Today the proponent of Ghazali’s cosmological argument stands comfortably within the scientific mainstream in holding that the universe began to exist.

Second Scientific Argument

As if this weren’t enough, there is actually a second scientific confirmation of the beginning of the universe, this one from the Second Law of Thermodynamics. According to the Second Law, unless energy is being fed into a system, that system will become increasingly disorderly.

Now already in the nineteenth century scientists realized that the Second Law implied a grim prediction for the future of the universe. Given enough time, all the energy in the universe will spread itself out evenly throughout the universe. The universe will become a featureless soup in which no life is possible. Once the universe reaches such a state, no significant further change is possible. It is a state of equilibrium . Scientists called this the “heat death” of the universe.

But this unwelcome prediction raised a further puzzle: if, given enough time, the universe will inevitably stagnate in a state of heat death, then why, if it has existed forever, is it not now in a state of heat death? If in a finite amount of time, the universe will reach equilibrium, then, given infinite past time, it should by now already be in state of equilibrium. But it’s not. We’re in a state of dis equilibrium, where energy is still available to be used and the universe has an orderly structure.

The nineteenth century German physicist Ludwig Boltzmann proposed a daring solution to this problem. Boltzmann suggested that perhaps the universe is, in fact, in a state of overall equilibrium. Nevertheless, by chance alone, there will arise more orderly pockets of disequilibrium here and there. Boltzmann refers to these isolated regions of disequilibrium as “worlds.” Our universe just happens to be one of these worlds. Eventually, in accord with the Second Law, it will revert to the overall state of equilibrium.

Contemporary physicists have universally rejected Boltzmann’s daring Many Worlds Hypothesis as an explanation of the observed disequilibrium of the universe. Its fatal flaw is that if our world is just a chance fluctuation from a state of overall equilibrium, then we ought to be observing a much smaller patch of order. Why? Because a small fluctuation from equilibrium is vastly more probable than the huge, sustained fluctuation necessary to create the universe we see, and yet a small fluctuation would be sufficient for our existence. For example, a fluctuation that formed a world no bigger than our solar system would be enough for us to be alive and would be incomprehensibly more likely to occur than a fluctuation that formed the whole universe we see!

In fact, Boltzmann’s hypothesis, if consistently carried out, would lead to a strange sort of illusionism: in all probability we really do inhabit a smaller world, and the stars and the planets we observe are just illusions, mere images on the heavens. For that sort of world is much more probable than a universe which has, in defiance of the Second Law of Thermodynamics, moved away from equilibrium for billions of years to form the universe we observe.

The discovery of the expansion of the universe in the 1920s modified the sort of heat death predicted on the basis of the Second Law, but it didn’t alter the fundamental question. Recent discoveries indicate that the cosmic expansion is actually speeding up. Because the volume of space is increasing so rapidly, the universe actually becomes farther and farther from an equilibrium state in which matter and energy are evenly distributed. But the acceleration of the universe’s expansion only hastens its demise. For now the different regions of the universe become increasingly isolated from one another in space, and each marooned region becomes dark, cold, dilute, and dead. So again, why isn’t our region in such a state if the universe has already existed for infinite time?

The obvious implication of all this is that the question is based on a false assumption, namely, that the universe has existed for infinite time. Today most physicists would say that the matter and energy were simply put into the universe as an initial condition, and the universe has been following the path plotted by the Second Law ever since its beginning a finite time ago.

Of course, attempts have been made to avoid the beginning of the universe predicted on the basis of the Second Law of Thermodynamics. But none of them has been successful. Skeptics might hold out hope that quantum gravity will serve to avert the implications of the Second Law of Thermodynamics. But in 2013, the cosmologist Aron Wall of the University of California was able to formulate a new singularity theorem which seems to close the door on that possibility. Wall shows that, given the validity of the generalized Second Law of Thermodynamics in quantum gravity, the universe must have begun to exist, unless one postulates a reversal of the arrow of time (time runs backwards!) at some point in the past, which, he rightly observes, involves a thermodynamic beginning in time which “would seem to raise the same sorts of philosophical questions that any other sort of beginning in time would.”  [5]  Wall reports that his results require the validity of only certain basic concepts, so that “it is reasonable to believe that the results will hold in a complete theory of quantum gravity.”

So once again the scientific evidence confirms the truth of the second premise of Ghazali’s cosmological argument.

On the basis, therefore, of both philosophical and scientific evidence, we have good grounds for believing that the universe began to exist. It therefore follows that the universe has a cause of its beginning.

What properties must this cause of the universe possess? This cause must be itself uncaused because we’ve seen that an infinite series of causes is impossible. It is therefore the Uncaused First Cause. It must transcend space and time, since it created space and time. Therefore, it must be immaterial and non-physical. It must be unimaginably powerful, since it created all matter and energy.

Finally, Ghazali argued that this Uncaused First Cause must also be a personal being. It’s the only way to explain how an eternal cause can produce an effect with a beginning like the universe.

Here’s the problem: If a cause is sufficient to produce its effect, then if the cause is there, the effect must be there, too. For example, the cause of water’s freezing is the temperature’s being below 0 degrees Celsius. If the temperature has been below 0 degrees from eternity, then any water around would be frozen from eternity. It would be impossible for the water to begin to freeze just a finite time ago. Now the cause of the universe is permanently there, since it is timeless. So why isn’t the universe permanently there as well? Why did the universe come into being only 14 billion years ago? Why isn’t it as permanent as its cause?

Ghazali maintained that the answer to this problem is that the First Cause must be a personal being endowed with freedom of the will. His creating the universe is a free act which is independent of any prior determining conditions. So his act of creating can be something spontaneous and new. Freedom of the will enables one to get an effect with a beginning from a permanent, timeless cause. Thus, we are brought not merely to a transcendent cause of the universe but to its Personal Creator.

This is admittedly hard for us to imagine. But one way to think about it is to envision God existing alone without the universe as changeless and timeless. His free act of creation is a temporal event simultaneous with the universe’s coming into being. Therefore, God enters into time when He creates the universe. God is thus timeless without the universe and in time with the universe.

Ghazali’s cosmological argument thus gives us powerful grounds for believing in the existence of a beginningless, uncaused, timeless, spaceless, changeless, immaterial, enormously powerful, Personal Creator of the universe.

Al-Gha-zalı-, Kitab al-Iqtisad fi’l-I’tiqad , cited in S. de Beaurecueil, “Gazzali et S. Thomas d’Aquin: Essai sur la preuve de l’exitence de Dieu proposée dans l’Iqtisad et sa comparaison avec les ‘voies’ Thomiste,” Bulletin de l’Institut Francais d’Archaeologie Orientale 46 (1947): 203.

Christopher Isham, “Creation of the Universe as a Quantum Process,” p. 378.

Audrey Mithani and Alexander Vilenkin, “Did the universe have a beginning?” arXiv:1204.4658v1 [hep-th] 20 Apr 2012, p. 5. For an accessible video, see see http://www.youtube.com/watch?v=NXCQelhKJ7A (accessed February 23, 2014), where Vilenkin concludes, “there are no models at this time that provide a satisfactory model for a universe without a beginning.”

A.Vilenkin, cited in “Why physicists can't avoid a creation event,” by Lisa Grossman, New Scientist (January 11, 2012).

Aron C. Wall, “The Generalized Second Law implies a Quantum Singularity Theorem,” arXiv: 1010.5513v3 [gr-qc] 24 (Jan 2013), p. 38.

Aquinas – the cosmological argument for the existence of God

the cosmological argument essay

The cosmological argument stems from the idea that the world and everything that is in it is dependent on something other than itself for its existence.  Even though the world may appear to be self-perpetuating, it is necessary to understand the source. Before Thomas Aquinas, both Plato and Aristotle too argued that something could not come from nothing.  There has to be something, which exists to cause a movement.  The three major arguments put forth by Aquinas known as the Cosmological Argument will be discussed here.

In his work, Summa Theologica Thomas Aquinas offered five ‘proofs’ for the existence of God. The first argument was the Argument of Motion. Aquinas’s argument has to be understood keeping in mind Aristotle’s discussion of Astronomy. Aristotle argued that the planetary position, which causes the seasons to change, requires an unmoved mover to maintain the order of things. Aquinas’ argument was based on this very premise that without God the heaven and earth would not exist. This implies that any event in the universe is the result of some cause. The argument is that this chain of events either has a cause or does not. While Aristotle left it at the Uncaused cause Aquinas named this uncaused cause as ‘God’.  Astronomers refute this theory and rely on the Big Bang Theory, which is the scientific theory that the universe emerged from an enormously dense and hot state nearly 14 billion years ago. Bertrand Russell too disagrees and says that the ‘universe just is’ without any cause.

Aquinas further argues that there can be no effect without an ‘efficient cause’. There is nothing in the world that can be the efficient cause of itself. To take away the cause is to take away the effect. Therefore, if there were no first cause among efficient causes, there will be no ultimate, or intermediate, cause. There cannot be an endless regression of cause and effect and hence the first cause must be God.

The third argument is based on possibility and necessity. According to Aquinas, it is logically possible that the universe has already existed for an infinite amount of time, and will continue to exist for an infinite amount of time. If the universe could exist or could not exist, that is to say, it is contingent, then its existence must have a cause. Objects have contingent existence but God has necessary existence. Aquinas argues that if everything can not-be, then at one time there was nothing in existence. If at one time nothing was in existence, it would have been impossible for anything to have begun to exist; and thus even now nothing would be in existence – which is absurd.

Aquinas further argues on the degrees of perfection and goodness that can be seen in the world. Humans have the capacity for both good and bad deeds. Therefore, the maximum in the genus (group of things) of morality must be God (the perfect being), who is the ‘first cause’, or source, of all goodness and perfection.

the cosmological argument essay

The cosmological argument not only seeks to reason the existence of God but could also be said to provide a meaning to life in the world. For instance, if we know where we have come from then surely, it could be argued, we have some idea of where we are going. Hence, Aquinas comes to the same conclusion that God exists, whether there was a first event in the universe or not.

  • ✔️ Social Commentary
  • ✝️ Christianity
  • Islamophobia

the cosmological argument essay

SEP logo

  • Table of Contents
  • New in this Archive
  • Chronological
  • Editorial Information
  • About the SEP
  • Editorial Board
  • How to Cite the SEP
  • Special Characters
  • Support the SEP
  • PDFs for SEP Friends
  • Make a Donation
  • SEPIA for Libraries
  • Entry Contents

Bibliography

Academic tools.

  • Friends PDF Preview
  • Author and Citation Info
  • Back to Top

Cosmological Argument

The cosmological argument is less a particular argument than an argument type. It uses a general pattern of argumentation ( logos ) that makes an inference from particular alleged facts about the universe ( cosmos ) to the existence of a unique being, generally identified with or referred to as God. Among these initial facts are that particular beings or events in the universe are causally dependent or contingent, that the universe (as the totality of contingent things) is contingent in that it could have been other than it is, that the Big Conjunctive Contingent Fact possibly has an explanation, or that the universe came into being. From these facts philosophers infer deductively, inductively, or abductively by inference to the best explanation that a first or sustaining cause, a necessary being, an unmoved mover, or a personal being (God) exists that caused and/or sustains the universe. The cosmological argument is part of classical natural theology, whose goal is to provide evidence for the claim that God exists.

On the one hand, the argument arises from human curiosity as to why there is something rather than nothing or than something else. It invokes a concern for some full, complete, ultimate, or best explanation of what exists contingently. On the other hand, it raises intrinsically important philosophical questions about contingency and necessity, causation and explanation, part/whole relationships (mereology), infinity, sets, the nature of time, and the nature and origin of the universe. In what follows we will first sketch out a very brief history of the argument, note the two basic types of deductive cosmological arguments, and then provide a careful analysis of examples of each: first, two arguments from contingency, one based on a relatively strong version of the principle of sufficient reason and one based on a weak version of that principle; and second, an argument from the alleged fact that the universe had a beginning and the impossibility of an infinite temporal regress of causes. In the end we will consider an inductive version of the cosmological argument and what it is to be a necessary being.

1. Historical Overview

2. typology of cosmological arguments, 3. complexity of the question, 4.1 a deductive argument from contingency, 4.2 objection 1: the universe just is, 4.3 objection 2: explaining the individual constituents is sufficient, 4.4 objection 3: the principles of causation and of sufficient reason are suspect, 4.5 objection 4: problems with the concept of a necessary being, 5. argument from the weak principle of sufficient reason, 6.1 the causal principle and quantum physics, 6.2 impossibility of an actual infinite, 6.3 successive addition cannot form an actual infinite, 6.4 the big bang theory of cosmic origins, 6.5 the big bang is not an event, 6.6 a non-finite universe, 6.7 personal explanation, 7. an inductive cosmological argument, 8. necessary being, other internet resources, related entries.

Although in Western philosophy the earliest formulation of a version of the cosmological argument is found in Plato’s Laws , 893–96, the classical argument is firmly rooted in Aristotle’s Physics ( VIII, 4–6) and Metaphysics (XII, 1–6). Islamic philosophy enriches the tradition, developing two types of arguments. Arabic philosophers ( falasifa ), such as Ibn Sina (c. 980–1037), developed the argument from contingency, which is taken up by Thomas Aquinas (1225–74) in his Summa Theologica (I,q.2,a.3) and his Summa Contra Gentiles (I, 13). Influenced by John Philoponus (5th c) (Davidson 1969) the mutakallimūm —theologians who used reason and argumentation to support their revealed Islamic beliefs—developed the temporal version of the argument from the impossibility of an infinite regress, now referred to as the kalām argument. For example, al-Ghāzāli (1058–1111) argued that everything that begins to exist requires a cause of its beginning. The world is composed of temporal phenomena preceded by other temporally-ordered phenomena. Since such a series of temporal phenomena cannot continue to infinity because an actual infinite is impossible, the world must have had a beginning and a cause of its existence, namely, God (Craig 1979: part 1). This version of the argument enters the medieval Christian tradition through Bonaventure (1221–74) in his Sentences ( II Sent . D.1,p.1,a.1,q.2).

Enlightenment thinkers, such as Georg Wilhelm Leibniz and Samuel Clarke, reaffirmed the cosmological argument. Leibniz (1646–1716) appealed to a strengthened principle of sufficient reason, according to which “no fact can be real or existing and no statement true without a sufficient reason for its being so and not otherwise” ( Monadology , §32). Leibniz uses the principle to argue that the sufficient reason for the “series of things comprehended in the universe of creatures” (§36) must exist outside this series of contingencies and is found in a necessary being that we call God. The principle of sufficient reason is likewise employed by Samuel Clarke in his cosmological argument (Rowe 1975: chap. 2).

Although the cosmological argument does not figure prominently in Asian philosophy, a very abbreviated version of it, proceeding from dependence, can be found in Udayana’s Nyāyakusumāñjali I,4. In general, philosophers in the Nyāya tradition argue that since the universe has parts that come into existence at one occasion and not another, it must have a cause. We could admit an infinite regress of causes if we had evidence for such, but lacking such evidence, God must exist as the non-dependent cause. Many of the objections to the argument contend that God is an inappropriate cause because of God’s nature. For example, since God is immobile and has no body, he cannot properly be said to cause anything. The Naiyāyikas reply that God could assume a body at certain times, and in any case, God need not create in the same way humans do (Potter 1977: 100–07).

The cosmological argument came under serious assault in the 18 th century, first by David Hume and then by Immanuel Kant. Hume (1748) attacks both the view of causation presupposed in the argument (that causation is an objective, productive, necessary power relation that holds between two things) and the Causal Principle—every contingent being has a cause of its existence—that lies at the heart of the argument. Kant contends that the cosmological argument, in identifying the necessary being, relies on the ontological argument, which in turn is suspect. We will return to these criticisms below.

Both theists and nontheists in the last part of the 20 th century and the first part of the 21 st century generally have shown a healthy skepticism about the argument. Alvin Plantinga concludes “that this piece of natural theology is ineffective” (1967: chap. 1). Richard Gale contends, in Kantian fashion, that since the conclusion of all versions of the cosmological argument invokes an impossibility, no cosmological arguments can provide examples of sound reasoning (1991: chap. 7). (However, Gale seems to have changed his mind and in recent writings proposed and defended his own version of the cosmological argument, which we will consider below.) Similarly, Michael Martin (1990: chap. 4), John Mackie (1982: chap. 5), Quentin Smith (Craig and Smith 1993), Bede Rundle (2004), Wes Morriston (2000, 2002, 2003, 2010), and Graham Oppy (2006: chap. 3) reason that no current version of the cosmological argument is sound. Yet dissenting voices can be heard. Robert Koons (1997) employs mereology and modal and nonmonotonic logic in taking a “new look” at the argument from contingency. In his widely discussed writings William Lane Craig marshals multidisciplinary evidence for the truth of the premises found in the kalām argument. Richard Gale and Alexander Pruss propose a new version based on a so-called weak principle of sufficient reason that leads to a finite God that is not omnibenevolent, and Richard Swinburne, though rejecting deductive versions of the cosmological argument, proposes an inductive argument that is part of a larger cumulative case for God’s existence.

There is quite a chance that if there is a God he will make something of the finitude and complexity of a universe. It is very unlikely that a universe would exist uncaused, but rather more likely that God would exist uncaused. The existence of the universe…can be made comprehensible if we suppose that it is brought about by God. (1979: 131–32)

In short, contemporary philosophers continue to contribute increasingly detailed and complex arguments on both sides of the debate.

Philosophers employ diverse classifications of the cosmological arguments. Swinburne distinguishes inductive from deductive versions. Craig distinguishes three types of deductive cosmological arguments in terms of their approach to an infinite regress of causes. The first, advocated by Aquinas, is based on the impossibility of an essentially ordered infinite regress. The second, which Craig terms the kalām argument, holds that an infinite temporal regress of causes is impossible because an actual infinite is impossible, and even if it were possible it could not be temporally realized. The third, espoused by Leibniz and Clarke, is overtly founded on the Principle of Sufficient Reason (Craig 1980: 282–83). Craig notes that the distinction between these types of arguments is important because the objections raised against one version may be irrelevant to other versions. So, for example, a critique of a particular version of the Principle of Sufficient Reason (PSR), which one finds developed by William Rowe or Richard Gale, might not be telling against the Thomistic or kalām versions of the argument. Another way of distinguishing between versions of the argument is in terms of the relevance of time to the argument. In Aquinas’s version, consideration of the essential ordering of the causes or reasons proceeds independent of temporal concerns. The relationship between cause and effect is treated as real but not temporal, so that the first cause is not a first cause in time but a sustaining cause. In the kalām version, however, the temporal ordering of the causal sequence is central, introducing issues of the nature of time into the discussion.

