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.

A Level Philosophy & Religious Studies

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 the universe having a cause, being in motion or being contingent and therefore requiring a creator. These observations form the premises of cosmological arguments. On the basis of those observations, an inference is then made to the nature of the origin of the universe being God.

Cosmological arguments from causation

Aquinas’ first and second ways.

Aquinas’ first two ways are developed from Aristotle’s theory of efficient causation, which is an attempt to explain the change we observe. Aristotle thought that change required a prime mover which sustains the motion and causation we experience.

Efficient causation involves sustaining causes, those which bring about their effect continuously, such that if they ceased to exist then their effect would also cease to exist. E.g., the gravity of the earth causes the moon to be in orbit, which in turn causes the sea tides to rise and fall on earth.

Aquinas’ 1 st way (motion)

P1. We observe that there are things in motion. P2. Motion is the actualization of a thing’s potential to be in motion. P3. A thing can only come to be in motion by being moved. P4. A mover must be something that is actual, i.e., in a state of actuality. P5. A thing cannot move itself. C1. So, all things in motion must have been moved by a mover, which was also moved by another mover. P6. There cannot be an infinite regress of movers, otherwise there would be no first mover and then no motion. C2. Therefore, there must be a first mover which must itself be unmoved (as it is pure actuality). That thing we call God.

Change requires going from potential to actual, which depends on something that is actual, which cannot depend merely on other potential things, so there must be something of pure actuality. A thing that is purely actual with no potential cannot change, it is an unmoved mover or uncaused causer.

Aquinas’ 2 nd way (atemporal causation)

P1. We observe efficient causation. P2. Nothing can cause itself. P3. There is a logical order to sustaining causes: the first cause, then intermediate causes, then an ultimate effect. P4. If A is the efficient cause of B, then if A doesn’t exist neither does B. C1. There must be a first sustaining cause, otherwise P1 would be false as there would be no further sustaining causes or effects. C2. As there is a first cause, there cannot be an infinite regress of causes. C3. The first cause must itself be uncaused. That thing we call God.

Aquinas’ first two ways treat the relationship between cause and effect as ontologically real but not temporal, although they are consistent with a temporal understanding of cause and effect. They point to the logical implications of there being sustaining causes. This is why especially Aquinas’ 2 nd way is called a cosmological argument from ‘atemporal causation’.

The first and second way attempt to show God must exist as the first mover or causer. The word ‘first’ in the concept of a first cause or first mover is not meant to indicate it being ‘first’ in time, but ontologically first in the sense that motion and causation are ontologically dependent on it.

The Kalam cosmological argument from temporal causation

The Kalam cosmological argument explicitly involves temporal causation, which is when a cause brings about its effect after it and the continued existence of the effect is independent of the existence of the cause.

The Kalam infers a beginning cause rather than a sustaining cause. The causal sequence being temporal, with God as the beginning cause, is a central feature of the argument.

One advantage of the Kalam argument is that it’s easier to explain how an atemporal God could create the world in one act compared to a sustained act of creation over time.

W. L. Craig brought this argument to prominence in the late 20 th century and named it ‘Kalam’ after the Islamic philosophy which first invented in the 11 th century.

P1. Everything that begins to exist has a cause of its existence. P2. The universe began to exist (an infinite regress is not possible). C1. So, the universe has a cause of its existence.

Further steps are required to show that the cause of the universe is God. Craig firstly argues that scientific explanation applies within the universe and therefore cannot apply to its actual creation. The cause of the universe must therefore have a personal explanation, i.e., intentionally created by an intelligent mind. This being must have the power to create a universe from nothing (ex nihilo). It must be outside time and space since it created time and space. As a timeless, eternal being, God did didn’t begin to exist so it’s then no contradiction in claiming that God doesn’t have a cause. These are qualities that God would have, so the cause of the universe is God.

A strength of the Kalam argument is W. L. Craig’s arguments for the impossibility of an ‘actual infinite’ meaning an infinite in reality. The problem is that sets with infinite members can be equal in size to their subsets. Craig uses the illustration of a library with an infinite number of books, half of which are red. Half of infinity is still infinity, so half of infinity is not actually smaller than infinity. This might make sense theoretically, but Craig claims problems arise when applying it to reality. It would mean the set of red books ‘X’ is apparently smaller than the set of all the books ‘Z’, and yet paradoxically is also equal in size, since both are infinite. Infinities therefore cannot actually exist, since then they could be both smaller than and the same size as other infinities.

Counter: An infinite series is actually possible.