It is said that philosophy begins in wonder. So it was for the ancients, who wondered what constituted the basic stuff of the world around them, how this basic stuff changed into the diverse forms they experienced, and how it came to be. Those origination questions related to the puzzle of existence that, in its metaphysical dimensions, is the subject of our concern.

First, why is there anything at all? Why is there something, no matter what it is, even if different or even radically different from what currently exists? This question becomes clearer when put in contrastive form, Why is there something rather than nothing? We can ask this question even in the absence of contingent beings, though in this context it is likely to prove unanswerable. For example, if God or the universe is logically or absolutely necessary, something would not only exist but would have to exist even if nothing else existed. At the same time, probably no reason can be given for why logically necessary things exist.

Some doubt whether we can ask this question because there being nothing is not an option. John Heil asks, “What exactly is nothing at all ? What would nothing be ?” (Heil 2013: 174). He analogizes nothing with the notion of empty space, in terms of which, he thinks, we can conceptualize nothing. He reasons that we cannot achieve a notion of empty space simply by removing its contents one at a time, for space (the void) would still exist. But we need not analogize nothing in terms of empty space, and even if we do, we surely can conceive of removing space. If we think of space as a particular type of relation between objects, the removal of all objects (everything) would leave nothing, including relations. The key point is that “leaving nothing” is not to be understood in the sense that nothing is or has existence. We can easily be misled by the language of there being nothing at all, leading to the notion that nothing has being or existence. Heil suggests that nothing might be a precursor to the Big Bang. But this too is a misconception—though one widely held by those who think that the universe arose out of nothing, e.g., a vacuum fluctuation. A vacuum fluctuation is itself not nothing “but is a sea of fluctuating energy endowed with a rich structure and subject to physical laws” (Craig and Sinclair 2009: 183, 191). The contrastive question is comprehensible: “Why is there something rather than there never having been anything whatsoever?”

Rutten (2012, Other Internet Resources: 15–16), using the modal proposition S5, develops an a priori argument for the impossibility of there being nothing. Suppose that there is nothing. If there is nothing, then there are no possible states of affairs, since nothing is actual to bring them about. But since I am actual, there is at least one possible state of affairs S . But if S is possible, then by S5, necessarily, S is possible. But this contradicts the original assumption that total nothingness is metaphysically possible. Hence, total nothingness cannot be actual. Rutten’s argument rests on the Principle of Causation, about which we will have a lot more to say below.

Second, why are there contingent beings? That is, “Why does a universe exist?” Several possible responses have been given: it might not be explicable (the universe just exists; its existence is a brute fact); it has always existed, which also leaves it as a brute fact; it was brought about by the intentional act of a supernatural being. The traditional cosmological arguments consider these options and determine that the last provides the best explanation for the existence of a contingent universe.

Third, why are there these particular contingent beings? The starting point here is the existence of particular things, and the question posed asks for an explanation for there being these particular things. If we are looking for a causal explanation and accept a full explanation (in terms of contemporary or immediately prior causal conditions and the relevant natural laws or intentions that together necessitate the effect), the answer emerges from an analysis of the relevant immediate causal conditions present in each case. As Hume argues, explanation in terms of immediately conjunctory factors is satisfactory. Theists counter that if we seek a complete causal explanation where nothing of the causal event remains unexplained, the response can lead to the development of the cosmological argument.

Heil suggests that the answer depends on how one understands the Big Bang (2013: 178). If it was spontaneous, the question has no answer. If not spontaneous, there might be an answer. Theists take up the latter cause, broadening the explanatory search to include final causes or intentions appropriate to a personal cause. It leads us to ask the question, “Supposing that God exists, why did he bring about contingent beings?” This assumes that God exists and now inquires about the reasons for creation. On the one hand, we might argue that this question is unanswerable in that only God would know his reasons for bringing the universe of contingent beings into existence (O’Connor 2008). On the other hand, God acts out of his nature; Swinburne (2004: 47, 114–23) emphasizes his goodness, from which we can infer possible reasons for what he brings about. God also acts from his intentions (Swinburne 1993: 139–45; 2007: 83–84), so that God could reveal his purposes for his act of creating.

Fourth, why do things exist now or at any given point? This is the question that Thomas Aquinas posed. Aquinas was interested not in a beginning cause but in a sustaining cause, for he believed that the universe could be eternal—although he believed on the basis of revelation that it was not eternal. He constructed his cosmological arguments around the question of what sustains things in the universe in their existence.

Fifth, if the universe has a beginning, what is the cause of that beginning? This is the question that is addressed by the kalām cosmological argument, given its central premise that everything that begins to exist has a cause.

Two things should be obvious from this discussion. First, questions about existence are more nuanced than usually addressed (Heil 2013: 177). It is important to be more precise about what one is asking when one asks this broader metaphysical question about why there is something rather than nothing. Second, it becomes clear that the cosmological argument lies at the heart of attempts to answer the questions, and to this we now turn.

4. Argument for a Non-contingent Cause

Thomas Aquinas held that among the things whose existence needs explanation are contingent beings that depend for their existence upon other beings. Richard Taylor (1992: 84–94) discusses the argument in terms of the world (“everything that ever does exist, except God, in case there is a god”, 1992: 87) being contingent and thus needing explanation. Arguing that the term “universe” refers to an abstract entity or set, William Rowe rephrases the issue, “Why does that set (the universe) have the members that it does rather than some other members or none at all?” (Rowe 1975: 136). Put broadly, “Why is there anything at all?” (Smart, in Smart and Haldane, 1996: 35; Rundle 2004). The response of defenders of the cosmological argument is that what is contingent exists because of the action of a necessary being.

As an a posteriori argument, the cosmological argument begins with a fact known by experience, namely, that something contingent exists. We might sketch out a version of the argument as follows.

  • A contingent being (a being such that if it exists, it could have not-existed or could cease to exist) exists.
  • This contingent being has a cause of or explanation [ 1 ] for its existence.
  • The cause of or explanation for its existence is something other than the contingent being itself.
  • What causes or explains the existence of this contingent being must either be solely other contingent beings or include a non-contingent (necessary) being.
  • Contingent beings alone cannot provide a completely adequate causal account or explanation for the existence of a contingent being.
  • Therefore, what causes or explains the existence of this contingent being must include a non-contingent (necessary) being.
  • Therefore, a necessary being (a being such that if it exists, it cannot not-exist) exists.
  • The universe is contingent.
  • Therefore, the necessary being is something other than the universe.

In the argument, steps 1–7 establish the existence of a necessary or non-contingent being; steps 8–9 attempt in some way to identify it.

Over the centuries philosophers have suggested various instantiations for the contingent being noted in premise 1 . In his Summa Theologica (I,q.2,a.3), Aquinas argued that we need a causal explanation for things in motion, things that are caused, and contingent beings. [ 2 ] Others, such as Richard Swinburne (2004), propose that the contingent being referred to in premise 1 is the universe. The connection between the two is supplied by John Duns Scotus, who argued that even if the essentially ordered causes were infinite, “the whole series of effects would be dependent upon some prior cause” (Scotus [c. 1300] 1964: I,D.2,p.1,q.1,§53). Gale (1999) calls this the “Big Conjunctive Contingent Fact”). Whereas the contingency of particular existents is generally undisputed, not the least because of our mortality, the contingency of the universe deserves some defense (see section 4.2 ). Premise 2 invokes a moderate version of the Principle of Causation or the Principle of Sufficient Reason, according to which if something is contingent, there must be a cause of its existence or a reason or explanation why it exists rather than not exists. The point of 3 is simply that something cannot cause or explain its own existence, for this would require it to already exist (in a logical if not a temporal sense). Premise 4 is true by virtue of the Principle of Excluded Middle: what explains the existence of the contingent being either are solely other contingent beings or includes a non-contingent (necessary) being. Conclusions 6 and 7 follow validly from the respective premises.

For many critics, premise 5 (along with premise 2 ) holds the key to the argument’s success or failure. The truth of 5 depends upon the requirements for an adequate explanation. According to the Principle of Sufficient Reason (PSR), what is required is an account in terms of sufficient conditions that provides an explanation why the cause had the effect it did, or alternatively, why this particular effect and not another arose. Swinburne (2004: 75–79), and Alexander Pruss (2006: 16–18) after him, note diverse kinds of explanations. In a full explanation the causal factors—in scientific causation, causal conditions and natural laws; in personal causation, persons and their intentions— are sufficient for the occurrence of an event. They “together necessitate the occurrence of the effect” (Swinburne 2004: 76).

It does not allow a puzzling aspect of the explanandum to disappear: anything puzzling in the explanandum is either also found in the explanans or else explained by the explanans. (Pruss 2006: 17)

It suffices to explain why something comes about given the immediately present causal conditions, but leaves unexplained why those explanatory causal conditions and/or reasons themselves hold.

In a complete explanation, every aspect of the explanandum and explanans at the time of the occurrence is accounted for; nothing puzzling remains.

A complete explanation of the occurrence of E is a full explanation of its occurrence in which all the factors cited are such that there is no [further] explanation (either full or partial) of their existence of operation in terms of factors operative at the time of their existence or operation. (Swinburne 2004, 78)

Pruss and Swinburne argue that the kind of explanation required by the PSR is a complete explanation.

Quinn argues that an adequate explanation need not require a complete explanation (2005: 584–85); a partial explanation might do just as well, depending on the context. Among these adequate explanations of why this actual world obtains rather than another possible world (including one with no contingent beings) is that the universe is an inexplicable brute fact and that God strongly actualized the world (although not everything in it). He refuses to take sides on the debate between explanations, except to say that science cannot provide an adequate explanation if the explanatory chain is infinite, for the chain of causes is itself contingent or it ends in an initial contingency not scientifically accountable. However, not only does Quinn not clarify what constitutes an adequate explanation, but as Pruss contends, the PSR “is not compatible with an infinite chain of explanations that has no ultimate explanans” (2006: 17), for in an infinite chain something puzzling remains to be explained, with the result that the PSR would again be invoked to explain what is puzzling. But, as we will question below, is the brute fact of the universe any more unacceptable as a complete explanation than the brute fact of a necessary being?

One worry with understanding the PSR is that it may lead to a deterministic account that not only bodes ill for the success of the argument but on a libertarian account may be incompatible with the contention that God created freely. Pruss, however, envisions no such difficulty.

What gives sufficiency to explanation is that mystery is taken away, for example, through the citing of relevant reasons, not that probability is increased. (Pruss 2006: 157)

Giving reasons neither makes the event deterministic nor removes freedom.

Once we have said that \(x\) freely chose \(A\) for \(R\), then the only thing left that is unexplained is why \(x\) existed and was both free and attracted by \(R\). (2006: 158)

Finally, some (disputedly, see below) argue that explanations must be either natural (impersonal) or non-natural (personal). 8 & 9 assert that if the contingent being identified in 1 is the universe, given that the universe encompasses all natural existents and the laws and principles governing them, the explanation must be in terms of a non-natural, eternal, necessary being that provides an intentional, personal, ultimate explanation. Since the argument proceeds independent of temporal considerations, the argument does not necessarily propose a first cause in time, but rather a first or primary sustaining cause of the universe. As Aquinas noted, the philosophical arguments for God’s existence as first cause are compatible with the eternity of the universe ( On the Eternity of the World ).

The question whether the necessary being to which the argument concludes is God is debated. Some hold that the assertion that the first sustaining cause is God (steps 8 and 9 ) is not part of the cosmological argument per se ; such defenders of the argument sometimes create additional arguments to identify the first cause. Others, however, contend that from the concept of a necessary being other properties appropriate to a divine being flow. O’Connor (2004) argues that being a necessary being cannot be a derivative emergent property, otherwise the being would be contingent. Likewise the connection between the essential properties must be necessary. Hence, the universe cannot be the necessary being since it is mereologically complex. Similarly, the myriad elementary particles cannot be necessary beings either, for their distinguishing distributions are externally caused and hence contingent. Rather, he contends that a more viable account of the necessary being is as a purposive agent with desires, intentions, and beliefs, whose activity is guided but not determined by its goals, a view consistent with identifying the necessary being as God. Koons (as are Craig and Sinclair 2009: 192–94) also is willing to identify the necessary being as God, constructing corollaries regarding God’s nature that follow from his construction of the cosmological argument. Oppy (1999), on the other hand, expresses significant skepticism about the possibility of such a deductive move.

Critics have objected to key premises in the argument. We will consider the most important objections and responses.

Interpreting the contingent being in premise 1 as the universe, Bertrand Russell denies that the universe needs an explanation ( premise 2 ); it just is. Russell, following Hume (1779), contends that since we derive the concept of cause from our observation of particular things, we cannot ask about the cause of something like the universe that we cannot experience. The universe needs no explanation; it is “just there, and that’s all” (Russell 1948 [1964]: 175). This view was reiterated by Hawking (1987: 651).

Swinburne replies that

uniqueness is relative to description. Every physical object is unique under some description,… yet all objects within the universe are characterized by certain properties, which are common to more than one object.… The objection fails to make any crucial distinction between the universe and other objects; and so it fails in its attempt to prevent at the outset a rational inquiry into the issue of whether the universe has some origin outside itself. (Swinburne 2004: 134–35)

We don’t need to experience every possible referent of the class of contingent things to be able to conclude that a contingent thing needs a cause. “To know that a rubber ball dropped on a Tuesday in Waggener Hall by a redheaded tuba player will fall to the ground”, I don’t need a sample that includes tuba players dropping rubber balls at this location (Koons 1997: 202).

Morriston (2002: 235) responds that although it is true that we don’t need to experience every instance to derive a general principle, the universe is a very different thing from what we experientially reference when we say that things cannot come into existence without a cause. Tuba players are not “anything remotely analogous to the ‘initial singularity’ that figures in the Big Bang theory of the origin of the universe”.

Defenders of the argument respond that there is a key similarity between the universe and the experienced content, namely, both tuba players (and the like) and the cosmos are contingent. Given our experience with contingent particulars, we do not need to experience every member of a contingent cosmos to know that it is caused.

But why should we think that the cosmos is contingent? Defenders of the view contend that if the components of the universe are contingent, the universe itself is contingent. Russell replies that the move from the contingency of the components of the universe to the contingency of the universe commits the Fallacy of Composition, which mistakenly concludes that since the parts have a certain property, the whole likewise has that property. Hence, whereas we legitimately can ask for the cause of particular things, to require a cause of the universe or the set of all contingent beings based on the contingency of its parts is mistaken.

Russell correctly notes that arguments of the part-whole type can commit the Fallacy of Composition. For example, the argument that since all the bricks in the wall are small, the wall is small, is fallacious. Yet it is an informal fallacy of content, not a formal fallacy. Sometimes the totality has the same quality as the parts because of the nature of the parts invoked—the wall is brick (composed of baked clay) because it is built of bricks (composed of baked clay). The universe’s contingency, theists argue, resembles the second case. If all the contingent things in the universe, including matter and energy, ceased to exist simultaneously, the universe itself, as the totality of these things, would cease to exist. But if the universe can cease to exist, it is contingent and requires an explanation for its existence (Reichenbach 1972: chap. 5).

Whether this argument for the contingency of the universe is similar to that advanced by Aquinas in his Third Way depends on how one interprets Aquinas’s argument. Aquinas holds that “if everything can not-be, then at one time there was nothing in existence” ( ST I,q.2,a.3). Plantinga, among others, points out that this may commit a quantifier mistake (Plantinga 1967: 6; Kenny 1969: 56–66). However, Haldane (Smart and Haldane 1996: 132) defends the cogency of Aquinas’s reasoning on the grounds that Aquinas’s argument is fallacious only on a temporal reading, but Aquinas’s argument employs an atemporal ordering of contingent beings. That is, Aquinas does not hold that over time there would be nothing, but that in the per se ordering of causes, if every contingent thing in that order did not exist, there would be nothing.

William Rowe treats the argument temporally and contends (1975: 160–67) that this argument for the contingency of the universe is fallacious, for even if every contingent being were to fail to exist in some possible world, it may be the case that there is no possible world that lacks a contingent being. That is, although no being would exist in every possible world, every world would possess at least one contingent being. In such a case, although each being is contingent, it is necessary that something exist. Rowe gives the example of a horse race.

We know that although no horse in a given horse race necessarily will be the winner, it is, nevertheless, necessary that some horse in the race will be the winner. (1975: 164)

Rowe’s example will work only if it is necessary that some horse will finish the race, for otherwise it is possible that all the horses break a leg and none finishes the race—a condition he notes in that “it is necessary that there exists at least one member of the collection”. So why should we think that it is necessary that something exist, even if it is contingent? Rowe does not say why, but one argument given in defense of this thesis is that the existence of one contingent being may be necessary for the nonexistence of some other contingent being. But although the fact that something’s existence is necessary for the non-existence of something else holds for certain relational properties (for example, the existence of a spouse is necessary for a man not to be a bachelor), it is doubtful that something’s existence is necessary for the non-existence of something else per se , which is what is needed to support the argument that denies the contingency of the universe.

Rowe (1975: 166) develops a different argument to support the thesis that the universe must be contingent. He argues that it is necessary that if God exists, then it is possible that no dependent beings exist. Since it is possible that God exists, it is possible that it is possible that no dependent beings exist. (This conclusion is licensed by the modal principle: If it is necessary that if \(p\) then \(q\), then if it is possible that \(p\), it is possible that \(q\).) Hence, it is possible that there are no dependent beings; that is, that the universe is contingent. Rowe takes the conditional as necessarily true in virtue of the classical concept of God, according to which God is free to decide whether or not to create dependent beings.

To avoid any hint of the Fallacy of Composition and to avoid its complications, Koons (1997: 198–99) formulates the argument for the contingency of the universe as a mereological argument. If something is contingent, it contains a contingent part. The whole and part overlap and, by virtue of overlapping, have a common part. Since the part in virtue of which they overlap is wholly contingent, the whole likewise must be contingent.