G. Cantor argued that the mathematical properties of infinite sets/things are simply radically different to those that are finite, making Craig’s library or Hilbert’s Hotel not absurd. Craig takes it to be obviously absurd for a subset (red books) to be both smaller than and equal to its set (all the books). However, it’s only absurd for finite sets. For infinite sets it’s not absurd, it’s actually their defining characteristic. When we think of libraries or hotels, we have in mind our ideas about finite sets of things, but Cantor argued such intuitions are not applicable to infinite sets. Infinite sets simply have different mathematical properties, one of which is the possibility of a one-to-one relation between the number of members of infinite sets and those of their subsets.

Hume’s objection to the ‘causal principle’

The strength of cosmological arguments from causation is that they are based on the causal principle, which is that every event has a cause, or that every contingent being has a cause of its existence.

Ex nihilo nihil fit (nothing comes from nothing) goes back to the ancient Greeks like Parmenides who influenced Aristotle and later Aquinas. It’s not possible for an event to happen without a cause, otherwise something could come from nothing, which is absurd.

Arguing for the Kalam cosmological argument, William Lane Craig claims that the causal principle is:

“based on the metaphysical intuition that something cannot come out of nothing.” – W, L. Craig

A causal principle is one of the essential pillars of cosmological arguments, alongside arguments against the possibility of infinite regression and inferences to the divine nature of the ultimate origin of the universe.

Weakness: Hume’s objection to the causal principle.

Hume’s fork tells us that propositions (such as the causal principle) can either be analytic or synthetic.

Hume argues that the causal principle is not true by definition (analytic). There doesn’t appear to be anything incoherent in the idea of an event or thing existing without a cause. It is conceivable and not obviously self-contradictory. We can imagine something popping into existence without a cause. The idea of a four-sided triangle is obviously self-contradictory because the idea of triangle contradicts the idea of four-sides. Yet, the idea of an event doesn’t seem contradicted by the idea of no cause.

So, the causal principle can only be justified on a posteriori grounds, which makes it a synthetic truth. The problem is, claims based on experience cannot be known with certainty to be true in all cases. All we can justifiably claim is that every event we have observed has a cause. It does not follow from this that all events have a cause, since we have not experienced all events.

So, the universe could exist without a cause. The cosmological argument therefore fails because in attempting to argue for God’s existence as the required explanation of the universe, it assumes that the universe has a cause.

Lawrence Krauss & Alan Guth add evidence from modern theoretical physics which resonates with Hume’s point. The universe has zero total energy and therefore required no energy to be created So, it could have come from nothing.

Evaluation defending the cosmological argument:

The cosmological argument could be defended by arguing that the universality of the causal principle is justified through induction. We have experienced many causal interactions, all of which involved the constant conjunction of cause and effect. From this we can infer that all effects have a cause. It is possible that this may be false, that in some cases the causal principle may not hold. However, the evidence so far suggests that it always holds. Therefore, we are empirically justified in accepting the causal principle.

Evaluation criticizing the cosmological argument:

Furthermore, we cannot even be empirically justified in holding the causal principle. All the evidence that we have for the causal principle comes from our observation of change within the universe itself. We are not justified in assuming the relevance of that experience to the conditions under which the universe itself came to exist. Those conditions, whatever they are, could be radically different to anything we now observe within the universe. So, we cannot even justify the causal principle on an empirical basis.

Cosmological arguments from contingency

Aquinas’ 3rd way (contingency).

P1. We observe that there are contingent beings. P2. A series of contingent beings cannot regress infinitely into the past. C1. So, a series of contingent beings must be finite. P3. If this finite series was all that existed, then before it would be nothing. P4. If there was once nothing, there would be nothing now, which is absurd. C1. So, there must be more than this finite series of contingent beings, i.e., a necessary being. P5. There cannot be an infinite regress of necessary beings. C3. There must be a necessary being “having of itself its own necessity … That thing we call God.”

Leibniz’ principle of sufficient reason

Leibniz improves on Aquinas’ 3 rd way by removing unnecessary reasoning about nothing once existing.

Leibniz’ argument is a priori, it doesn’t require inference from experience. The downside of a posteriori arguments is that they are defeasible, meaning in principle future experiences could always prove them false. A priori arguments based on logic are stronger.

Leibniz bases his argument on the principle of sufficient reason, which does both the job of a causal principle and an argument against the infinite regress. It shows that there must be not just any causal explanation, but a causal explanation which provides an ultimately sufficient reason for everything that exists. This strengthens the argument by making it dependent on only one claim.