One might approach Russell’s thesis regarding the universe from a different direction. If theists are willing to accept the existence of God as the necessary being as a brute fact, why cannot nontheists accept the existence of the universe as a brute fact, as a necessary being? Bede Rundle, for example, argues that what has necessary existence is causally independent. Matter has necessary existence, for although it undergoes change as manifested in particular bits of matter, the given volume of matter found in the universe persists, and as persisting matter/energy does not have or need a cause. This accords with the Principle of Conservation of Mass-Energy, according to which matter and energy are never lost but rather transmute into each other. As indestructible, matter/energy is the necessary being. Consequently, although the material components of the universe are contingent vis-à-vis their form, they are necessary vis-à-vis their existence. On this reading, there is not one but there are many necessary beings, all internal to the universe. Their particular configurations are contingent, but since matter/energy is conserved it cannot be created or lost.

Interestingly enough, this approach was anticipated by Aquinas in his third way in his Summa Theologica (I,q.2,a.3). Once Aquinas concludes that necessary beings exist, he then goes on to ask whether these beings have their existence from themselves or from another. If from another, then we have an unsatisfactory infinite regress of explanations. Hence, there must be something whose necessity is uncaused. As Kenny points out, Aquinas understands this necessity in terms of being unable to cease to exist (Kenny 1969: 48). Although Aquinas understands the uncaused necessary being to be God, Rundle takes this to be matter/energy itself.

One question that arises with Rundle’s view is whether there could have been more or less matter/energy than there is. That is, if there is \(n\) amount of matter/energy in the world, could there be a possible world with \(+n\) or \(-n\) amounts of matter/energy? We do not know how much matter/energy existed in the first \(10^{-35}\) seconds of the universe. Even if the universe currently operates according to the principle of the Conservation of Matter and Energy, Rundle’s thesis depends on the contention that during the very early phase of rapid expansion, a period of time we know little about, this principle held. A second significant problem concerns what follows from the existence of necessary beings. If the matter/energy nexus constitutes the necessary being, what causally follows from that nexus is itself necessary, and contingency, even in the composing relations within the universe, would disappear. Everything in the universe would be necessary, which is a disquieting position. Third, O’Connor (2004) argues that since the necessary being provides the ultimate explanation, there is no explanation of the differentiation of the kinds of matter or of contingencies that matter/energy causally undergo, for example, in terms of space-time location. Perhaps one way to rescue Rundle’s thesis would be to invoke an indeterministic presentation of quantum phenomena, which would allow contingency of individual phenomena but not of the overall probabilistic structure.

Whereas Russell argued that the universe just is, David Hume held that when the parts are explained the whole is explained.

But the whole , you say, wants a cause. I answer that the uniting of these parts into a whole… is performed merely by an arbitrary act of the mind, and has no influence on the nature of things. Did I show you the particular causes of each individual in a collection of twenty particles of matter, I should think it very unreasonable should you afterwards ask me what was the cause of the whole twenty. This is sufficiently explained in explaining the parts. (Hume 1779: part 9)

Hume contends that uniting the parts or individual constituents into a whole is a mental act. In reality, all that exist are individual, causally-related events, not whole sets of events. When we have provided an account of each of these individual, causally-related events we have explained the whole. We don’t need anything more.

Rowe objects to what he terms the Hume-Edwards principle—that by explaining the parts we have explained the whole:

When the existence of each member of a collection is explained by reference to some other member of that very same collection , then it does not follow that the collection itself has an explanation. For it is one thing for there to be an explanation of the existence of each dependent being and quite another thing for there to be an explanation of why there are dependent beings at all. (Rowe 1975: 264)

Pruss (1999) expands on Rowe’s argument. An explanation of the parts may provide a partial but not a complete explanation. The explanation in terms of parts may fail to explain why these parts exist rather than others, why they exist rather than not, or why the parts are arranged as they are. Each member or part will be explained either in terms of itself or in terms of something else that is contingent. The former would make them necessary, not contingent, beings. If they are explained in terms of something else, they still remain unaccounted for, since the explanation would invoke either an infinite regress of causes or a circular explanation. Pruss employs the chicken/egg sequence: chickens account for eggs, which account for chickens, and so on where the two are paired. But appealing to an infinite chicken/egg regress or else arguing in a circle explains neither any given chicken nor egg.

Richard Swinburne notes that an explanation is complete when “any attempt to go beyond the factors which we have would result in no gain of explanatory power or prior probability” (2004: 89). But explaining why something exists rather than something else or than nothing and why it is as it is gives additional explanatory power in explaining why a universe exists at all. Gale (1991: 257–58) concludes that if we are to explain the parts of the universe and their particular concatenation, we must appeal to something other than those parts.

Critics of the cosmological argument contend that the Causal Principle or, where applicable, the broader Principle of Sufficient Reason (PSR) that underlies versions of the argument, is suspect. As Hume argued, there is no reason for thinking that the Causal Principle is true a priori , for we can conceive of events occurring without conceiving of their being caused, and what is conceivable is possible in reality (1748: IV). Neither can an argument for the application of the Causal Principle to the universe be drawn from inductive experience. Even if the Causal Principle applies to events in the world, we cannot extrapolate from the way the world works to the world as a whole (Mackie 1982: 85).

Several replies are in order. First, Hume’s conceivability to possibility argument is unsound. For one thing, whose conceivability is being appealed to here? For another, someone who fails to understand a necessarily true proposition might conceive of it being false, but from this it does not follow that it possibly is false. A person might think (wrongly) that pi is a determinate number, but it does not follow that it is so. In the phenomenology of conceivability, what is really conceivable is difficult if not impossible to differentiate from what some might think is conceivable. And even if something is conceivable, say in a logical sense, it does not follow that it is metaphysically or factually possible. One might conceive that, since heads can be distinguished from tails on a coin, they can actually be separated, but metaphysically such is impossible. What is distinguishable is not necessarily separable. Hume, it seems, confuses epistemic with ontological conditions. Hence, the argument based on conceivability is suspect (Reichenbach 1972: 57–60).

Second, there is reason to think that the Causal and Sufficient Reason principles are true. Some suggest a pragmatic-type of argument: the principles are necessary to make the universe intelligible. Critics reply that the principles then only have methodological or practical and not ontological justification. As John Mackie argues, we have no right to assume that the universe complies with our intellectual preferences for causal order. We can simply work with brute facts. Perhaps so, but without such principles, science itself would be undercut. As Pruss (2006: 255) points out, “Claiming to be a brute fact should be a last resort. It would undercut the practice of science”. Utilization of the principles best accounts for the success of science, indeed, for any investigatory endeavor (Koons 1997; see also Koons 2008: 111–12, where he argues that it is “a subjectively required presumption for needed for immunity to internal defeaters”). The best explanation of the success of science and other such rational endeavors is that the principles are really indicative of how reality operates.

Pruss goes further to suggest that the PSR in particular is “self-evident, obvious, intuitively clear, in no need of argumentative support” (2006: 189). For example, he holds that the principle of sufficient reason—“necessarily, every contingently true proposition has an explanation” (he defers on whether the principle also applies to necessarily true propositions)—is self-evident in the sense that anyone who understands it correctly understands that it is true. These persons might not know it to be self-evidently true, but they do understand it to be true. This is consistent with other persons denying it is self-evident, for those who deny it might misunderstand the principle in various ways. They might experience a conceptual blindness to the nature of contingency or they might be “talked out of” understanding the principle because of its controversial implications (e.g., the existence of a necessary being).

The problem with the claim of self-evidence is that it is a conversation ender, not a starter. One who denies its self-evidence might think that those who hold to it are the ones who experience conceptual blindness. In contrast to analyticity, self-evidence holds in relation to the knowers themselves, and here diversity of intuitions varies, perhaps according to philosophical or other types of perspectives. Furthermore, if the principle truly is self-evident, it would be strange to respond to skeptics by attempting to give reasons to support that contention, and were such demanded, the request would itself invoke the very principle in question.

Pruss responds that being self-evident is not incompatible with providing arguments for self-evident propositions, and he thinks that arguments can show the truth of the PSR to those who deny its self-evidence. Among the numerous arguments he advances is a modal argument employing a Weak Principle of Sufficient Reason, according to which “every contingent proposition possibly has a complete explanation” (Pruss 2006: 234–35). We will develop this in section 5 .

Peter van Inwagen (1983: 202–04) argues that the PSR must be rejected. If the PSR is true, every contingent proposition has an explanation. Suppose P is the conjunction of all contingent true propositions. Suppose also that there is a state of affairs S that provides a sufficient reason for P . S cannot itself be contingent, for then it would be a conjunct of P and entailed by P , and as both entailing and entailed by P would be P , so that it would be its own sufficient reason. But no contingent proposition can explain itself. Neither can S be necessary, for from necessary propositions only necessary propositions follow. Necessary propositions cannot explain contingent propositions, for if x sufficiently explains y , then x entails y , and if x is necessary so is y . So \(S\) cannot be either contingent or necessary, and hence the PSR is false. Thus, if the cosmological argument appeals to the PSR to establish the existence of a necessary being whose existence is expressed by a necessary proposition as an explanation for contingent beings, it fails in that it cannot account for the contingent beings it purportedly explains. But, as Pruss notes (2006: chaps. 6 & 7),

The word sufficient can be read in two different ways: the reason given can be logically sufficient for the explanandum, or it can sufficiently explain the explanandum. (2006: 103)

We need not hold to the strong claim of logical sufficiency about the relation between explaining and entailment in cases where the explanation is brought about by libertarian free agency. Although God is a necessary being, his connection with the world is through his free agency, and free actions explain but do not entail the existence of particular contingent states.

Clearly, the soundness of the deductive version of the cosmological argument hinges on whether principles such as that of Causation or Sufficient Reason are more than methodologically true and on the extent to which these principles can be applied to individual things or to the universe. Critics of the argument will be skeptical regarding the universal application of the principles; defenders of the argument generally not so. Perhaps the best one can say, with Taylor, is that even those who critique the PSR invoke it when they suggest that defenders have failed to provide a sufficient reason for thinking it is true.

The principle of sufficient reason can be illustrated in various ways,…but it cannot be proved…. If one were to try proving it, he would sooner or later have to appeal to considerations that are less plausible than the principle itself. Indeed, it is hard to see how one could even make an argument for it without already assuming it. For this reason it might properly be called a presupposition of reason itself. (1992: 87)

We will return to the Principle of Causation below with respect to the kalām argument.

Immanuel Kant objected to the use of “necessary being” throughout the cosmological argument, and hence to the conclusion that a necessary being exists. Kant held that the cosmological argument, in concluding to the existence of an absolutely necessary being, attempts to prove the existence of a being whose nonexistence “is impossible”, is “absolutely inconceivable” ( Critique B621). Kant indicates that what he has in mind by an “absolutely necessary being” is a being whose existence is logically necessary, where to deny its existence is contradictory. The only being that meets this condition is the most real or maximally excellent being—a being with all perfections, including existence. This concept lies at the heart of the ontological argument (see entry on ontological arguments ). Although in the ontological argument the perfect being is determined to exist through its own concept, in fact nothing can be determined to exist in this manner; one has to begin with existence. In short, the cosmological argument presupposes the cogency of the ontological argument. But since the ontological argument is defective for the above (and other) reason, the cosmological argument that depends on or invokes it likewise must be defective ( Critique B634).

Kant’s contention that the necessity found in “necessary being” is logical necessity was common up through the 1960s. J.J.C. Smart wrote,

And by “a necessary being” the cosmological argument means “a logically necessary being”, i.e., “a being whose non-existence is inconceivable in the sort of way that a triangle’s having four sides is inconceivable”.… Now since “necessary” is a word which applies primarily to propositions, we shall have to interpret “God is a necessary being” as “The proposition ‘God exists’ is logically necessary”. (in Flew and MacIntyre 1955: 38; in a later work Smart (Smart and Haldane 1996: 41–47) broadened his notion of necessity.)

Many recent discussions of the cosmological argument, both supporting and critiquing it, interpret the notion of a necessary being as a being that cannot not exist (O’Connor 2008: 78, 2013: 38). For example, Gale-Pruss contend that speaking about necessary beings does not differ from speaking of the necessity of propositions (see section 5 ). As such, as Plantinga notes, if a necessary being is possible, it exists ( God, Freedom and Evil , 1967: 110). It is a being that exists in all possible worlds. The only question that remains is whether God’s existence is possible. This notion is similar to, if not a modernization of, Aquinas’s contention that God’s essence is to exist. Aquinas attempts to avoid the accusation that this invokes the ontological argument on the grounds that we do not have an adequate concept of God’s essence ( ST I,q.2,a.1). However, if we understand “necessary being” in this sense, we can dispose of the cosmological argument as irrelevant; what is needed rather is an argument to establish that God’s existence is possible, for if it is possible that it is necessary that God exists, then God exists (by Axiom S5).

But this need not be the sense in which “necessary being” is understood in the cosmological argument. A more adequate notion of necessary being is that the necessity is metaphysical or factual. A necessary being is one that if it exists, it neither came into existence nor can cease to exist, and correspondingly, if it does not exist, it cannot come into existence (Reichenbach 1972: 117–20). If it exists, it eternally maintains its own existence; it is self-sufficient and self-sustaining. So understood, the cosmological argument does not rely on notions central to the ontological argument. Rather, instead of being superfluous, the cosmological argument, if sound, gives us reason to think that the necessary being exists rather than not.

Mackie replies that if God has metaphysical necessity, God’s existence is logically contingent, such that some reason is required for God’s own existence (Mackie 1982: 84). That is, if God necessarily exists in the sense that if he exists, he exists in all possible worlds, it remains logically possible that God does not exist in any (and all) possible worlds. Hence, God, as Swinburne notes (2004: 79, 148) is a logically contingent being, and so could have not-existed. Why, then, does God exist? The PSR can be applied to the necessary being.

The theist responds that the PSR does not address logical contingency but metaphysical contingency. One is not required to find a reason for what is not metaphysically contingent. It is not that the necessary being is self-explanatory; rather, a demand for explaining its existence is inappropriate. Hence, the theist concludes, Hawking’s question “Who created God?” (Hawking 1988: 174) is out of place (Davis 1997). We will return to this discussion in section 8 .

Recently Richard Gale and Alexander Pruss (1999) advanced a modal version of the cosmological argument. They reject the strong version of the PSR, according to which “for every proposition \(p\), if \(p\) is true, then there is a proposition, \(q\), that explains \(p\)”. In its place they favor using a weak version of the PSR—it is possible that for every true proposition, there is a proposition, \(q\), that explains \(p\)—that they believe is less question-begging and more initially acceptable to critics. They phrase the argument in terms of contingent and necessary propositions. A contingent proposition is one that is both possibly true and possibly false (i.e., true in some worlds and false in others); a necessarily true proposition is true in every possible world. In its simplest form, the argument is (1) if it is possible that it is necessary that a supernatural being of some sort exists, then it is necessary that a supernatural being of that sort exists. Since (2) it is possible that it is necessary that a supernatural being of some sort exists, (3) it is necessary that this being exists. The being that Gale has in mind is a very powerful and intelligent designer-creator, not the all perfect God of Anselm, for this perfect God who would exist in all possible worlds would be incompatible with the existence of gratuitous and horrendous evils to be found in some of those possible worlds.

If one grants modal Axiom S5 (if it is possible that it is necessary that \(p\), then it is necessary that \(p\)), the critical premise in the argument is the second, and Gale and Pruss proceed to defend it using their weak PSR. They begin with the notion of a Big Conjunctive Fact (BCF), which is the totality of propositions that would be true of any possible world were it actualized. Since all possible worlds would have the same necessary propositions, they are differentiated by their Big Conjunctive Contingent Fact (BCCF), which would contain different contingent propositions. Let \(p\) be the BCCF of the actual world \(W\). Suppose, further, that it is possible that \(p\) has an explanation, that is, that it is possible that there is some proposition \(q\) that explains \(p\). As such, there is a possible world \(W_{1}\) that contains \(p\), \(q\), and the proposition that \(q\) explains \(p\). The question now is whether \(W_{1}\) is the actual world, that is, whether there is a proposition \(q\) that explains \(p\) in the actual world. Gale argues that \(W_{1}\) (which contains \(q\) and the proposition that \(q\) explains \(p\)) is the actual world, for since \(W_{1}\) contains \(p\), there can be no property of \(p\) that is not found in \(W_{1}\). Every conjunct of \(p\) will be a conjunct of \(p_{1}\) (the BCCF of \(W_{1}\)) in \(W_{1}\). Suppose that \(r\) is a conjunct of \(p\) in \(W\), then not-\(r\) cannot be a conjunct of \(p_{1}\) in \(W_{1}\)\(.\) Since \(W\) and \(W_{1}\) have the same properties, \(W_{1}\) is the actual world. Therefore, since these worlds are identical, the actual world contains \(p\), \(q\), and the proposition that \(q\) explains \(p\). That is, there is something that explains the BCCF of the actual world. The explanation of the BCCF cannot be scientific, for such would be in terms of law-like propositions and statements about the actual world at a given time, which would be contingent and hence part of the BCCF. Since the only explanations we can conceive of are personal or scientific, \(q\) provides a personal explanation of the BCCF in terms of the intentional action of a necessary being who freely brings it about that the world exists. \(q\) cannot report the action of a contingent being, for then the being would be part of \(p\) and explained by \(q\). But something cannot explain itself. Hence, although contingent, \(q\) reports the action of a necessary being. Gale concludes that although this necessary being exists in every possible world, this tells little about its power, goodness, and other qualities. To make this being palatable to theists, he offers that the argument be supplemented by other arguments, such as the teleological arguments, to suggest that the necessary being is the kind of being that satisfies theistic requirements. Since

the actual world’s universe displays a wondrous complexity due to its law-like unity and simplicity, fine tuning of natural constants, and natural purpose and beauty,… there exists a necessary supernatural being who is very powerful, intelligent, and good and freely creates the actual world’s universe. (1999: 468–69)

(For the detailed 18 step deductive argument, see Gale and Pruss 1999: 462–69).

Several objections have been raised about the argument from the weak principle of sufficient reason. Almeida and Judisch construct their objection via two reductio arguments. They note that, according to Gale’s argument, \(q\) is a contingent proposition in the actual world that reports the free, intentional action of a necessary being. As such, since the actual world contains the contingent proposition \(q\), non-\(q\) is possible. That is, there is a possible world \(W_{2}\) that contains \(p\), non-\(q\), and that \(q\) does not explain \(p\). But by Gale’s own reasoning, \(W_{2}\) is identical to the actual world. But the actual world cannot contain both \(q\) and non-\(q\). Thus, \(q\) cannot be a contingent proposition.