P1. For every true fact or assertion, there is a “sufficient reason why it is thus and not otherwise.” P2. There are two types of truth: truths of reasoning and truths of fact. P2a. Truths of reasoning are necessary, so their opposite is impossible. The sufficient reason for truths of reasoning can be discovered a priori. P2b. Truths of fact are contingent, so their opposite is possible. The sufficient reason for truths of fact cannot be discovered through other contingent truths, because they too require a sufficient explanation, and so on. C1. A sufficient reason for contingent facts must be found outside a series of contingent things. C2. The sufficient reason for contingent facts must be a necessary substance. C3. That necessary substance is God. C4. So, God exists.

Leibniz claims that the principle of sufficient reason (P1) can be known as a necessary truth. Even if we can’t know or even find out what the reason is, there must be one. ‘From nothing, nothing comes’ because nothing is not sufficient to create something. Only a necessary being is sufficient to explain the universe because otherwise there would be an infinite chain of contingent beings, but an infinite regress alone can have no sufficient explanation.

If things have always existed going back forever then nothing would have a sufficient reason for its existence. Everything’s reason for existence would consist something for which its reason for existence consists in something else. There would simply be an infinite deferring of the reason for existence and thus there would not actually be a reason for existence. So, a necessary being must have begun the chain of contingent beings and is the sufficient explanation of the universe.

The fallacy of composition

The strength of cosmological arguments from contingency is their seeking an ultimate explanation rather than only a first cause.

They focus more fundamentally on the nature of things. Beings have causal relations, but also an ontological status, i.e., contingency or necessity. Contingent beings & series must have an external explanation since it is their nature to depend on something else for their existence.

Weakness: Hume & Russell on the fallacy of composition.

This criticism explicitly attacks cosmological arguments from contingency, but if successful would also undermine arguments from causation. It is a fallacy to assume that what is true of a thing’s part(s) must also be true of the whole. Bertrand Russell illustrated this by pointing out that just because every human (parts) has a mother, that doesn’t mean the human race (whole) has a mother.

Hume uses the example of a finite set of contingent beings (20 particles). Just because the parts of a set/series of contingent beings have an explanation, that doesn’t mean the whole set/series has an explanation.

Experience shows that parts of the universe are contingent & have a cause/explanation. This doesn’t mean the universe itself as a whole must also be contingent (have a cause/explanation).

The only empirical way to believe a whole series has an explanation is to commit the fallacy of composition by assuming the whole is like the parts. So, a posteriori cosmological arguments commit the fallacy of composition by assuming that the universe has a cause when all we experience is that parts of the universe have a cause. That’s the problem with Aquinas’ 3 rd way.

Evaluation defending the cosmological argument

However, Copleston rightly points out that contingency arguments don’t actually appear to make an inference from parts to whole.

They argue that a series of contingent things must have an external cause. Copleston argues:

P1. A series is either caused or uncaused. P2. If a series is uncaused, the reason for its existence must be internal to it, making it necessary. P3. No amount of contingent things can be necessary, not even an infinite number of them, so a series of contingent things cannot be necessary. C1. So, a series of contingent things must have an external cause like a necessary being.

However, Hume’s empiricism drives his point further. Hume questions how we know a series must have a cause/explanation. In P2, Copleston assumes that an uncaused series must have an internal explanation. Hume doesn’t see how experience can justify thinking a series must have an explanation at all.

However, Leibniz’ a priori argument may avoid this problem as it’s derived from intuitively known a priori necessary truths, not experience. Leibniz is a rationalist and believes the whole has an explanation because of his a priori principle of sufficient reason. Nothing can be true, like a whole series existing, without a sufficient explanation for why it is thus and not otherwise.

Copleston vs Hume & Russell on the universe as a brute fact

Hume and Russell take the fallacy of composition objection further in a way that can respond to Leibniz’ a priori approach.

Hume turns to the example of an infinite series of contingent beings. He argues it is possible and doesn’t need an ultimate sufficient reason. Each being is explained by the being it depends on.

Of course, Leibniz wants to object that this leaves the whole chain itself without explanation, without reason for existence.

Hume, however, denies the validity of claiming that there is ‘whole’ chain. If experience only shows that contingent beings exist, we can question whether there actually is a ‘whole’.