On the other hand, assume that \(q\) is a contingently necessary proposition, that is, that it is possible that \(q\) is necessary and possible that \(q\) is not necessary. By S5, we get that it is necessary that \(q\) is necessary, making it impossible that \(q\) is not necessary. As a result, it is both possible and not-possible that \(q\) is not necessary, which likewise shows that \(q\) cannot be a contingently necessary proposition. The only other option is that \(q\) is a necessary truth, which would beg the question. Thus, the argument fails by being unable to characterize \(q\). For rebuttals, see Gale and Pruss (2002) and Rutten (2012, Other Internet Resources: 84–87).

Graham Oppy (2000) similarly argues that suppose \(p_1\) is the BCF of some possible world, and \(p_{1}\) has no explanation. Then, given \(r\) (namely, that \(p_{1}\) has no explanation) there is a conjunctive fact \(p_{1}\) and \(r\). Since by hypothesis the conjunctive fact \(p_{1}\) and \(r\) is true in some world, on Gale’s account it is true in the actual world. Then by the weak PSR there is a world in which this conjunction of \(p_{1}\) and \(r\) possibly has an explanation. If there is an explanation for the conjunction of \(p_{1}\) and \(r\), there is an explanation for \(p_{1}\). Thus, we have the contradiction that \(p_{1}\) both has and does not have an explanation, which is absurd. Hence, no world exists where the BCF lacks an explanation, which is the strong principle of sufficient reason that Gale allegedly circumvented. Since accepting the weak PSR would commit the nontheist to the strong PSR and ultimately to a necessary being, the nontheist has no motivation to accept the weak PSR.

Gale and Pruss (2002) subsequently concede that their weak PSR does entail the strong PSR, but they contend that there still is no reason not to proceed with the weak PSR, which they think the nontheist would accept. The only grounds for rejecting it, they claim, is that it leads to a theistic conclusion, which is not an independent reason for rejecting it. Oppy, however, maintains that appealing to some initial instincts of acceptance is irrelevant. Perhaps the nontheists did not see what granting the weak PSR entailed, that it contradicted other things they had independent reasons to believe, or they did not fully understand the principle. There is a modus tolens reason to reject it, since there are other grounds for thinking that theism is false.

Jerome Gellman has argued that the Gale/Pruss conclusion to a being that is not necessarily omnipotent also fails; this being is essentially omnipotent and, if omnipotence entails omniscience, is essentially omniscient. This too Gale and Pruss concede, which means that the necessary being they conclude to is not significantly different from that arrived at by the traditional cosmological argument that appeals to the moderate version of the PSR.

Finally, there is doubt that Gale’s rejection of the traditional cosmological argument on the grounds that the necessary being could not be necessarily good is well grounded. Gale argues that since there are possible worlds with gratuitous or horrendous evils, and since God as necessary would exist in these worlds, God cannot be necessarily good. The problem here is that if indeed there is this incompatibility between a perfectly good necessary being (God) and gratuitous evils or even absolutely horrendous evils, then it would follow that worlds with God and such evils would not be possible worlds, for they would contain a contradiction. In all possible worlds where a perfectly good God as a necessary being would exist, there would be a justificatory morally sufficient reason for the evils that would exist, or at least, given the existence of gratuitous evils, for the possibility of the existence of such evils (Reichenbach 1982: 38–39). Beyond this, however, the point stands that the weak PSR entails the strong PSR, and as we argued above, defenders of the cosmological argument do not need such a strong version of the PSR to construct their argument.

6. The Kalām Cosmological Argument

A second type of cosmological argument, contending for a first or beginning cause of the universe, has a venerable history, especially in the Islamic mutakalliman tradition. Although it had numerous defenders through the centuries, it received new life in the recent voluminous writings of William Lane Craig. Craig formulates the kalām cosmological argument this way (in Craig and Smith 1993: chap. 1):

  • Everything that begins to exist has a cause of its existence.
  • The universe began to exist.
  • Therefore, the universe has a cause of its existence.
  • Since no scientific explanation (in terms of physical laws) can provide a causal account of the origin (very beginning) of the universe, the cause must be personal (explanation is given in terms of a personal agent).

This argument has been the subject of much recent debate, only some of which we can summarize here. (For greater bibliographic detail, see Craig and Sinclair 2009.)

The basis for the argument’s first premise is the Causal Principle that undergirds many cosmological arguments. (Oderberg [2002: 308] is mistaken when he tries to establish the uniqueness of the kalām argument by denying that the Causal Principle plays a role in kalām argument. It only does not play a role in supporting a particular premise in the argument.) Defenders and critics alike suggest that basing the argument on the Principle of Causation rather than on the more general Principle of Sufficient Reason is advantageous to the argument (Morriston 2000: 149). Craig holds that the first premise is intuitively obvious; no one, he says, seriously denies it (Craig, in Craig and Smith 1993: 57). Although at times Craig suggests that one might treat the principle as an empirical generalization based on our ordinary and scientific experiences (which might not be strong enough for the argument to succeed in a strong sense, although it might be supplemented by an inference to the best explanation argument that what best explains the success of science is that reality operates according to the causal principle), ultimately, he argues, the truth of the Causal Principle rests “upon the metaphysical intuition that something cannot come out of nothing” (Craig, in Craig and Smith 1993: 147). “No one sincerely believes that things, say, a horse or an Eskimo village, can just pop into being without a cause” (Craig and Sinclair 2009: 182), and this includes the universe.

The Causal Principle has been the subject of extended criticism. (We addressed objections to the Causal Principle as subsumed under the PSR from a philosophical perspective earlier in 4.4 .) Some critics deny that they share Craig’s intuitions about the Causal Principle (Oppy 2002). Morriston (2000) argues that, for one thing, it does not seem to be an a priori truth, for not only does it lack “a kind of ‘luminosity’ that makes it impossible not to believe it, but closer inspection does not make it clearer that it is true” (2000: 156–59). He points not only to the presence of serious doubters (which he thinks he should not be able to find if it were truly an a priori truth), but also to quantum phenomena, and thereby joins those who raise objections to the Causal Principle based on quantum physics (Davies 1984: 200). On the quantum level, the connection between cause and effect, if not entirely broken, is to some extent loosened. For example, it appears that electrons can pass out of existence at one point and come back into existence elsewhere. One can neither trace their intermediate existence nor determine what causes them to come into existence at one point rather than another. Neither can one precisely determine or predict where they will reappear; their subsequent location is only statistically probable given what we know about their antecedent states. Hence,

quantum-mechanical considerations show that the causal proposition is limited in its application, if applicable at all, and consequently that a probabilistic argument for a cause of the Big Bang cannot go through. (Smith, in Craig and Smith 1993: 121–23, 182)

Craig responds that appeals to quantum phenomena do not affect the kalām argument. For one thing, quantum events are not completely devoid of causal conditions. Even if one grants that the causal conditions are not jointly sufficient to determine the event, at least some necessary conditions are involved in the quantum event. But when one considers the beginning of the universe, he notes, there are no prior necessary causal conditions; simply nothing exists (Craig, in Craig and Smith 1993: 146; see Koons 1997: 203). Pruss (2006: 169) contends that in quantum phenomena causal indeterminacy is compatible with the causal principle in that the causes indeterministically bring about the effect. Morriston is rightly puzzled by this reply, for, he asks, what

makes a cause out of a bunch of merely necessary conditions. Apparently not that they are jointly sufficient to produce the effect. (2000: 158)

If conditions are not jointly sufficient, is there reason to think that premise 1 is true? More recently, Craig argues that

not all physicists agree that subatomic events are uncaused…. Indeed, most of the available interpretations of the mathematical formulation of [Quantum Mechanics] are fully deterministic. (Craig and Sinclair 2009: 183)

For another, a difference exists between predictability and causality. It is true that, given Heisenberg’s principle of uncertainty, we cannot precisely predict individual subatomic events. What is debated is whether this inability to predict is due to the absence of sufficient causal conditions, or whether it is merely a result of the fact that any attempt to precisely measure these events alters their status. The very introduction of the observer into the arena so affects what is observed that it gives the appearance that effects occur without sufficient or determinative causes. But we have no way of knowing what is happening without introducing observers into the situation and the changes they bring. In the above example, we simply are unable to discern the intermediate states of the electron’s existence apart from introducing conditions of observation. When Heisenberg’s indeterminacy is understood not as describing the events themselves but rather our knowledge of the events, the Causal Principle still holds and can still be applied to the initial singularity, although we cannot expect to achieve any kind of determinative predictability about what occurs on the sub-atomic level given the cause.

At the same time, it should be recognized that showing that indeterminacy is a real feature of the world at the quantum level would have significant negative implications for the more general Causal Principle that underlies the deductive cosmological argument. The more this indeterminacy has ontological significance, the weaker is the Causal Principle. If the indeterminacy has merely epistemic significance, it scarcely affects the Causal Principle. Quantum accounts allow for additional speculation regarding origins and structures of universes. In effect, whether Craig’s response to the quantum objection succeeds depends upon deeper issues, in particular, the epistemic and ontological status of quantum indeterminacy, the nature of the Big Bang as a quantum phenomenon, the nature and role of indeterminate causation, and whether realist theories about quantum phenomena have serious traction. Quantum physics is murky, as evidenced by Bell’s gedanken experiments, as described by Mermin.

In defense of premise 2, Craig develops both a priori and a posteriori arguments. His primary a priori argument is

  • An actual infinite cannot exist.
  • A beginningless temporal series of events is an actual infinite.
  • Therefore, a beginningless temporal series of events cannot exist.

Since (7) follows validly, if (5) and (6) are true the argument is sound. In defense of premise (5), he defines an actual infinite as a determinate totality that occurs when a part of a system can be put into a one-to-one correspondence with the entire system (Craig and Sinclair 2009: 104). Craig argues that if actual infinites that neither increase nor decrease in the number of members they contain were to exist in reality, we would have rather absurd consequences. For example, imagine a library with an actually infinite number of books. Suppose that the library also contains an infinite number of red and an infinite number of black books, so that for every red book there is a black book, and vice versa. It follows that the library contains as many red books as the total books in its collection, and as many red books as black books, and as many red books as red and black books combined. But this is absurd; in reality the subset cannot be equivalent to the entire set. Likewise, in a real library by removing a certain number of books we reduce the overall collection. But if infinites are actual, a library with an infinite number of books would not be reduced in size at all by removal of a specific number of books (short of all of them), for example, all the red books or those with even catalogue numbers (Craig and Smith 1993: 11–16). The absurdities resulting from attempting to apply basic arithmetical operations, functional in the real world, to infinities suggest that although actual infinites can have an ideal existence, they cannot exist in reality.

Craig’s point is this. Two sets \(A\) and \(B\) are the same size just in case they can be put into one-to-one correspondence, that is, if and only if every member of \(A\) can be correlated with exactly one member of \(B\) in such a way that no member of \(B\) is left out. In the case of infinite sets, this notion of “same size” yields results like the following: the set of all natural numbers (let this be \(A\)) is the same size as the set of squares of natural numbers (\(B\)), since every member of \(A\) can be correlated with exactly one member of \(B\) in a way that leaves out no member of \(B\) (correlate \(0\leftrightarrow 0\), \(1\leftrightarrow1\), \(2\leftrightarrow4\), \(3\leftrightarrow9\), \(4\leftrightarrow 16\),…). So this is a case—recognized in fact as early as Galileo ( Dialogues Concerning Two New Sciences )—where two infinite sets have the same size but, intuitively, one of them, as a subset, appears to be smaller than the other; one set consists of only some of the members of another, but you nonetheless never run out of either when you pair off their members.

Craig uses a similar, intuitive notion of “smaller than” in his argument concerning the library. It appears that the set \(B\) of red books in the library is smaller than the set \(A\) of all the books in the library, even though both have the same (infinite) size. Craig concludes that it is absurd to suppose that such a library is possible in actuality , since the set of red books would simultaneously have to be smaller than the set of all books and yet equal in size.

Critics fail to be convinced by these paradoxes of infinity. For example, Rundle (2004: 170) agrees with Craig that the concept of an actual infinite is paradoxical, but this, he argues, provides no grounds for thinking it is incoherent. The logical problems with the actual infinite are not problems of incoherence, but arise from the features that are characteristic of infinite sets. When the intuitive notion of “smaller than” is replaced by a precise definition, finite sets and infinite sets just behave somewhat differently, that is all. Cantor, and all subsequent set theorists, define a set \(B\) to be smaller than set \(A\) (i.e., has fewer members) just in case \(B\) is the same size as a subset of \(A\), but \(A\) is not the same size as any subset of \(B\). The application of this definition to finite and infinite sets yields results that Craig finds counter-intuitive but which mathematicians see as our best understanding for comparing the size of sets. They see the fact that an infinite set can be put into one-to-one correspondence with one of its own proper subsets as one of the defining characteristics of an infinite set, not an absurdity. Say that set \(C\) is a proper subset of \(A\) just in case every element of \(C\) is an element of \(A\) while \(A\) has some element that is not an element of \(C\). In finite sets, but not necessarily in infinite sets, when set \(B\) is a proper subset of \(A\), \(B\) is smaller than \(A\). But this doesn’t hold for infinite sets—as above where \(B\) is the set of squares of natural numbers and \(A\) is the set of all natural numbers.

Cantorian mathematicians argue that these results apply to any infinite set, whether in pure mathematics, imaginary libraries, or the real world series of concrete events. Thus, Smith argues that Craig begs the question by wrongly presuming that an intuitive relationship holds between finite sets and their proper subsets, namely, that a set has more members than its proper subsets must hold even in the case of infinite sets (Smith, in Craig and Smith 1993: 85). So while Craig thinks that Cantor’s set theoretic definitions yield absurdities when applied to the world of concrete objects, which entails that infinites cannot be actual, set theorists see no problem so long as the definitions are maintained. Further discussion is in Oppy 2006: 137–54.

Why should one think premise (6) is true—that a beginningless series, such as the universe up to this point, is an actual rather than a potential infinite? For Craig, an actual infinite is a determinate totality or a completed unity, whereas the potential infinite is not. Since the past events of a beginningless series can be conceptually collected together and numbered, the series is a determinate totality (1979: 96–97). And since the past is beginningless, it has no starting point and is infinite. If the universe had a starting point, so that events were added to or subtracted from this point, we would have a potential infinite that increased through time by adding new members. The fact that the events do not occur simultaneously is irrelevant.

Bede Rundle rejects an actual infinite, but his grounds for doing so—the symmetry of the past and the future—, if sustained, make premise 6 false. He argues that the reasons often advanced for asymmetry, such as those given by Craig, are faulty. It is true that the past is not actual, but neither is the future. Likewise, that the past, having occurred, is unalterable is irrelevant, for neither is the future alterable. The only time that is real is the present. Likewise, the argument that if the past were infinite, there would be no reason why we arrive at t0 now rather than earlier, fails.

For Rundle, the past and the future are symmetrical; it is only our knowledge of them that is asymmetrical. Any future event lies at a finite temporal distance from the present. Similarly, any past event lies at a finite temporal distance from the present. For each past or future series of events, beginning from the present, there can always be a subsequent event. Hence, for both series an infinity of events is possible, and, as symmetrical, the infinity of both series is the same. Since the series of future events is not an actual but a potential infinite (or, better, an “indefinitely extendible” series, 2004: 168), the series of past events is also indefinitely extendible. It follows that although the future is actually finite, it does not require an end to the universe, for there is always a possible subsequent event (2004: 180). Similarly, although any given past event of the universe is finitely distant in time from now, a beginning or initial event can be ruled out; for any given event there is a possible earlier event. But then, since there is a possible prior or possible posterior event in any past or future series respectively, the universe, although finite in time, is temporally unbounded (indefinitely extendible); both beginning and cessation are ruled out. (How Rundle [2004: 176–78] gets from the possibility of a subsequent event to actually ruling out cessation and beginning is less than clear.) Since there is no time when the material universe might not have existed, it is not contingent but necessary. Hence, although the principle of sufficient reason is still true, it applies only to the components of the material universe and not to the universe itself. No explanation of the universe is possible. The universe, as matter-energy, is neither caused nor destructible; not in the sense that it could have been caused or could cease, but in the sense that “the notions of beginning and ceasing to exist are inapplicable to the universe” (2004: 178).

But, one might wonder, are the past series and future series of events really symmetrical? It is true that one can start from the present and count either forward and backward in time. Rundle thinks that …\(x5\), \(x4\), \(x3\), \(x2\), \(x1\), \(t0\), \(y1\), \(y2\), \(y3\), \(y4\), \(y5\) are all on the same continuum, so that we cannot distinguish ontologically the time dimension of the future and past series. The two series, going into the past and into the future, would be the same in that however far we count from the present (\(t0\)) remains finite although indefinitely extendible. But is it true that, as he claims, with regard to the past, “any movement currently terminating can be redescribed as extending back”, that counting backward from the present is the same as counting from the past to the present (2004: 176)?

Craig says no, for in the actual world we do not start from now to arrive at the past; we move from the past to the present. To count backwards, we would start from a particular point in time, the present. From where would we start to count were the past indefinitely extendible? Both to count and to move from the past to the present, we cannot start from the indefinitely extendible. Indeed, if the past is indefinitely extendible, no matter where we started, we would have arrived at \(t0\) long before now.

Before the present event could occur the event immediately prior to it would have to occur; and before that event could occur, the event immediately prior to it would have to occur; and so on ad infinitum . One gets driven back and back into the infinite past, making it impossible for any event to occur. Thus, if the series of past events were beginningless, the present event could not have occurred, which is absurd. (Craig and Sinclair 2009: 118)

One cannot just reverse the temporal sequence of the past, for we don’t ontologically engage the sequence from the present to the past. Rundle’s two movements are quite disparate, such that the two sequences—of the past and of the future—are not symmetrical, which leaves intact Craig’s claim that a beginningless past would result in an actual and not a potential infinite.

Craig is well aware of the fact that he is using actual and potential infinite in a way that differs from the traditional usage in Aristotle and Aquinas. For Aristotle all the elements in an actual finite exist simultaneously, whereas a potential infinite is realized over time by addition or division. Hence, the temporal series of events, as formed by successively adding new events, was a potential, not an actual, infinite (Aristotle, Physics , III, 6). For Craig, however, an actual infinite is a timeless totality that cannot be added to or reduced. “Since past events, as determinate parts of reality, are definite and distinct and can be numbered, they can be conceptually collected into a totality” (Craig, in Craig and Smith 1993: 25). The future, but not the past, is a potential infinite, for its events have not yet happened.

Craig’s second argument addresses this very point.

  • The temporal series of events is a collection formed by successive addition.
  • A collection formed by successive synthesis is not an actual infinite.
  • Therefore, the temporal series of events cannot be an actual infinite (Craig 1979: 103).