“uniting of these parts into a whole, like the uniting of several distinct counties into one kingdom … is performed merely by an arbitrary act of the mind and has no influence on the nature of things.” – Hume

“I think the word “universe” is a handy word in some connections, but I don’t think it stands for anything that has a meaning … The whole concept of cause is one we derive from our observation of particular things; I see no reason whatsoever to suppose that the total has any cause whatsoever … it doesn’t need to be its own cause, what I’m saying is that the concept of cause is not applicable to the total.” – Russell.

We cannot say the ‘whole universe/series’ is even a valid concept, let alone that it is contingent. So, we can’t conclude there is a need for an explanation like a necessary being. Cosmological arguments from contingency fail. Note that this isn’t quite saying the universe could be necessary. A necessary being contains its own reason for existence, but Hume and Russell are proposing that the universe could simply have no explanation at all.

All we can justifiably say is that the universe/series is “just there, and that’s all” (Russell) . It is a brute fact, one which has no explanation, whether causal or otherwise. Russell points to evidence from Quantum mechanics:

“The physicists assure us that individual quantum transitions in atoms have no cause.” – Russell.

So, arguments from contingency baselessly assume that a series (as a whole) must have an explanation at all.

This argument also counters the arguments from causation. If we have no basis for thinking the universe has an explanation of any kind, that includes causal explanation.

Copleston firstly responds that only some interpretations of quantum mechanics propose uncaused events. Copleston further argues:

“I cannot see how science could be conducted on any other assumption than that of order an intelligibility in nature”  – Copleston.

Copleston’s argument is successful because the ‘brute fact’ argument is self-defeating. Science and philosophy are about finding out why things are the way they are; their causes and explanations. If we say that there is no reason, then we undermine the purpose of science and philosophy itself.

Furthermore, Russell argues that while a scientist may look for causes, they do not assume that there is one to find. Even if quantum transitions are only uncaused on some interpretations of quantum mechanics, the mere idea of it still shows that physicists are able to conceive of events that have no cause. It is therefore at least logically possible for events in nature to have no cause or explanation. So, a contingent series doesn’t need an external cause. The principle of sufficient reason cannot be a necessary truth. The universe could be one of those natural events that has no cause or explanation. Science should accept that possibility, since science should be open to whatever could be true.

Copleston attempts to show that Russell is assuming that there isn’t a cause and Russell attempts to show that Copleston is assuming that there is. Ultimately, defenders of the cosmological argument are the ones making the positive claim about reality, so they have the burden of proof. It looks like for cosmological arguments from contingency, the conclusion that God exists as the explanation/cause of the universe cannot be reached without assuming that the universe has an explanation/cause. In that case, the mere possibility of the universe being a brute fact is enough to undermine the cosmological argument. It’s not irrational to look for a cause, even a cause like a God, but since there might not be a cause it is irrational to think that there must be one.

Issues around the possibility of a ‘necessary being’

A strength of cosmological arguments from contingency is that their conclusions achieve more than arguments from causation. They can establish God’s necessity, meaning inability to cease existing, which is a key element of Christian theology. Aquinas understands necessity to mean the inability to cease existing, which fits with the concept of omnipotence.

Weakness: Hume’s rejection of the possibility of a necessary being.

Hume’s fork:

A priori reasoning can only tell us about the relations between ideas , i.e. analytic knowledge (true by definition). E.g. “a bachelor is an unmarried man”.

A posteriori reasoning can only tell us about matters of fact , i.e. synthetic knowledge (true by the way the world is). E.g. “The sun will rise tomorrow”.

“there is an evident absurdity in pretending to demonstrate a matter of fact, or to prove it by any arguments a priori” – Hume.

A being whose existence is logically necessary is an absurdity to Hume. A thing’s existence is a matter of fact. All matters of fact can be conceived false and denied without contradiction. So, anything which could exist could also not exist. The concept of a being which ‘must’ exist, whose existence cannot be denied without contradiction, is therefore absurd and meaningless.

Hume’s justification for the fork is that truths of logic/definition are necessary, because they will be true no matter what happens regarding the factual state of the universe. E.g., 1+1 will always = 2, because there is no possible change to the universe which could make it false.

This shows there is a disconnect between logical (analytic) truths which are necessary (cannot be otherwise) and factual (synthetic) truths which can. The term “necessary existence” violates this disconnect. It is about what factually exists, so must be synthetic. Yet, it somehow also has the “cannot be otherwise” property that only belongs to logical analytic truths. We cannot know that a being’s existence is logically necessary, since a being’s existence cannot be established through logic.

“The words, therefore, necessary existence, have no meaning.” – Hume.

Any argument which attempts to conclude that God exists necessarily therefore fails, including the ontological argument and some cosmological arguments.