The collection of historical events is formed by successively adding events, one following another. The events are not temporally simultaneous but occur over a period of time as the series continues to acquire new members. Even if an actual infinite were possible, it could not be realized by successive addition; in adding to the series, no matter how much this is done, even to infinity, the series remains finite and only potentially infinite. One can neither count to nor traverse the infinite (Craig and Sinclair 2009: 118).

It might be objected that this sounds very much like Zeno’s paradoxes that prohibit Achilles or anyone from either beginning to cross an area or succeeding in doing so. But, notes Craig, significant disanalogies disallow this conclusion. For one, Zeno’s argument rests on progressively-narrowing, unequal distances that sum to a finite distance, whereas in traversing the past the equal distances continue to the infinity of the future. Second, Zeno’s distances are potential because of divisibility, whereas the distances from the past are actual distances or times to be traversed.

Morriston (2003: 290) critiques Craig’s thesis that forming the infinite series of events by successive addition presupposes a beginning point. He asks,

Why couldn’t there have been an infinite series of years in which there was no first year? It’s true that in such a series we never “arrive” at infinity, but surely that is only because infinity is, so to speak, “always already there”. At every point in such a series, infinitely many years have already passed by.

Why do we need to “ arrive at infinity?” But Craig’s point is not that we cannot arrive at infinity in the past, but that we could not traverse the infinite to arrive at the present moment. Why this moment rather than another? But maybe, Morriston replies, that is just the way it is; “the past just is the series of events that have already happened”. To require a reason for the series of past events arriving at now is to appeal to the principle of sufficient reason, which he deems both suspect and inappropriate for Craig to invoke (Morriston 2003: 293). Furthermore, he argues, Craig’s argument mistakenly presupposes two independent series—a series of events and a series of segments of time they occupy—such that one can ask about how the former is mapped onto the latter, whereas in fact the two series are not independent. Craig (2010) replies that it is not a matter of sufficient reason, but that Morriston simply has not paid sufficient attention to the distinction between past and present tenses, on which potential and actual infinites are founded. A finite series that has the potential for further members, as with future events beginning with now, is actually finite and only potentially infinite. But a beginningless series of past events cannot add new members; it is actually, not potentially, infinite. There is a relevant distinction between the two series.

To illustrate his conclusion, Craig presents Bertrand Russell’s example of Tristram Shandy, who writes his autobiography. It takes him a year to write about one day of his life, so that as his life progresses so does his autobiography in which he gets progressively farther behind. Russell concludes that

if (Shandy) had lived forever, and had not wearied of his task, then, even if his life had continued as eventfully as it began, no part of his biography would have remained unwritten. ([1903] 1937: 358)

But, Oderberg (2002: 310) claims, Russell seems to have fallaciously moved from (1) For every day, there is a year such that, by the end of that year, Shandy has recorded that day, which is true, to (2) There is a year such that, for every day, by the end of that year Shandy has recorded that day. (2) is needed for Russell’s conclusion but fails to follow from (1). Shandy’s writing never catches up with his life; indeed, the longer he lives, even if for infinity, his writing would never catch up to his life but progressively would get farther behind. Indeed, if he has been living and writing from infinity, his autobiography is infinitely behind his life. Contrary to Russell, there will be days—an infinite number—about which he will be unable to write. As can be imagined, this example has been greatly contested, modified, and has generated a literature of its own. For samples, see Eells (1988), Oderberg (2002), and Oppy 2003.

Finally, it is objected that Craig’s argument presupposes an \(A\) view of time, where time flows from past to present to future and not all events tenselessly coexist. But can Craig’s argument be sustained if time is understood in the \(B\) sense, where all members of the series tenselessly coexist, being equally real (Grünbaum 1994)? On a \(B\) view of time there is no beginning, and it would seem that on this view the argument would collapse. Andrew Loke responds that even on a \(B\) view of time, one cannot realize the infinite by successive addition. It is impossible to count an actual infinite at any time because “an actual infinite has a greater number of elements than what could be counted by the process of counting”. This feature

is due to (i) the nature of the number of elements of an actual infinite set, which is an essential property of such a set, and (ii) the process of counting one element after another. (Loke 2014a: 76)

This response depends crucially on the distinction between an actual and a potential infinite.

Craig and Sinclair’s a posteriori argument for premise 2 invokes recent cosmology and the Big Bang theory of cosmic origins. Since the universe is expanding as the galaxies recede from each other, if we reverse the direction of our view and look back in time, the farther we look, the denser the universe becomes. If we push backwards far enough, we find that the universe reaches a state of compression where the density and gravitational force are infinite. This unique singularity constitutes the beginning of the universe—of matter, energy, space, time, and all physical laws. It is not that the universe arose out of some prior state, for there was no prior state. Since time too comes to be, one cannot ask what happened before the initial event. Neither should one think that the universe expanded from some state of infinite density into space; space too came to be in that event. Since the Big Bang initiates the very laws of physics, one cannot expect any scientific or physical explanation of this singularity.

One picture, then, is of the universe beginning in a singular, non-temporal event roughly 13–14 billion years ago. Something, perhaps a quantum vacuum, came into existence. Its tremendous energy caused it, in the first fractions of a second, to expand and explode, creating the four-dimensional space-time universe that we experience today. How this all happened in the first \(10^{-35}\) seconds and subsequently is a matter of serious speculation and debate; what advocates of premise 2 maintain is that since the universe and all its material elements originate in the Big Bang, the universe is temporally finite and thus had a beginning. (For a detailed consideration of cosmogonic theories from the kalām perspective, see Craig and Sinclair 2009: 125–182; for the counter discussion see Grünbaum 1991). By itself, of course, this reasoning, even if accurate, leaves it the case that premise 2 and hence conclusion 3 are only probably true, dependent on accepted cosmogenic theories.

Sobel (2004: 198) argues that if the universe began at \(t_1\), it is possible that the cause of what came to be itself came to be at \(t_2\), and what caused this came to be at \(t_4\), and so on to infinity, without ever getting to \(t=0\). The beginning converges to \(t=0\), but infinitely never reaches it. But this leads to the paradox that Koons (2014) recently developed, employing reasoning that parallels the Grim Reaper Paradox. He constructs a reductio of the assertion that it is possible that there is an infinite past with infinitely many sub-periods, as well as of the claim that there are an infinite number of periods of time, each of which is earlier than the previous. His constructed dilemma is this:

if time is intrinsically self-measuring, then any extended period of time is divided (in and of itself) into an infinity of actual sub-periods, and so no simple unit of time can be extended. If time is not self-measuring, then a simple period of time (a period in which no process begins or ends) cannot have temporal extension. Either way, an infinitely extended simple past is impossible. (Koons 2014: 261)

One critical response to the kalām argument from the Big Bang is that, given the Grand Theory of Relativity, the Big Bang is not an event at all. An event takes place within a space-time context. But the Big Bang has no space-time context; there is neither time prior to the Big Bang nor a space in which the Big Bang occurs. Hence, the Big Bang cannot be considered as a physical event occurring at a moment of time. As Hawking notes, the finite universe has no space-time boundaries and hence lacks singularity and a beginning (Hawking 1988: 116, 136). Time might be multi-dimensional or imaginary, in which case one asymptotically approaches a beginning singularity but never reaches it. And without a beginning the universe requires no cause. The best one can say is that the universe is finite with respect to the past, not that it was an event with a beginning. (Rundle 2004: chap. 8.)

Grünbaum defends this position by arguing that events can only result from other events.

Since the Big Bang singularity is technically a non-event, and \(t=0\) is not a bona fide time of its occurrence, the singularity cannot be the effect of any cause in the case of either event-causation or agent causation alike…. The singularity \(t=0\) cannot have a cause. (Grünbaum 1994; Rundle 2004: 168, writes, “[T]here is no event—the beginning of the universe—to be explained, events being possible only in time”)

One response to Grünbaum’s objection is to opt for broader notions of “event” and “cause”. We might broaden the notion of “event” by removing the requirement that it must be relational, taking place in a space-time context. In the Big Bang the space-time universe commences and then continues to exist in measurable time subsequent to the initiating singularity (Silk 2001: 456). Thus, one might consider the Big Bang as either the event of the commencing of the universe or else a state in which “any two points in the observable universe were arbitrarily close together” (Silk 2001: 63). As such, one might inquire why there was this initial state of the universe in the finite past. Likewise, one need not require that causation embody the Humean condition of temporal priority, but may treat causation counter-factually, or perhaps even, as traditionally, a relation of production. Any causal statement about the universe would have to be expressed atemporally, but for the theist this presents no problem provided that God is conceived atemporally and sense can be made of atemporal causation.

Furthermore, suppose Grünbaum is correct that the Big Bang singularity is not an event. Then, by his reasoning that events only arise from other events, subsequent so-called events cannot be the effect of that singularity. If they were, they would not be events either. This result that there are no events is absurd.

Given this understanding of space/time, we might reconceive the kalām argument.

  • If something has a finite past, its existence has a cause.
  • The universe has a finite past.
  • Since space-time originated with the universe and therefore similarly has a finite past, the cause of the universe’s existence must transcend space-time (must have existed aspatially and, when there was no universe, atemporally).
  • If the cause of the universe’s existence transcends space-time, no scientific explanation (in terms of physical laws) can provide a causal account of the origin of the universe.
  • If no scientific explanation can provide a causal account of the origin of the universe, the cause must be personal (explanation is given in terms of a personal agent).

Some critics see a problem with this reformulation of the kalām argument in premise 11 . Whereas behind premise 1 of the original argument lies the ancient Parmenidean contention that out of nothing nothing comes, it is alleged that no principle directly connects finitude with causation. They contend that we have no reason to think that just because something is finite it must have a cause of its coming into existence. Theists respond that this objection has merit only if the critic denies that the Principle of Causation is true or that it applies to events like the Big Bang.

Grünbaum (1991) also argues that defenders of the kalām argument cannot make sense of the claim that the universe began to exist.

The question of its beginning is not, “If the universe did have a bounded past of finite duration, what was the cause of its initial event \(t=0\)?” There simply did not exist any instants of time before \(t=0\)!

One simply cannot ask what happened before \(t=0\); the question makes no sense. And if we cannot ask that question, then we cannot inquire whether the Big Bang was an effect, for nothing temporal preceded it. Questions about creation occur in time in the universe, not outside of it (Hawking 1987: 650–51).

Grünbaum’s contention is that to begin to exist requires a previous time, and that there was no time prior to the Big Bang.

[T]here is no first instant of time at all , just as there is no leftmost point on an infinite Euclidean line that extends in both directions. Since here as elsewhere, the term ‘always’ refers to all actual past instants of time, the non-existence of time before \(t=0\) … allows that matter has always existed, despite the finitude of the age of the universe in both sets of models. (1991)

But then, as Craig observes, the series is finite, not infinite, even though it includes all past instants of time. Beginning to exist does not entail that one has a beginning point in time . Craig defines “\(x\) begins to exist” as “\(x\) exists at \(t\) and there is no time immediately prior to \(t\) at which \(x\) exists” (1992: 238).

Something has a beginning just in case that the time during which it has existed is finite.… So understood, deleting the beginning point of a thing’s existence does not imply that the thing no longer begins to exist and therefore came into being uncaused. (Craig and Sinclair 2009: 185).

Morriston (2000) suggests that this analysis of the universe’s coming to be no longer adequately supports premise 1 , for we have no reason to think that something could not just come into existence. Any appeal to ex nihilo nihil fit is either tautologous with the first premise or else appears mistakenly to treat nihilo as if it were “a condition of something”. In part, what Morriston rejects is the intuitiveness that Craig sees in the truth of premise 1 , where “beginning to exist” is understood as explicated as above (see our discussion in 6.1 ). It is not that 1 is false; it is just that it is unsupported and hence loses its plausibility. It has the same plausibility (or implausibility) as creation ex nihilo . Morriston thinks that premise 1 fares equally poorly if Craig attempts to justify it empirically, for we have many situations where the causes of events have not been discovered, and even if we could find the causes in each individual case, it provides no evidence that causation applies to the totality of cases (the universe). (See our discussion of this argument in 4.2 and 4.3 above.) Indeed, he argues, the inductive generalization involved in defending the causal principle stands at odds with similar inductive generalizations that conflict with the kalām argument—that something can be made without there being a prior stuff or that causes can bear no temporal relation to their effects. However, as Craig (2002) points out, although this may weaken the argument were the objection sound, it is not fatal to it unless there is good reason to think that premise 1 actually is false.

Some have suggested that since we cannot “exclude the possibility of a prior phase of existence” (Silk 2001: 63), it is possible that the universe has cycled through oscillations, perhaps infinitely, so that Big Bangs occurred not once but an infinite number of times in the past and will do so in the future. The current universe is a “reboot” of previous universes that have expanded and then contracted (Musser 2004).

The idea of an oscillating universe faces significant problems. For one, no set of physical laws accounts for a series of cyclical universe-collapses and re-explosions. That the universe once exploded into existence provides no evidence that the event could reoccur even once, let alone an infinite number of times, should the universe collapse. Second, even an oscillating universe seems to be finite (Smith, in Craig and Smith 1993: 113). Further, the cycle of collapses and expansions would not, as was pictured, be periodic (of even duration). Rather, entropy would rise from cycle to cycle, so that even were a series of universe-oscillations possible, they would become progressively longer (Davies 1992: 52). If the universe were without beginning, by now that cycle would be infinite in duration, without any hope of contraction. Fourth, although each recollapse would destroy the components of the universe, the radiation would remain, so that each successive cycle would add to the total.

The radiation ends up as blackbody radiation. Because we measure a specific amount of cosmic blackbody radiation in the background radiation, we infer that a closed (oscillating) universe can have undergone only a finite number of repeated bounces

or cycles, no more than 100 and certainly not the infinite number required for a beginningless series.

We reluctantly conclude that a future singularity is inevitable in a closed universe; hypothetical observers cannot pass through it, and so the universe probably cannot be cyclical. (Silk 2001: 380, 399)

The central thesis of the oscillating theory has been countered by recent discoveries that the expansion of the universe is actually speeding up. Observations of distant supernova show that they appear to be fainter than they should be were the universe expanding at a steady rate.

Relative dimness of the supernovae showed that they were 10% to 15% farther out than expected,… indicating that the expansion has accelerated over billions of years. (Glanz 1998: 2157)

The hypothesis that these variations in intensity are caused by light being absorbed when passing through cosmic dust is no longer considered a viable explanation because the most distant supernova yet discovered is brighter than it should be if dust were the responsible factor (Sincell 2001). Some force in the universe not only counteracts gravity but pushes the galaxies in the universe apart ever faster. This increased speed appears to be due to dark energy, a mysterious type of energy, characterized by a negative pressure, composing as much as 70% of the universe. Dark matter, it seems, is overmatched by dark energy. [ 3 ]

Finally, something needs to be said about statement 4 , which asserts that the cause of the universe is personal. Defenders of the cosmological argument suggest two possible kinds of explanation. [ 4 ] Natural explanation is provided in terms of precedent events, causal laws, or necessary conditions that invoke natural existents. Personal explanation is given “in terms of the intentional action of a rational agent” (Swinburne 2004: 21; also Gale and Pruss 1999). We have seen that one cannot provide a natural causal explanation for the initial event, for there are no precedent natural events or natural existents to which the laws of physics apply. The line of scientific explanation runs out at the initial singularity, and perhaps even before we arrive at the initial singularity (at \(10^{-35}\) seconds). If no scientific explanation (in terms of physical laws) can provide a causal account of the origin of the universe, the explanation must be personal, that is, in terms of the intentional action of an intelligent, supernatural agent.

Morriston (2000) questions whether Craig’s argument for the cause being personal goes through. Craig argues that if the cause were an eternal, nonpersonal, mechanically operating set of conditions, then the universe would exist from eternity. Since the universe has not existed from eternity, the cause must be a personal agent who chooses freely to create an effect in time. But, notes Morriston, if the personal cause intended to create the world, and if the intention alone to create is sufficient to bring about the effect, then there is no reason to postulate a personal cause of the universe. Craig (2002) replies that it is not intention alone that must be considered, but the personal agent also employs its personal causal power to bring about the world.

Morriston responds that Craig has equivocated on two notions of eternity: eternity as timelessness and eternity as beginninglessness and endlessness of temporal duration. But the issue seems not to be one of eternity at all, but as in 11–16 above, with finitude. In one sense the universe is eternal: if time came into being with the universe, the universe has existed at every time. But since time came into existence with the universe, both time and the universe are finite in terms of the past and thus need a cause.

Paul Davies argues that one need not appeal to God to account for the Big Bang. Its cause, he suggests, is found within the cosmic system itself. Originally a vacuum lacking space-time dimensions, the universe “found itself in an excited vacuum state”, a “ferment of quantum activity, teeming with virtual particles and full of complex interactions” (Davies 1984: 191–2), which, subject to a cosmic repulsive force, resulted in an immense increase in energy. Subsequent explosions from this collapsing vacuum released the energy in this vacuum, reinvigorating the cosmic inflation and setting the scenario for the subsequent expansion of the universe. But what is the origin of this increase in energy that eventually made the Big Bang possible? Davies’s response is that the law of conservation of energy (that the total quantity of energy in the universe remains fixed despite transfer from one form to another), which now applies to our universe, did not apply to the initial expansion. Cosmic repulsion in the vacuum caused the energy to increase from zero to an enormous amount. This great explosion released energy, from which all matter emerged. Consequently, he contends, since the conclusion of the kalām argument is false, one of the premises of the argument—in all likelihood the first—is false.

Craig responds that if the vacuum has energy, the question arises concerning the origin of the vacuum and its energy. Merely pushing the question of the beginning of the universe back to some primordial quantum vacuum does not escape the question of what brought this vacuum laden with energy into existence. A quantum vacuum is not nothing (as in Newtonian physics) but

a sea of continually forming and dissolving particles that borrow energy from the vacuum for their brief existence. A vacuum is thus far from nothing, and vacuum fluctuations do not constitute an exception to the principle

enunciated in premise 1 (Craig, in Craig and Smith 1993: 143–4). Hence, he concludes, the appeal to a vacuum as the initial state is misleading.

One might wonder, as Rundle (2004: 75–77) does, how a supernatural agent could bring about the universe. He contends that a personal agent (God) cannot be the cause because intentional agency needs a body and actions occur within space-time. But acceptance of the cosmological argument does not depend on an explanation of the manner of causation by a necessary being. When we explain that the girl raised her hand because she wanted to ask a question, we can accept that she was the cause of the raised hand without understanding how her wanting to ask a question brought about her raising her hand. As Swinburne notes, an event is “fully explained when we have cited the agent, his intention that the event occur, and his basic powers” that include the ability to bring about events of that sort (2004: 36). Similarly, theists argue, we may never know why and how creation took place. Nevertheless, we may accept it as an explanation in the sense that we can say that God created that initial event, that he had the intention to do so, and that such an event lies within the power of an omniscient and omnipotent being; not having a body is irrelevant.

The issues raised by the kalām argument concern not only the nature of explanation and when an explanation is necessary, but even whether an explanation of the universe is possible (given the above discussion). Whereas all agree that it makes no sense to ask about what occurs before the Big Bang (since there was no prior time) or about something coming out of nothing, the dispute rests on whether there needs to be a cause of the first natural existent, whether something like the universe can be finite and yet not have a beginning, and the nature of infinities and their connection with reality.

Richard Swinburne contends that the cosmological argument is not deductively valid; if it were,

it would be incoherent to assert that a complex physical universe exists and that God does not exist. There would be a hidden contradiction buried in such co-assertions…. [A]ttempts to derive obviously incoherent propositions from such co-assertions have failed through commission of some elementary logical error. (2004: 136)

Swinburne is correct that if someone believes that a deductive cosmological argument (proof) for God’s existence is sound, then it would be incoherent for that same person to then deny that God exists. But in their respective proofs defenders of the deductive cosmological arguments make a claim about incoherence, namely, that it would be contradictory for the same person to affirm the premises of the argument and to claim that God or a personal necessary being does not exist. And for them, the respective premises not only have the intuitiveness or argumentative support that Swinburne deems necessary, but believe that the argument has not committed some “elementary error of logic”.

Has Swinburne shown incoherence? Whereas propositions are true and deductive arguments are valid independent of anyone’s beliefs that they are true and valid (the proposition that the earth orbits the sun is true regardless of whether anyone believes it), the acceptance of premises as true, of deductive arguments as valid, and of an argument’s use as a proof is not independent of those same beliefs. An argument that one person takes as being sound another might believe not to be sound, in that the person rejects one of the premises or holds that the conclusion fails to properly follow; arguments are person-relative in their persuasive value or assessment of coherence. Swinburne himself notes that arguments of coherence and incoherence are persuasive only to the extent that someone accepts other statements inherent to the proof as coherent or incoherent and that one statement entails another (1993: 39). Elsewhere Swinburne admits to having

some doubt about whether men have enough initial consensus about what is coherent and what entails what, are clever enough and have enough imagination to reach agreed proofs which would settle all disputes about whether a statement is coherent or incoherent. (1993: 45)

As such, Swinburne cannot so easily dismiss deductive cosmological arguments, although he is justified in wondering whether “reasons less strong than compelling proofs can be given for thinking some statements coherent and others incoherent” (1993: 45).

In place of a deductive argument, Swinburne develops an inductive cosmological argument that appeals to the inference to the best explanation. Swinburne distinguishes between two varieties of inductive arguments: those that show that the conclusion is more probable than not (what he terms a correct P-inductive argument) and those that further increase the probability of the conclusion (what he terms a correct C-inductive argument). In The Existence of God (1979) he presents a cosmological argument that he claims falls in the category of C-inductive arguments. However, for him this argument is part of a larger, cumulative case for a P-inductive argument for God’s existence that includes as its evidence the orderliness of the universe, the existence of consciousness, miracle reports, and religious experience.

Swinburne notes that “a cosmological argument argues that the fact that there is a universe needs explaining” (2004: 9–10). But he emphasizes that his approach differs from those we have already considered in that he rejects the Principle of Sufficient Reason understood as “everything not ‘metaphysically necessary’ has an explanation in something ‘metaphysically necessary’” (2004: 148), for the PSR leads, as it does in Leibniz, to a being that is logically necessary, and such a being cannot explain the logically contingent. From the logically necessary only the logically necessary follows. In place of using the PSR to construct a deductive argument, he employs a “basic theorem of confirmation theory”, Bayes Theorem, to construct an inductive argument (2004: 67). (In making this claim about the need for an explanation of the universe, however, it is hard not to see that he invokes some formulation of the PSR.)

Swinburne begins his discussion with the existence of a physical universe that contains odd events that cannot be fitted into the established pattern of scientific explanation (e.g., miracles, the appearance of conscious beings), is too big in that science cannot explain why there are states of affairs at all or why the fundamental natural laws to which science appeals to explain things hold), and is complex (its matter-energy has relevant powers). (2004: 74, 150)

It is not logically necessary that the existence of the universe needs explanation; we could accept this universe as a brute, inexplicable fact, but Swinburne thinks that to do so fails to accord with the example of the sciences, which seek the best explanation for any given phenomena. Since “it is reasonable to suppose” that there is an explanation (2004: 75), the issue, then, is which view is more reasonable: that science can provide a natural explanation for the existence of this universe, or whether the universe and its phenomena exist because of the intentional causal activity of a personal being that also is a brute fact.

To find the explanatory hypothesis most likely to be true, especially about something that might be unobservable, he claims to follow the example of science. Using Bayes Theorem, he looks for a hypothesis \(h\) such that \(p(e\mid h \mathbin{\&} k) \gt p(e\mid k)\) where \(p\) is probability, \(e\) is the existence of a complex universe, and \(k\) is the background data. A hypothesis is more likely to be true (1) in so far as it has high explanatory power, in that it makes probable the evidence of the observation; this may be predictive but can be postdictive as well (Swinburne 1996: 34, 2001: 80–81), and insofar as the evidence is very un likely to occur if the hypothesis is false. And (2), it has a greater prior probability. The prior probability of a hypothesis encompasses three features: (a) how well it fits with our background knowledge (2001: 81). The broader the scope, the less relevant this criterion becomes (2004: 60). Since there are no “neighboring fields of inquiry related to the origin of the universe”, Swinburne treats this condition in the cosmological argument as irrelevant or reducing to the feature of simplicity (1996: 29). (b) The scope of the hypothesis (the extent of its claims)—the broader the scope, the less likely it is to be true. For example, all crows are black is less likely to be true than all crows along the upper Mississippi River are black. Since both scientific naturalism and theism have the same scope—explaining the universe, this does not factor into his calculations for explaining the complex universe (2001: 82); and (c) simplicity, which for Swinburne holds the key (2001: 82–83).

He holds that we are looking for a complete explanation, where

we may reasonably conclude that the criteria for supposing that factors have no further explanation (scientific or personal) in terms of factors acting at the time and so that any explanation is a complete explanation overall (not just a complete explanation within scientific or within personal explanation) are that any attempt to go beyond the factors that we have would result in no gain of explanatory power or prior probability. (2004: 89)

A scientific explanation fails to give a complete explanation. It leaves us not with a simple but with a very complex explanatory hypothesis, in that “ultimate explanation stops at innumerable, different stopping points, many of them … having exactly the same powers and liabilities as each other” (1996: 42). It presents us with the brute fact of the existence of the universe, not an explanation for it. It explains in terms of a full cause the events at any moment, but it cannot provide a complete explanation of the universe, “for there are no physical causes apart from the universe itself and parts thereof” (1984: 144).

On the other hand, a personal explanation, given in terms of the intentional actions of a person, is simpler and no explanatory power is lost. Further, a personal explanation can be understood, as in the case of explaining basic actions, without knowing or understanding any of the natural causal conditions that enable one to bring it about. In the case of the cosmological argument, personal explanation is couched in terms of a being that has beliefs, purposes, and intentions, and possesses both the power to bring about the complex universe and a possible reason for doing so.

Swinburne argues that a personal explanation of the universe satisfies the above probability criteria. It satisfies condition (1) in that appealing to God as an intentional agent has explanatory power. It leads us to have certain expectations about the universe: that it manifests order, is comprehensible, and favors the existence of beings that can comprehend it. It makes probable the existence of the complex universe because God could have reasons for causing such a universe, whereas we would have no reasons at all if all we had was the brute fact of the material universe. Among these reasons is that the universe would be “a theatre for finite agents to develop and make of it what they will” (Swinburne 1979: 131).

Michael Martin objects at this point. Martin contends that if Swinburne is to compare the a priori probability of there being a complex universe given our background knowledge with the a priori probability of a complex universe given our background knowledge and the existence of God, he has to be clear on how he interprets the probability. Martin notes that herein lies crucial ambiguity that disables calculating the a priori probability. If one compares the very many possible complex universes with there being no universe, on the basis of assigning equal probability to all possibilities the probability of there being a complex universe is nearly 1. But if one compares the probability of there being a complex universe with there being no universe at all, it is 50 percent (Martin 1986: 155). Furthermore, Martin wonders whether complexity is an issue at all. According to Swinburne, as free God can create any kind of world or no world at all. But then the existence of God is compatible with any number of scenarios: the existence of no world, a simpler world than we have, one like ours, or any number of more complex universes. Consequently, the complexity of this world does not matter in constructing an inductive argument for God’s existence (1986: 155). Put another way, adding the existence of God to our background knowledge does not increase the likelihood of there being a complex universe, let alone of there being this particular universe or a universe at all (1986: 158).

In short, Martin does not see how Swinburne can establish an a priori probability for the existence of a complex universe, to be compared with an a priori probability for the existence of God based on simplicity, a feature of Swinburne’s Bayesian argument. This introduces the theme of simplicity, to which Swinburne devotes much attention.

Swinburne goes on to argue that a personal explanation in terms of God satisfies condition (2) because of its simplicity. If one is going to construct an explanatory hypothesis using the criterion of simplicity, God rather than science is more likely to be the focus of the true explanatory hypothesis. God is one and of one kind; polytheism is ruled out. Moreover, God is the simplest kind of person there can be because a person is a being with power (to do intentional actions), knowledge, and freedom (to choose, uncaused, which actions to do), and in God these properties are infinite, and having infinite properties is simpler than having properties with limits, as humans do. “It is always simpler to postulate infinite or zero degrees of some property than a certain precise finite value of it” (Swinburne 1983: 385). Furthermore, God engages in simple causation, that is, causation by simple intention. Swinburne concludes that although the prior likelihood of neither God nor the universe is particularly high, the prior probability of a simple God exceeds that of a complex universe. Hence, if anything is to occur unexplained, it would be God, not the universe.

Consequently, if we are to explain the universe, we must appeal to a personal explanation

in terms of a person who is not part of the universe acting from without. This can be done if we suppose that such a person (God) brings it about at each instant of time, that (the laws of nature) \(L\) operate. (Swinburne 1979: 126, 2004: 142)

Although for Swinburne this argument does not make the existence of God more probable than not (it is not a P-inductive argument), it does increase the probability of God’s existence (is a C-inductive argument) because it provides a more reasonable explanation for the universe than merely attributing it to the brute fact of the universe's existence.

Theism does not make [certain phenomena] very probable; but nothing else makes their occurrence in the least probable, and they cry out for an explanation. A priori , theism is perhaps very unlikely, but it is far more likely than any rival supposition. Hence, our phenomena are substantial evidence for the truth of theism. (Swinburne 1979: 290)

In his critique of Swinburne, J. L. Mackie wonders whether personal explanations are reducible to natural, scientific explanations. To implement intentionality requires an entire system of neurological and macro-biological conditions. Not only does God as nonphysical lack these biological conditions, but these conditions are exceedingly complex, not simple. “Only by ignoring such key features [the role of the body] do we get an analogue of supposed divine action” (Mackie 1982: 100). When we incorporate these features, the simplicity disappears.

Swinburne replies that Mackie has misunderstood his argument. “The simplicity of the relation between intention and its realization has nothing to do with how our will or intentions are realized in practice” (Swinburne 1983: 386). Even if we understand all the neural connections and firings, we may not achieve any better explanation of why persons intended to act as they did than simply asking them why they acted as they did. This indicates that the existence of intermediate physical causal links is not an essential part of personal explanation. In fact, Swinburne argues, since it is easier to understand the function of intention without invoking any physical causal limitations, it makes it easier to understand the case of God who as nonphysical has no need for intermediary physical processes. Thus, he claims, Mackie missed the point about God when he invokes the complexity of physical accounts. The point is that God can will to act on his intentions directly, and this provides a simple account or explanation of why things came to exist.

The critical aspect of Swinburne’s argument is his almost total reliance in his inductive cosmological argument on simplicity as the deciding factor between competing hypotheses regarding the cause of the existence of the universe (2004: 333–34; 2010, 9; Ostrawick). Swinburne has at least six understandings that one hypothesis is simpler than another. (1) It invokes the fewest number of entities (2004: 106, 1983: 386, 2001: 87; 2010, 5). This is a quantitative understanding. (2) It invokes the fewest kinds of entities—a qualitative understanding (1983: 386, 2001: 87). (3) It invokes entities with simple or few properties (1983: 386) Swinburne invokes a subcriterion that explanations are simpler when the properties they invoke are observable (2010, 6). (4) It invokes powers, acquisition of beliefs, and consistency of intentions similar to ours when applied to personal explanation of rational behavior (2004: 61–64). (5) The explanation invokes the simplest organization of the features functioning in the explicans, e.g., laws or variables (2001: 83, 89–90). (6) Simplicity can be found in the explicans in that it does not invoke extraneous features that are not necessary to explain the effect (2001: 81).

Swinburne holds that the appeal to God as an explanation is simpler in all of these ways. [ 5 ] Not only is there one entity and that entity is simple, the explanation effectively has no organization of the features. The explanation itself is simple. The appeal to God’s causal activity satisfies interpretation 6 in that it involves no extraneous entities to do the explaining and requires no intermediaries. God can bring about the effect by himself alone.

Several important questions about simplicity arise. First, is simplicity the criterion we should use to decide between hypotheses? For one thing, simplicity is not always a reliable criterion for determining which hypothesis is true or which hypothesis provides the best explanatory account. The rise of quantum explanations suggests that the simplest account of the universe, for example, that of Newton, is not a complete and fully adequate account. The events in the subatomic realm are far from simply explained. For another, although an explanation in terms of four factors might make an explanation simpler, the reverse might hold: an explanation in terms of ten factors might be simpler than an explanation in terms of four because the relationships that hold between the ten facts are less complex than those that hold between the four, making for a simpler explanation (Ostrowick 2012). In reply, Swinburne might grant this, but argue that in these much more limited cases explanatory power, background knowledge, and scope now come into significant play in a way that they don’t when addressing hypotheses explaining the oddness, bigness, and complexity of the universe.

Second, why think that theism is simpler than naturalism? Oppy argues that whereas both naturalism and theism equally fit the data and have the same scope, naturalism is simpler, for theism is

committed to two kinds of entities (the natural and the supernatural), two kinds of external relations (the natural and the supernatural), two kinds of causation (the natural and the supernatural), two kinds of non-topic-neutral properties (the natural and the supernatural), and so on, whereas naturalism is committed to only one kind in each of these categories. (2013: 52)

Swinburne’s reply is that naturalism might provide a full explanation for individual things, but it cannot do so for the universe itself, for the reasons noted above.

In conclusion, Swinburne contends that it is very unlikely that a universe would exist uncaused, but more likely that God would exist uncaused. It is likely that if there is a God, he will make something like the finite and complex universe. The puzzling existence of the universe can be made comprehensible (explicable) if we suppose it is brought about by a personal God with intentional beliefs and the power to bring intentions to fruition (2004: 152). Whether simplicity can bear the weight of his argument is the key matter in question.

Finally, even if the cosmological argument is sound or cogent, the difficult task remains to show, as part of natural theology, that the necessary being to which the cosmological argument concludes is the God of religion, and if so, of which religion. Rowe suggests that the cosmological argument has two parts, one to establish the existence of a first cause or necessary being, the other that this necessary being is God (1975: 6). It is unclear, however, whether the second contention is an essential part of the cosmological argument. Although Aquinas was quick to make the identification between God and the first mover or first cause, such identification seems to go beyond the causal reasoning that informs the argument (although one can argue that it is consistent with the larger picture of God and his properties that Aquinas paints in his Summae ). Some (Rasmussen, O’Connor, Koons) have plowed ahead in developing this stage 2 process by showing how and what properties—simplicity, unity, omnipotence, omniscience, goodness, and so on—might follow from the concept of a necessary being. It “has implications that bring it into the neighborhood of God as traditionally conceived” (O’Connor 2008: 67). Others have proposed a method of correlation, where to give any religious substance to the concept of a necessary being, one conducts a lengthy discussion of the supreme beings found in the diverse religions and carefully correlates the properties of a necessary being with those of a religious being. This is done to discern compatibilities and incompatibilities (Attfield 1975).

Regardless of the connection of a necessary being with religion, it is necessary to flesh out the nature of the necessary being if one is to hold that the cosmological argument is informative. As O’Connor notes, the mere concept of a necessary being is “quite thin”. Along with classical Islamic defenders of the argument (e.g., al-Baghdadi (c. 1000); O’Connor (2008: 88) concludes that there is a necessary connection between a personal necessary being and its nature. He suggests that there is not a contingent but a “subtle entailment relation” between certain essential properties like being perfectly powerful, perfectly free, and perfectly knowledgeable. For example, the

extent of power seems to be a function of at least two variables: the amount of work that can be performed in a single task and the range of tasks one is able to perform in a given circumstance.… Perfect power and freedom would require an essentially unlimited knowledge, corresponding to the unlimited range of possibilities. (2008: 89)

A necessary being must also be causally independent for its existence and thus transcendent (2008: 92). Similarly, Swinburne ties God’s perfection to his simplicity that, as we have seen, functions centrally in his argument.

Two notions of necessity are found in the conclusion to the deductive argument: “Necessarily, a necessary being exists”. The first is conditional necessity: the proposition is necessary given that the premises are true and the argument valid. The other use concerns what is meant by “necessary being”. O’Connor writes that God is absolutely necessary, by which he means that God is “absolutely invulnerable to nonexistence” (2008: 70).