The masked man fallacy. Hume’s argument depends on conceivability entailing possibility. It is therefore susceptible to the masked man fallacy, which shows that we can conceive of the impossible. Imagine someone heard of a masked man robbing a bank. They can conceived that it is not their father. Yet, if it was their father, then it is impossible that it is not their father. Yet, that was what they conceived of. So, we can conceive of the impossible. When Hume argues that our ability to conceive of God not existing shows that it is possible for God to not exist and that God therefore cannot be necessary, he assumes that conceivability entails possibility. Conceiving of God’s non-existence could be conceiving of something impossible because God is necessary.

Hume’s argument is successful because it is epistemological, i.e., focused on what we can know. Hume isn’t really saying a necessary being is ‘impossible’, just that it is impossible for us to know that such a being exists. Aquinas, Leibniz and Craig say that it is a being with the property of the impossibility of non-existence, but Hume simply throws up his hands and says he cannot see how anyone can ascribe meaning to that statement. We might technically be able to understand the words, but we cannot conceive or understand how they could map onto reality. If we are to accept the term on that basis, Hume rightly goes on to say that we might as well say that, even though we can’t understand how, reality itself has this mysterious property of the impossibility of non-existence, so positing a God is unnecessary anyway.

Issues around the possibility of an infinite series

A crucial strength of all cosmological arguments are the justifications its proponents provide for the premise that an infinite series is impossible. This is essential, for if there is an infinite regress going back in time forever then all forms of the cosmological argument fail. God could not be concluded to exist as the origin of what exists if there was no origin. These arguments attempt to show that the idea of an infinite regress leads to paradox or absurdity and therefore must be false.

Aquinas argues an infinite regress is impossible. If there was an infinite regress, there would be an infinite amount of time before the present moment. That means to get to the present moment, an infinite amount of time must have passed. However, an infinite amount of time cannot pass. No matter how long you wait, even if you never stop waiting, the amount of time passed can never reach infinity. So, there cannot be an infinite amount of time before the present moment and therefore there cannot be an infinite regress. You cannot traverse an infinite through successive addition.

Weakness: The possibility of an infinite series.

The idea of an infinite regress may be difficult to comprehend, but there is no obvious logical contradiction in the idea. The concept of ‘time’ does not seem contradicted by the predicate of ‘infinitely regressing into the past’.

Critics of the infinite regress attempt to show that while there is no obvious contradiction, nonetheless paradoxes arise from the concept. This aims to show that it is illogical and therefore impossible.

However, Hume thinks we simply lack the evidence on which to properly judge whether it is possible or not. We can strengthen Hume’s point with modern science which is full of examples where physicists speculate about ways an infinite regress could exist without paradox.

Some physicists argue for the block universe, that the passing of time is an illusion. Others suggest the universe eternally cycles between expansion and collapse and that a new timeline begins each cycle. In that case, an infinite amount of time never passes (so Aquinas fails) and in fact a time line containing an actual infinite never existed (so Craig fails). Yet, an infinite series of cycles happened (just not on any particular time-line).

The meta point of these examples is that we know very little about what time and infinity are and how they work, so claims about the impossibility of an infinite series are unjustified. Philosophers who think they can draw conclusions about time through pure logic alone fail because our actual concept of time is so poorly scientifically understood and could bare very little relation to what time actually is. It seems Hume would say we should admit that we are not in a position to conclude either that there is or is not an infinite regress.

There is certainly much we don’t know about time and infinity. However, what we currently understand suggests Aquinas’ reasoning is right. The eternal bang/crunch model may be logically possible, but we have no reason to believe it. The evidence suggests the universe started with a big bang and there is no evidence it will collapse again.  We are justified in believing what we currently have most reason to believe.

Hume’s stance is the right one. We simply lack the required evidence to have a reasonable belief one way or the other regarding whether time regresses infinitely or not. Ultimately it is a matter for scientists and mathematicians and since they are far from decided, we are not justified in claiming an infinite regress is possible or impossible. Cosmological arguments therefore fail as they have the burden of proof. They are making a positive claim about what exists on the assumption of an infinite regress’ impossibility.

Extra credit:

Modern science’s rejection of the cosmological argument.

A reasonably popular view in physics had been the steady state theory, which was that the universe had always existed and that there is a continuous creation of matter which expands. However in the mid 20 th century, the discovery of the microwave background radiation provided very strong evidence against steady state and for the big bang theory. Steady state theory leaves no room for a creator God because it claims the universe was not created but has always existed, whereas the big bang arguably does leave room for God because the big bang could simply be how God created the universe. So, the big bang theory was actually welcomed by many Christians for this reason.