The concept of a necessary being is of one that could not have failed to exist, absolutely speaking. For such a being to be possible, it must be such that it would exist in every possible circumstance, including the actual one. That’s precisely why the question of its existence cannot arise, thereby ending the regress of explanation nonarbitrarily. (O’Connor 2013: 42)

For him necessary existence is necessarily tied up with a particular nature (otherwise the existence would be contingent) but not derivative from it; God’s existence entails his nature (2008: 88). God’s necessity is not logical (for there is no contradiction in denying that such a being exists) but made possible on explanatory grounds (the cosmological argument). But, we might inquire, if God could not have failed to exist, how does an absolutely necessary being differ from a logically necessary being? O’Connor goes on to argue that God’s absolute necessity does not invoke the ontological argument. He agrees that by S5, if it is possible that a necessary being exists, it necessarily exists (2008: 71), but denies that this invokes the ontological argument, since it “gives no reason to think that the nature in question is genuinely possible, and not merely logically consistent”. But, one might wonder, what would one have to establish to show that it is genuinely possible that a necessary being exists? (see Plantinga, God, Freedom, and Evil , 1967: 112). Gale himself admits that, given this view of necessity and S5, the ontological argument works although we don’t know how to properly construct it (Gale and Pruss 1999: 462).

One way to understand the necessary being is as factually or metaphysically necessary. In this understanding, the necessary being is “sheer, ultimate, unconditioned reality, without beginning or end” (Hick 1960: 730). God’s necessity refers to his aseity, in that God does not depend on anything else for his existence. It is from God’s aseity that his eternity follows.

God is not one fact amongst others, but is related asymmetrically to all other facts as that which determines them. This is the ultimate given circumstance, which it is not possible to go with either question or explanation. For to explain something means either to assign a cause to it or to show its place within some wider context in relation to which it is no longer puzzling to us. But the idea of the self-existent Creator of everything other than himself is the idea of a reality which is beyond the scope of these explanatory procedures. (Hick 1960: 733–34)

Given this reading of “necessary being”, God as the necessary being possesses metaphysical necessity and logical contingency (Hick 1960; Swinburne 2004: 79). If the necessary being exists at any time, then necessarily it exists at all times. If the necessary being does not exist, it cannot come into existence. Nothing can bring it into existence or cause it to cease to exist. Thus, if God exists now, it is not coherent to suppose that any agent can make it false that God exists (Swinburne 2004: 249, 266). O’Connor objects that if the necessary being is contingent, it just happens to exist (2008: 70; see White for further objections). But one might reply that God does not just happen to exist; God exists because of his nature (although his nature does not precede his existence).

Much more can be said about necessity and the other properties associated with a necessary being. While defenders of the cosmological argument point to the relevance and importance of connecting the necessary being with natural theology, critics find themselves freed from such endeavors.

  • al–Baghdadi, Abu Mansur, [c.1000] 1981, ‘Usul al–din , 3rd edition Beirut: Dar al-kutub al–‘ilmiyya.
  • Aquinas, Thomas, Summa Contra Gentiles , Bk. 1. Notre Dame: University of Notre Dame Press, 1975. [ Summa Contra Gentiles available online ]
  • –––, [ST] Summa Theologica , I, q. 2. [ Summa Theologica available online ]
  • Attfield, Robin, 1975, “The God of Religion and the God of Philosophy”, Religious Studies , 9(1): 1–9. doi:10.1017/S0034412500006259
  • Beck, W. David, 2002, “The Cosmological Argument: A Current Bibliographical Appraisal”, Philosophia Christi , 2(2): 283–304. [ Beck 2002 available online ]
  • Bonaventure, c. 1250, Commentary on the Sentences: Philosophy of God. Works of St. Bonaventure XVI, R.E. Houser and Timothy B. Noone (eds), St. Bonaventure, NY: Franciscan Institute Press, 2014.
  • Brown, Patterson, 1966, “Infinite Causal Regression”, Philosophical Review , 75(4): 510–25. doi:10.2307/2183226
  • Copan, Paul and William Lane Craig, 2004, Creation out of Nothing , Grand Rapids: Baker.
  • Craig, William Lane, 1979, The Kalām Cosmological Argument , London: Macmillan Press.
  • –––, 1980, The Cosmological Argument from Plato to Leibniz , London: Macmillan Press.
  • –––, 1992, “The Origin and Creation of the Universe: A Reply to Adolf Grünbaum”, British Journal for the Philosophy of Science , 43(2): 233–40. doi:10.1093/bjps/43.2.233.
  • –––, 1997, “In Defense of the Kalām Cosmological Argument”, Faith and Philosophy , 14(2): 236–47. doi:10.5840/faithphil19971422 [ Craig 1997 available online ]
  • –––, 2002, “Must the Beginning of the Universe Have a Personal Cause?: A Rejoinder”, Faith and Philosophy , 19(1): 94–105. doi:10.5840/faithphil20021917
  • –––, 2010, “Taking Tense Time Seriously in Differentiating Past and Future: A Response to Wes Morriston”, Faith and Philosophy , 27(4): 451–56. doi:10.5840/faithphil201027445
  • Craig, William Lane and James P. Moreland, (eds.), 2009, The Blackwell Companion to Natural Theology , London: Blackwell,
  • Craig, William Lane and James D. Sinclair, 2009, “The Kalām Cosmological Argument”, in Craig and Moreland 2009: 101–201. [This contains an exhaustive bibliography on the Kalām cosmological argument.] [ Craig and Sinclair 2009 preprint available online ]
  • Craig, William Lane and Quentin Smith, 1993, Theism, Atheism, and Big Bang Cosmology , New York: Oxford University Press. doi:10.1093/acprof:oso/9780198263838.001.0001
  • Davey, Kevin and Rob Clifton, 2001, “Insufficient Reason in the ‘New Cosmological Argument’”, Religious Studies , 37(4): 485–490. doi:10.1017/S0034412501005819
  • Davidson, Herbert A., 1969, “John Philoponus as a Source of Medieval Islamic and Jewish Proofs of Creation”, Journal of the American Oriental Society , 89(2, Apr.–June): 357–91. doi:10.2307/596519
  • Davies, Paul, 1984, Superforce , New York: Simon and Schuster.
  • –––, 1992, The Mind of God , New York: Simon and Schuster.
  • Davis, Stephen, 1997, God, Reason & Theistic Proofs , Grand Rapids: Eerdmans.
  • Dvali, George, 2004, “Out of the Darkness”, Scientific American , Feb. 290(2): 68–75.
  • Eells, Ellery, 1988, “Quentin Smith on Infinity and the Past”, Philosophy of Science , 55(3): 453–455. doi:10.1086/289451 [ Eells 1988 preprint available online ]
  • Fakry, Majid, 1957, “The Classical Islamic Arguments for the Existence of God”, The Muslim World , 47(2): 133–45. doi:10.1111/j.1478-1913.1957.tb02956.x [ Fakry 1957 available online ]
  • Flew, Anthony and Alasdair C. MacIntyre (eds.), 1955, New Essays in Philosophical Theology , London: SCM Press.
  • Gale, Richard M. (ed.), 1991, On the Nature and Existence of God , Cambridge: Cambridge University Press. 10.1017/CBO9781316499054
  • Gale, Richard M. and Alexander R. Pruss, 1999, “A New Cosmological Argument”, Religious Studies , 35(4): 461–476. Reprinted in Gale and Pruss 2003: 365–80. [ Gale and Pruss 1999 available online ]
  • –––, 2002, “A Response to Oppy, and to Davey and Clifton”, Religious Studies , 38(1): 89–99. doi:10.1017/S0034412501005923
  • ––– (eds.), 2003, The Existence of God , Burlington, VT: Ashgate.
  • Gellman, Jerome, 2000, “Prospects for a sound stage 3 of cosmological arguments”, Religious Studies , 36(2): 195–2001. Reprinted in Gale and Pruss 2003: 381–87.
  • Glanz, James, 1998, “Cosmic Motion Revealed”, Science , 282(5397): 2156–2157. doi:10.1126/science.282.5397.2156a
  • Goldschmidt, Tyron (ed.), 2013, The Puzzle of Existence: Why Is There Something Rather than Nothing? New York: Routledge.
  • Grünbaum, Adolf, 1991, “Creation as a Pseudo-Explanation in Current Physical Cosmology”, Erkenntnis , 35(1–3): 233–54. doi:10.1007/BF00388287 [ Grünbaum 1991 available online ]
  • –––, 1994, “Some Comments on William Craig’s ‘Creation and Big Bang Cosmology’”, Philosophia Naturalis , 31(2): 225–236. [ Grünbaum 1994 available online ]
  • Hawking, Stephen W., 1987, “Quantum Cosmology”, in Stephen W. Hawking and Werner Israel (ed.), Three Hundred Years of Gravitation , Cambridge: Cambridge University Press, 631–51.
  • –––, 1988, A Brief History of Time , New York: Bantam Books.
  • Heil, John, 2013, “Contingency”, in Goldschmidt 2013: 167–81.
  • Hick, John, 1960, “God as Necessary Being”, Journal of Philosophy , 57(22/23): 725–33.
  • Hume, David, 1779, Dialogues Concerning Natural Religion , London. Reprinted Indianapolis: Hackett, 1980. [ Hume 1779 available online ]
  • –––, 1748, An Enquiry Concerning Human Understanding , London. Reprinted Indianapolis: Hackett, 1993. [ Hume 1748 available online ]
  • Kant, Immanuel, 1781/1787, Critique of Pure Reason , Cambridge: Cambridge University Press, 1998. [ Kant Critique available online ]
  • Kenny, Anthony, 1969, The Five Ways , New York: Schocken Books.
  • Koons, Robert C., 1997, “A New Look at the Cosmological Argument”, American Philosophical Quarterly , 34(2): 193–211.
  • –––, 2008, “Epistemological Foundations for the Cosmological Argument”, in Jonathan L. Kvanvig, (ed.), Oxford Studies in the Philosophy of Religion , Oxford: Oxford University Press.
  • –––, 2014, “A New Kalam Argument: Revenge of the Grim Reaper”, Noûs , 48(2): 256–67. doi:10.1111/j.1468-0068.2012.00858.x
  • Leftow, Brian, 2012, God and Necessity , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199263356.001.0001
  • Leibniz, Gottfried, 1714, The Monadology , Pittsburgh, PA: University of Pittsburgh Press, 1991. [ Leibniz 1714 available online ]
  • Leslie, John, 1979, Value and Existence , Totowa, NJ: Rowman and Littlefield.
  • Loke, Andrew, 2014a, “A Modified Philosophical Argument for the Beginning of the Universe”, Think , 13(36): 71–83. doi:10.1017/S147717561300033X
  • –––, 2014b, “No Heartbreak at Hilbert’s Hotel: a Reply to Landon Hedrick”, Religious Studies , 50(1): 47–50. doi:10.1017/S0034412513000346
  • Mackie, John L., 1982, The Miracle of Theism: Arguments for and against the Existence of God , Oxford: Clarendon Press.
  • Martin, Michael, 1990, Atheism: A Philosophical Justification , Philadelphia: Temple University Press.
  • –––, 1991, The Case Against Christianity , Philadelphia: Temple University Press.
  • –––, 1986, “Swinburne’s Inductive Cosmological Argument”, Heythrop Journal , 27(2): 151–162. doi:10.1111/j.1468-2265.1986.tb00085.x
  • Mermin, N. David, 1985, “Is the moon there when nobody looks? Reality and the quantum theory”, Physics Today , 38(4): 38–47. doi:10.1063/1.880968
  • Miethe, Terry L., 1978, “The Cosmological Argument: A Research Bibliography”, New Scholasticism , 52(2): 285–305. doi:10.5840/newscholas197852210
  • Morriston, Wes, 2000, “Must the Beginning of the Universe Have a Personal Cause? A Critical Examination of the Kalam Cosmological Argument”, Faith and Philosophy , 17(2): 149–69. doi:10.5840/faithphil200017215
  • –––, 2002, “Causes and Beginnings in the Kalam Argument: Reply to Craig ”, Faith and Philosophy , 19(2): 233–44. doi:10.5840/faithphil200219218
  • –––, 2003, “Must Metaphysical Time Have a Beginning”, Faith and Philosophy , 20(3): 288–306. doi:10.5840/faithphil200320338
  • –––, 2010, “Beginningless Past, Endless Future, and the Actual Infinite”, Faith and Philosophy , 27(4): 439–50. doi:10.5840/faithphil201027444
  • Musser, George, 2004, “Four Keys to Cosmology”, Scientific American 290(2, February): 42–43.
  • Nowacki, Mark R., 2007, The “Kalām” Cosmological Argument for God , New York: Barnes and Noble.
  • O’Connor, Timothy, 2004, “‘And This All Men Call God’”, Faith and Philosophy , 21(4): 417–35. doi:10.5840/faithphil200421436 [ O’Connor 2004 available online ]
  • –––, 2008, Theism and Ultimate Explanation: the Necessary Shape of Contingency , London: Wiley-Blackwell.
  • –––, 2013, “Could There Be a Complete Explanation of Everything ?” in Goldschmidt 2013: 22–45.
  • Oderberg, David S., 2002, “Traversal of the Infinite: the ‘Big Bang,’ and the Kalam Cosmological Argument”, Philosophia Christi , 4(2): 305–344. [ Oderberg 2002 available online ]
  • –––, 2013, “The Cosmological Argument”, in Chad Meister and Paul Copan (eds.), The Routledge Companion to Philosophy of Religion , London: Routledge, 401–10.
  • Oppy, Graham, 1999, “Koons’ Cosmological Argument”, Faith and Philosophy , 16(3): 379–389. doi:10.5840/faithphil199916335
  • –––, 2000, “On ‘A New Cosmological Argument’”, Religious Studies , 36(3): 345–353.
  • –––, 2002, “Arguing about the Kalam Cosmological Argument”, Philo , 5(1): 34–61. doi:10.5840/philo2002513 [This article contains helpful references tracing the history of the debate between Craig and his critics like Oppy, Grünbaum, Mackie, and Smith.]
  • –––, 2003, “From the Tristram Shandy Paradox to the Christmas Shandy Paradox: A Reply to Oderberg”, Ars Disputandi , 3: 1–24. [ Oppy 2003 available online ]
  • –––, 2006, Arguing about Gods . Cambridge: Cambridge University Press.
  • –––, 2009, “Cosmological Arguments”, Noûs , 43(1): 31–48. doi:10.1111/j.1468-0068.2008.01694.x
  • –––, 2013, “Ultimate Naturalistic Causal Explanation”, in Goldschmidt 2013: 46–63.
  • Ostrowick, John, 2012, “Is Theism a Simple, and Hence, Probable, Explanation for the Universe?” South African Journal of Philosophy , 31(2): 354–68. doi:10.1080/02580136.2012.10751781
  • Plantinga, Alvin, 1967, God and Other Minds , Ithaca: Cornell University Press.
  • Potter, Karl H. (ed.), 1977, Indian Metaphysics and Epistemology: The Tradition of Nyāya-Vaiśeşika up to Gangeśa , Princeton: Princeton University Press.
  • Pruss, Alexander R., 1999, “The Hume-Edwards Principle and the Cosmological Argument”, in Gale and Pruss 2003: 347–63.
  • –––, 2006, The Principle of Sufficient Reason: A Reassessment , Cambridge: Cambridge University Press.
  • –––, 2009, “The Leibnizian Cosmological Argument”, in Craig and Moreland 2009: 24–100. [ Pruss 2009 available online ]
  • Quinn, Philip, 2005, “Cosmological Contingency and Theistic Explanation”, Faith and Philosophy , 22(5): 581–600. doi:10.5840/faithphil200522520
  • Rasmussen, Joshua, 2009, “From a Necessary Being to God”, International Journal for Philosophy of Religion , 66(1): 1–13. doi:10.1007/s11153-008-9191-8
  • –––, 2010, “A New Argument for a Necessary Being”, Australasian Journal of Philosophy , 89(2): 351–56. doi:10.1080/00048402.2010.523706
  • Reichenbach, Bruce R., 1972, The Cosmological Argument: A Reassessment , Springfield: Charles Thomas.
  • –––, 1982, Evil and a Good God , New York: Fordham University Press.
  • –––, 2004, “Explanation and the Cosmological Argument”, in Michael L. Peterson and Raymond J. VanArragon (eds.), Contemporary Debates in Philosophy of Religion , London: Blackwell, 97–114.
  • Rowe, William L., 1962, “The Fallacy of Composition”, Mind , 71(281): 87–92. doi:10.1093/mind/LXXI.281.87
  • –––, 1968, “The Cosmological Argument and the Principle of Sufficient Reason”, Man and World , 1(2): 278–92. doi:10.1007/BF01258405
  • –––, 1975, The Cosmological Argument , Princeton: Princeton University Press.
  • –––, 1997, “Circular Explanations, Cosmological Arguments, and Sufficient Reasons”, Midwest Studies in Philosophy , 21(1): 188–99. doi:10.1111/j.1475-4975.1997.tb00523.x
  • Rundle, Bede, 2004, Why There Is Something Rather than Nothing? Oxford: Clarendon Press.
  • Russell, Bertrand, 1937, The Principles of Mathematics , second edition (first edition 1903), London: George Allen & Unwin.
  • Russell, Bertrand and Frederick Copleston, 1948, “Debate on the Existence of God”, Reprinted in John Hick (ed.), 1964, The Existence of God , New York: Macmillan, 167–90.
  • Scotus, John Duns, c. 1300 [1964], Ordinatio , in Philosophical Writings: A Selection , Allan Wolter (trans.), Indianapolis: Bobbs-Merrill Co., 1964.
  • Silk, Joseph, 2001, The Big Bang , San Francisco: W.H. Freeman.
  • Sincell, Marc, 2001, “Farthest Supernova Yet Bolsters Dark Energy”, Science , 292(5514): 27–28. doi:10.1126/science.292.5514.27a
  • Small, Robin, 1986, “Tristram Shandy’s Last Page”, British Journal for the Philosophy of Science , 37(2): 213–16. doi:10.1093/bjps/37.2.213
  • Smart, J.J.C. and J.J. Haldane, 1996, Atheism and Theism , Oxford: Wiley-Blackwell.
  • Sobel, Jordan H., 2004, Logic and Theism: Arguments For and Against Beliefs in God , Cambridge: Cambridge University Press, chaps. 5 & 6.
  • Swinburne, Richard, 1977, The Coherence of Theism , Oxford: Clarendon Press.
  • –––, 1979, The Existence of God , Oxford: Clarendon Press.
  • –––, 1983, “Mackie, Induction, and God”, Religious Studies , 19(3): 385–91. doi:10.1017/S0034412500015316
  • –––, 1993. The Coherence of Theism , revised edition, Oxford: Clarendon Press.
  • –––, 1996, Is There a God? Oxford: Oxford University Press.
  • –––, 2001, Epistemic Justification , Oxford: Oxford University Press.
  • –––, 2004, The Existence of God , revised edition, Oxford: Oxford University Press.
  • –––, 2007, Revelation: From Metaphor to Analogy , 2 nd edition, Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199212460.001.0001
  • –––, 2010, “God as the Simplest Explanation of the Universe” European Journal for Philosophy of Religion , 2(1): 1–24.
  • –––, 2012, “What Kind of Necessary Being Could God Be?” European Journal for Philosophy of Religion , 4(2), 1–18.
  • –––, 2016, The Coherence of Theism , 2 nd edition, Oxford: Clarendon Press. doi:10.1093/0198240708.001.0001
  • Taylor, Richard, 1992, Metaphysics , Englewood Cliffs: Prentice-Hall.
  • Udayana, [c. 900] 1996, Nyāyakusumāñjali , Indian Council of Philosophical Research, New Delhi: Distributed by Munshiram Manoharlal Publishers.
  • van Inwagen, Peter, 1983, An Essay on Free Will , New York: Oxford University Press.
  • White, John D., 1979, “God’s Necessity”, International Journal for Philosophy of Religion , 10(1): 177–87. doi:10.1007/BF00143165
How to cite this entry . Preview the PDF version of this entry at the Friends of the SEP Society . Look up this entry topic at the Indiana Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers , with links to its database.
  • Defense of the kalām argument .
  • Reply to Oppy (Number 2).
  • Reply to Sobel.
  • Reply to Oppy (Number 1).
  • Oppy, Graham, Critiques of the kalām argument .
  • Muehlhauser, Luke, Bibliography on the theistic arguments and the cosmological argument .
  • Rutten, Emanuel, 2012, A Critical Assessment of Contemporary Cosmological Arguments: Toward a Renewed Case for Theism , Amsterdam.
  • Smith, Quentin, Critiques of the cosmological argument .