Pope Pius XII declared in 1951 that the big bang theory does not conflict with the Catholic conception of creation.

The question then becomes whether science can explain why the big bang happened without God. One possibility is the oscilating universe theory, that what we call our universe is just one of the cycles of big bangs and big crunches that has been oscilating forever.

However, another theory with more evidence for it is Alan Guth’s inflation theory. Laurence Krauss , a physicist, explains Guth’s theory that the universe came from nothing because it actually is nothing. Gravity has negative energy, the total amount of which in the universe happens to exactly cancel out the positive energy of the matter in the universe, so the total energy of the universe is zero; it is nothing. Krauss claims this answers Leibniz’ question of why there is something rather than nothing. Quantum mechanics has existed eternally and causes quantum fluctuations which can create a zero-energy universe from nothing because such a universe requires zero energy to create.

W. L. Craig’s response to Krauss is that Krauss’ definition of nothing is faulty since it’s actually something. Nothing really means total absence of anything.

Defence of Krauss: Arguably the definition of nothing philosophers have traditionally used is in fact not what nothing actually is though. Science has proven the philosophical conception of nothing to be inaccurate.

Encyclopedia Britannica

  • History & Society
  • Science & Tech
  • Biographies
  • Animals & Nature
  • Geography & Travel
  • Arts & Culture
  • Games & Quizzes
  • On This Day
  • One Good Fact
  • New Articles
  • Lifestyles & Social Issues
  • Philosophy & Religion
  • Politics, Law & Government
  • World History
  • Health & Medicine
  • Browse Biographies
  • Birds, Reptiles & Other Vertebrates
  • Bugs, Mollusks & Other Invertebrates
  • Environment
  • Fossils & Geologic Time
  • Entertainment & Pop Culture
  • Sports & Recreation
  • Visual Arts
  • Demystified
  • Image Galleries
  • Infographics
  • Top Questions
  • Britannica Kids
  • Saving Earth
  • Space Next 50
  • Student Center

Candles Burning On Table In Church

cosmological argument

Our editors will review what you’ve submitted and determine whether to revise the article.

  • Stanford Encyclopedia of Philosophy - 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. Each cause is itself the effect of a further cause or set of causes; this chain moves in a series that either never ends or is completed by a first cause , which must be of a radically different nature in that it is not itself caused. Such a first cause is an important aspect, though not the entirety, of what Christianity means by God. The argument from contingency follows by another route a similar basic movement of thought from the nature of the world to its ultimate ground.

Essay Plan: Cosmological Argument

October 30, 2012.

Print Friendly, PDF & Email

AS Religious Studies Revision: The Cosmological Argument

Ao1 material: i.e. ‘what goes in part a)’, a) explain the cosmological argument. (25).

The cosmological argument begins with the observation that the universe exists. It then asks the question ‘why is the universe here?’ or ‘why is there something rather than nothing?’

P1: Everything in the universe has a cause. P2: The universe itself must have a cause. P3: To avoid infinite regress of causes there must be an uncaused cause. C: This uncaused cause is God.

Type of argument

Inductive: inductive reasoning is where the premises support the conclusion, but they do not entail it. It is usually based upon information coming from the senses (the order and complexity we observe with our eyes). It is therefore not deductive, which is where the premises of an argument do entail the conclusion, i.e. the conclusion is necessary e.g. 1+1=2..

A posteriori: it is based upon experience: it comes ‘after the fact’ of order and complexity, it is not a priori which is based upon reasoning before experiencing. Synthetic: a proposition whose predicate concept is not contained in its subject concept. In other words, if I say ‘all triangles have three sides’; the fact that a triangle has three sides (predicate) is contained in the definition (subject). In the statement ‘there is design in the universe’ there is doubt because the predicate (design) is not contained in the subject (universe). We have to use our senses to verify the truth of this statement.

Scholars whose versions of the argument you must explain…(you need to do it in detail)

Thomas Aquinas: The first 3 of the 5 Ways.

Unmoved mover: everything that moves is moved by something else. There must be an initial cause of movement in the universe. There is an unmoved mover called God.

Uncaused cause: everything has a cause. There cannot be an infinite regress of causes. The first cause is God. Possibility and Necessity: everything in the universe exists contingently i.e. it could not exist. It is conceivable for everything in the universe to go out of existence. There must be something which cannot not exist (that exists necessarily). That something is God.

Aquinas argued that the world depends on God now for its existence.

The Kalam Argument:

This was suggested by Islamic scholars such as al-Ghazzali and argues that God is the originating cause of the universe. It goes: everything that has a beginning of existence must have a cause; the universe began to exist; the universe has a cause; the cause is God. Modern versions come from scholars such as William Lane-Craig.

Frederick Copleston: The Radio Debate

Copleston debated the existence of God with Bertrand Russell on the radio. Copleston provided a new version of Aquinas’ Five ways.

AO2: Critical evaluation i.e. ‘what do I put in part b)?’

b. The strengths fo the cosmological argument outweigh the weaknesses. Discuss (10)

Remember to read the question on the exam paper first before just regurgitating.

The strengths of the cosmological argument

The strengths of the cosmological argument are the strengths of inductive reasoning: inductive arguments begin with something that we can observe. Inductive reasoning begins with experience which may be universal (i.e. everyone has had it) or it may at least be testable. The argument does not rely upon fixed definitions that we must accept (unlike the Ontological Argument). The Cosmological argument fits in with the God of classical theism (omnipotent, omnibenevolent, omniscient). It makes sense to think that there is an initial cause to the universe: this fits with our experience of events within the universe. Most scientists would argue that the universe has a beginning, which fits in with the Cosmological argument.

The weaknesses of the Cosmological Argument.

The weaknesses of this argument are the weaknesses of inductive reasoning: the conclusion does not necessarily follow from the premises. Just because things in the universe have causes, doesn’t mean that the universe as a whole has a cause: we have no experience of universes being caused so cannot claim we know that they need a cause. (Bertrand Russell). The universe is just here and that is that: we do not need to ask why. (Bertrand Russell). There is no logical absurdity in claiming that things can come into existence without a cause. (Hume). Only analytic propositions (e.g. 1+1=2) can exist necessarily. (Kant). Perhaps the universe has always been here (Buddhism) and fluctuates in and out of existences (Big Crunch theory).

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

Leave a Reply Cancel

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

Home — Essay Samples — Philosophy — Philosophy of Life — An Analysis of Aquinas’ Cosmological Argument

test_template

An Analysis of Aquinas’ Cosmological Argument

  • Categories: Existence of God Philosophy of Life

About this sample

close

Words: 1378 |

Published: Oct 26, 2018

Words: 1378 | Pages: 3 | 7 min read

Works Cited

  • Aquinas, T. (1265-1274). Summa Theologica [Electronic version].
  • Retrieved from https://www.newadvent.org/summa/
  • Craig, W. L. (1979). The Kalam cosmological argument. London: Macmillan Press.
  • Craig, W. L. (2000). The Kalam cosmological argument. In W. L. Craig, Reasonable faith: Christian truth and apologetics (pp. 92-117). Wheaton, Ill.: Crossway Books.
  • Dawkins, R. (2006). The God Delusion. London: Bantam Press.
  • Feser, E. (2010). The Last Superstition: A Refutation of the New Atheism. South Bend, Indiana: St. Augustine's Press.
  • Leibniz, G. W. (1686). Discourse on Metaphysics [Electronic version].
  • Retrieved from http://www.earlymoderntexts.com/assets/pdfs/leibniz1686.pdf
  • Mackie, J. L. (1982). The miracle of theism: Arguments for and against the existence of God. Oxford: Oxford University Press.
  • Oppy, G. (2006). Arguing about Gods. Cambridge, UK: Cambridge University Press.
  • Rowe, W. L. (1975). The cosmological argument. Princeton, NJ: Princeton University Press.
  • Schaffer, J. (2005). Is there a fundamental level? Noûs, 39(4), 1-51.

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

Verified writer

  • Expert in: Religion Philosophy

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

4 pages / 1689 words

1 pages / 670 words

1 pages / 657 words

3 pages / 1234 words

Remember! This is just a sample.

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

121 writers online

An Analysis of Aquinas’ Cosmological Argument Essay

Still can’t find what you need?

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

Related Essays on Philosophy of Life

Escapism is a phenomenon deeply ingrained in human nature, offering a respite from the complexities and challenges of everyday life. The desire to retreat into alternate realities, whether through literature, entertainment, or [...]

Revenge is a powerful emotion that has been the subject of numerous literary works, films, and psychological studies. It is a natural human response to feeling wronged or mistreated and is often seen as a form of justice or [...]

Life is a journey filled with ups and downs, and often when we face challenges or difficulties, we may lose hope or feel overwhelmed. In such moments, it is essential to remember the age-old saying, "Every cloud has a silver [...]