Aquinas, Saint Thomas | Bonaventure, Saint | Clarke, Samuel | Duns Scotus, John | Hume, David | Hume, David: on religion | Kant, Immanuel | Leibniz, Gottfried Wilhelm | ontological arguments | Russell, Bertrand

Copyright © 2016 by Bruce Reichenbach < reichen @ augsburg . edu >

Support SEP

Mirror sites.

View this site from another server:

  • Info about mirror sites

Stanford Center for the Study of Language and Information

The Stanford Encyclopedia of Philosophy is copyright © 2016 by The Metaphysics Research Lab , Center for the Study of Language and Information (CSLI), Stanford University

Library of Congress Catalog Data: ISSN 1095-5054

Marked by Teachers

  • TOP CATEGORIES
  • AS and A Level
  • University Degree
  • International Baccalaureate
  • Uncategorised
  • 5 Star Essays
  • Study Tools
  • Study Guides
  • Meet the Team
  • Religious Studies & Philosophy
  • Philosophy & Ethics

Explain the cosmological argument for existence of God

Authors Avatar

The cosmological argument is an a posterior argument which has a long history, going back to the great classical philosophers of Plato, Aristotle, Leibnitz and Kant. All of them believed that the universe was the result of a transcendent being called G-d. Although these philosophers may have had different ideas about G-d, they all agreed that the universe was not self explanatory and must have had a sole cause in order for it to come into existence. Although the cosmological argument had various forms, each version focused on a key fundamental question: Why the universe began, why it was created and who or what created it. The case for the Cosmological Argument is best and most famously put forward by St Thomas Aquinas in his book Summa Theologicae which contained the ‘Five ways’

The argument starts off with his rejection of the ontological argument, as he says “[...] an argument that says G-d’s existence is self-evident we cannot use [...] as we can’t see the self evidence.” He argued that one first needs to argue about G-d from evidence we find in the world today. This is quite an Aristotelian concept; Aristotle was a philosopher who Aquinas studied in Cologne and translated his works.

His first argument was the “Unmoved mover” argument. The argument is concerned with things which change. Everything that is in motion is moved by something else, infinite regress is impossible; therefore there must be a first mover. The movement, to which Aquinas is referring, is the movement from one state to another, from potentiality to actuality. This is not an argument relating to the beginning of the universe; rather it relates to the way everything depends on something else for the changes to occur. For Aquinas, the changes that occur from moment to moment depend on the first mover (i.e. G-d).

This first argument is very similar to the next argument which Aquinas called the uncaused cause, or the first cause argument. In the world we find an order of efficient causes. There is no case where a thing is found to be its own efficient cause; it would have to exist before itself, which is impossible. Efficient causes cannot go to infinity because the first cause is the cause of middle and the middle cause of end cause. Without a cause, there is no effect. If causes went to infinity, there would be no intermediate cause and no present effect for us. There must be a first efficient cause, which is in itself uncaused.  The focus of this argument is again on dependency, that everything depends upon something else to cause it. The difference between this argument and the first argument is that this argument is focused upon the things that causes something to change, rather than the things themselves which change which is what the unmoved mover is concerned about. The Unmoved mover focuses on the present moment whereas the first cause focuses on the past up to the present moment. It is therefore logical to see how the two fit together to give a greater understanding. For example wood has the potential to turn into fire, but it needs the cause of a spark in order to move from potentiality to actuality and turn into the fire.

Join now!

This is a preview of the whole essay

The third part of his argument is the necessity and contingency argument. The argument states that some contingent beings exist, if any contingent being exist, then a necessary being must exist (because contingent beings require a necessary being as their ultimate cause), therefore there exists a necessary being (which is the ultimate cause of the existence of contingent beings). C.S. Evans expounds the argument “[...] Ultimately the explanation of contingent beings’ existence will be incomplete unless there exists a necessary being, a being which cannot fail to exist, who is the cause of all contingent beings[...].” The argument needs an explanation of the definition of contingency and necessity. The necessary being is formally known as G-d who is needed to start off the chain of dependant beings which need a necessary being in order for their existence to come about. A Dependant being cannot exist without being caused to exist by something which is not dependant and in itself has not been caused to exist which would suggest a dependency.

The kalam cosmological argument stems from Aquinas’s cosmological argument. It has recently been restored to popularity by William Lane Craig. Like all cosmological arguments, the kalam cosmological argument is an argument from the existence of the world or universe to the existence of God.

This argument has the following logical structure: Everything that has a beginning of its existence has a cause of its existence. The universe has a beginning of its existence. The universe has a cause of its existence. If the universe has a cause of its existence then that cause is G-d, therefore G-d exists.

What distinguishes the kalam cosmological argument from other forms of cosmological argument is that it rests on the idea that the universe has a beginning in time. Modal forms of the cosmological argument are consistent with the universe having an infinite past. According to the kalam cosmological argument, however, it is precisely because the universe is thought to have a beginning in time that its existence is thought to stand in need of explanation.

In conclusion, all three arguments are interconnected to form the cosmological argument. The unmoved mover is concerned with the things themselves that change whilst the uncaused cause is focused on the things which cause them to change and the necessity and contingency argument explains the theory behind, all have the same consistency about the universe not being infinite, that it has a beginning which G-d caused, because G-d is the necessary being which causes all things contingent to exist in a chain of causes.

 To what extent is the cosmological argument convincing

The philosopher Leibniz supports Aquinas and his argument that there must be a necessary reality or being “We assume that things in the world happen for a reason, why can’t we assume this about the world as a whole?” here, he is saying that if we broaden the perspective of understanding from experience, we can apply it to the Universe, to find what out what created it. This supports Thomas Aquinas by using a posterior experience to suggest that the universe and world were created by G-d.

However contrary to the logical outline of Thomas Aquinas’s argument, there have been many criticisms against it. The idea of infinite regress is one that Aquinas was against, he says that there must be a first cause, however it is possible there was no such first cause.  William Temple argued “[...]It is impossible to imagine infinite regress [...] but it is not impossible to conceive it.” He meant that something is unthinkable if we cannot hold the concept without contradiction. But infinite does not contradict regress; you can imagine infinity but not think of it, the world makes sense as a concept and can be understood. It is impossible to imagine which shows the limitation of the imagination but not of things. If infinite regress exists, then in that amount of time anything could happen, for example someone trying to roll a dice to get a six one million times in a row, however unlikely it may seem, if an infinite amount of time was had, it would happen. Although logical, Aquinas is detracting people from a possible argument for the existence of G-d by refusing to concede the possibility of infinite regression.

David Hume also argued against Aquinas’ argument. Hume believed “there is no necessary connection between cause and effect” meaning that we look at one thing and believe that we are seeing it causing an effect on another, what we are really seeing is one thing happening close to another. Relating to the world Hume believed that even if it did have a cause for being, then it would be impossible to see what it is, as we have not experienced the creation of the universe and cannot get out of the universe to see what caused it when nothing was before. Hume believed that all knowledge comes from experience, and thus the experiences we have had are not adequate for us to know if G-d created the universe.

Hume’s views are also supported by Russell, who argued that just because we see individual things as having a cause; it doesn’t mean that the universe also has a cause. Russell believed that the universe was a “Brute Fact” and it was “just there, and that’s all there is to say about it.” Although the Universe is “there” it doesn’t mean it had a beginning.

Modern scientists say that Aquinas’ argument rests on the assumption no longer necessarily holds true that is “Everything must have a cause. However, scientists have proven that sub-atomic particles can come into existence without a cause. Scientists today do not believe that it is a law that everything must have a cause, especially at the subatomic level in modern physics. This is damaging to Aquinas’ argument because one of the conditions is that everything in the universe had a cause, and this leads back to the original or first caused. Stephen Hawking comments “If the universe was completely self contained it could neither be created nor destroyed” which means that there is no need for an outside cause in the creation of the universe or a beginning.

The continuous questioning is part of human nature; we cannot accept that the universe may be a reality itself as it has not been proved. We therefore need to believe in some external intelligent creator. Aquinas’ argument is the product of this human belief that we are the subject of design, in a series of causes and effects that can be traced to a definite cause which itself is uncaused. Aquinas is logical and his argument understandable, but it is in human logic and experience which it is trapped and ultimately flawed where it fails to look beyond human experience, something that we cannot ourselves image but rather to speculate over, as we will never know what there was before the universe. Hume’s criticisms consider these possibilities, making sense logically and outlining limitations of human experience. “We are prepared to argue that because there are causes of things within the universe, there is a cause for the universe as a whole?” we do not know because there is no way of knowing. Furthermore, the advances made in recent years in science shows that an effect does not follow on from a cause from the advances in quantum physics, particles can come into existence with no explanation.

Aquinas’ logic is understandable, and the argument believable, but it is the expansion of human understanding and discovery that will eventually disprove logic as we encounter the unbelievable and so his argument will become less convincing over time.

Explain the cosmological argument for existence of God

Document Details

  • Word Count 1838
  • Page Count 3
  • Level AS and A Level
  • Subject Religious Studies & Philosophy

Related Essays

Explain the cosmological argument for existence of God

Explain the Cosmological Argument for the existence of God, according to Aq...

Explain Aquinas cosmological argument for the existence of God.    Humes criticisms alone completely discredit the cosmological argument  Discuss.

Explain Aquinas cosmological argument for the existence of God. Humes cr...

outline the cosmological argument for the existence of God

outline the cosmological argument for the existence of God

IMAGES

  1. The Cosmological Argument

    the cosmological argument essay

  2. Anselm's Cosmological Argument Essay Example

    the cosmological argument essay

  3. Cosmological argument essay

    the cosmological argument essay

  4. Cosmological Argument by Annie Davey

    the cosmological argument essay

  5. The Cosmological Argument for the Existence of God

    the cosmological argument essay

  6. ⇉Explain the Cosmological Argument for the Existence of God Essay

    the cosmological argument essay

VIDEO

  1. Cosmological Argument Episode 2

  2. Cosmological Argument Episode 9

  3. Cosmological Argument Episode 6

  4. Cosmological Argument Episode 13

  5. Cosmology Model Wrong And Universe Isn't What We Think

  6. Cosmological Argument Episode 12

COMMENTS

  1. Essay: Cosmological Argument

    This essay, of A grade standard, has been submitted by a student. PB. The Cosmological argument is an argument put forward by the Christian Philosopher St. Thomas Aquinas (1225-1274) in an attempt to prove God's existence. However, it is important to take into account that Aquinas already had a strong belief in God when putting this theory ...

  2. The Cosmological argument

    The Cosmological argument. Cosmological arguments attempt to justify the conclusion that God exists as the required explanation of the existence of the universe. A posteriori. Cosmological arguments are typically a posteriori arguments, which means they are based on experience. The cosmological argument is based on observation of everything in ...

  3. Cosmological Argument

    The cosmological argument is less a particular argument than an argument type. It uses a general pattern of argumentation (logos) that makes an inference from particular alleged facts about the universe (cosmos) to the existence of a unique being, generally identified with or referred to as God.Among these initial facts are that particular beings or events in the universe are causally ...

  4. Cosmological argument

    cosmological argument, Form of argument used in natural theology to prove the existence of God. Thomas Aquinas, in his Summa theologiae, presented two versions of the cosmological argument: the first-cause argument and the argument from contingency.The first-cause argument begins with the fact that there is change in the world, and a change is always the effect of some cause or causes.

  5. 6.3 Cosmology and the Existence of God

    Teleological arguments consider the level of design found in living organisms, the order displayed on a cosmological scale, and even how the presence of order in general is significant. Aquinas's Design Argument. Thomas Aquinas's Five Ways is known as a teleological argument for the existence of God from the presence of design in experience.

  6. Cosmological Arguments for the Existence of God

    Cosmological arguments for God's existence propose that God is the ultimate explanation or cause of everything. Such arguments begin with an empirical observation of the world—that there is motion, or causes, or just ordinary things that exist—and conclude this observation is explained by God's existence.[1] This essay surveys three ...

  7. Cosmological argument

    A cosmological argument, in natural theology, is an argument which claims that the existence of God can be inferred from facts concerning causation, explanation, change, motion, contingency, dependency, or finitude with respect to the universe or some totality of objects. [1] [2] [3] A cosmological argument can also sometimes be referred to as ...

  8. PDF The cosmological argument

    THE ARGUMENT FROM CONTINGENT EXISTENCE. This version of the cosmological argument, defended by Frederick Copleston in a radio debate with Bertrand Russell, emphasises the need to explain what exists. 1. Things in the universe exist contingently, they might not have existed or they might stop existing.

  9. PDF The Cosmological Argument for the Existence of God

    Aquinas had five proofs for the existence of God, of which three are cosmological; they are the First Cause Argument, the Prime Mover Argument and the Argument from Contingency. The First Cause Argument is concerned with the fact that all things have a cause, and if we trace back the chain of causes, there must be an initial cause which began ...

  10. The Cosmological Argument

    The cosmological argument is the argument to God from the existence of a complex physical universe. God can create such a universe (by keeping it in existence as long as it exists — whether for a finite or infinite time) and has good reason to do so because it is necessary for the existence of humanly free persons. But it is a far less simple ...

  11. The Cosmological Argument

    Abstract. The cosmological argument is the argument to God from the existence of a complex physical universe. God can create such a universe (by keeping it in existence as long as it exists ‐ whether for a finite or infinite time) and has some reason to do so because it is a theatre for finite agents which they can shape and in which they can ...

  12. Philosophy of Cosmology

    Philosophy of Cosmology. First published Tue Sep 26, 2017. Cosmology (the study of the physical universe) is a science that, due to both theoretical and observational developments, has made enormous strides in the past 100 years. It began as a branch of theoretical physics through Einstein's 1917 static model of the universe (Einstein 1917 ...

  13. Clarke's Cosmological Argument Essay (Critical Writing)

    Clarke's argument stands out because the author draws a very sharp contrast between contingent and necessary beings. We will write a custom essay on your topic. This paper looks at Clarke's cosmological argument. Specifically, it provides a thesis and proceeds to give a discussion on the argument, objections, and responses by different ...

  14. The Cosmological Argument

    The Cosmological argument is defined as the argument which proves that the universe was created by God. This argument clearly claims that the world is in existence because it was created by God. The first arguments are said to have been generated by Aristotle around three hundred years before Christ. We will write a custom essay on your topic.

  15. The Cosmological argument summary notes

    Aquinas created three versions of the cosmological argument. 1st way: motion. Everything is in motion. There can't be an infinite regress of motion. It cannot be that there is just an infinite chain of movers going back in time forever. There has to have been a first mover - a start to the motion we observe.

  16. Cosmological Argument: St. Thomas Aquinas

    Get original essay. The Cosmological Argument has got its basis from St. Thomas Aquinas, who in his book "Summa Theologica" has proved the existence of God in five ways. However, it is the first three proofs that are Cosmological and explain about the existence of God. These three Cosmological proofs are: a) the theory of First Mover, b ...

  17. The Kalam Cosmological Argument

    Ghazali formulates his argument very simply: "Every being which begins has a cause for its beginning; now the world is a being which begins; therefore, it possesses a cause for its beginning.". [1] Ghazali's reasoning involves three simple steps: 1. Whatever begins to exist has a cause of its beginning. 2.

  18. Aquinas

    The three major arguments put forth by Aquinas known as the Cosmological Argument will be discussed here. In his work, Summa Theologica Thomas Aquinas offered five 'proofs' for the existence of God. The first argument was the Argument of Motion. Aquinas's argument has to be understood keeping in mind Aristotle's discussion of Astronomy.

  19. Cosmological Argument

    The cosmological argument is less a particular argument than an argument type. It uses a general pattern of argumentation (logos) that makes an inference from particular alleged facts about the universe (cosmos) to the existence of a unique being, generally identified with or referred to as God.Among these initial facts are that particular beings or events in the universe are causally ...

  20. COSMOLOGICAL ARGUMENT HIGH MARK ESSAY

    Lastly, I will explore the modern science arguments made by Quentin Smith and Werner Heisenberg. Ultimately, I shall agree with Kant's criticisms of Aquinas and disagree with the statement. Aquinas posed an a posteriori argument based on empirical evidence via observation of the world, known as the cosmological argument.

  21. Explain the cosmological argument for existence of God

    Explain the cosmological argument for existence of God. The cosmological argument is an a posterior argument which has a long history, going back to the great classical philosophers of Plato, Aristotle, Leibnitz and Kant. All of them believed that the universe was the result of a transcendent being called G-d.

  22. Cosmological Argument Essay

    The cosmological argument is an a posteriori argument which intends to prove that there is an intelligent being that exists; the being is distinct from the universe, explains the existence of the universe, and is omniscient, omnipotent, omnipresent and omnibenevolent. The basic notion of cosmological arguments is that the world and everything ...