Immortality and mortality are two fundamental themes in Natalie Babbitt's novel, Tuck Everlasting. The novel explores the concept of eternal life and its implications, as well as the inevitability of death. Through the Tuck [...]

In this paper I shall briefly define what induction is and attempt to explain David Hume’s problem of induction through examining the thre most common problems of induction, which are, the problem of the uniformity of nature, [...]

Philosophy is an academic field concerned with the generation of a clear understanding of such phenomenon as nature, the confines of knowledge, and the rules of moral judgment. As such, the field is made up of five branches: [...]

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

cosmological argument essay conclusion

Rudolf Carnap’s View on Cosmological Argument 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

There are many arguments and justifications regarding the existence of God. Some people argue that there is God, while others believe that God does not exist. The arguments that support the existence of God can be divided into two categories; the posterior arguments and the ‘ a priori’ arguments (Rowe 16). The former refers to the arguments that are developed based on the premise that is only known through the experience that people have in the world.

The latter, on the other hand, is based on the premises that can be known without the world experiences. One of the arguments that have been developed is the Cosmological Argument. This is an argument that supports the fact that there is a God. The other two arguments are the Ontological Argument and the Design Argument. The Cosmological Argument is a posteriori argument. The Design Argument is also a posteriori argument, but the Ontological Argument is a priori argument .

The Cosmological Argument has two parts. The first part attempts to prove the existence of a being that is not like other beings, while the second part of the argument proves that this special being has some features, such as perfect goodness, omniscience, and omnipotence. Moreover, this being is the creator of the world, and it is separate and independent of the world. The two parts of the Cosmological Argument face objections from other philosophers, with the first part being the most objected.

The first part states that “every being (that exists or ever did exist) is either a dependent being or a self-existing being. Not every being can be a dependent being; therefore, there exists a self-existing being” (Rowe 18). A dependent being is a kind of being whose existence can only be accounted for by the results of the activities carried out by other things. In other words, it is the kind of being whose existence can only be explained by another being.

On the other hand, a self-existent being is a being whose existence does not have to be explained by another being, but its own nature accounts for its existence. God is a self-existing being (Rowe 17). The premises of the first part of the Cosmological Argument and the part that has been focused on in the article have been criticized for not being deductively valid. In other words, they are said to be true, but their conclusions are not true.

For instance, the argument states that God is a self-existing being, but it fails to provide enough grounds to prove that God really exists. The premises are true, as God is truly a self-existing being. However, the conclusion that God exists is not true because they fail to prove that God truly exists. The argument is good because it is able to justify the existence of a special being, but it is not deductively valid because it fails to prove the existence of God.

The Carnap’s discussion about the explanation using the laws can be connected to the Cosmological Argument. The argument by Carnap does not explain why God exists. Instead, it tries to explain how God exists.

It gives the laws that try to explain how God exists through three premises. For instance, the third premise, which could be termed as the third law, explains that God is a self-existent being. In other words, God’s existence does not depend on the existence of another being. Therefore, the Cosmological Argument is consistent with Carnap’s discussion of explanation by the laws (Carnap 142).

Works Cited

Carnap, Rudolf. The Value of Laws . New York: IMG, 1966. Print.

Rowe, L. William. The Cosmological Argument. New York: Fordham University Press, 1998. Print.

  • Problem of Evil and Varieties of Atheism
  • Philo and Cleanthes Views on Natural Religion
  • A Theory of the A Priori: Philosophical Perspectives
  • The Ontological Argument for the Existence of God
  • Teachings of Jesus and Kant
  • Can God and Real Evil Be Reconciled?
  • Differences Between Confucianism and Daoism
  • Asian Philosophy: Concept of Samsara
  • Postmodern Plural Approach to Religion and Ratzinger’s "Dictatorship of Relativism"
  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2020, April 29). Rudolf Carnap's View on Cosmological Argument. https://ivypanda.com/essays/rudolf-carnaps-view-on-cosmological-argument/

"Rudolf Carnap's View on Cosmological Argument." IvyPanda , 29 Apr. 2020, ivypanda.com/essays/rudolf-carnaps-view-on-cosmological-argument/.

IvyPanda . (2020) 'Rudolf Carnap's View on Cosmological Argument'. 29 April.

IvyPanda . 2020. "Rudolf Carnap's View on Cosmological Argument." April 29, 2020. https://ivypanda.com/essays/rudolf-carnaps-view-on-cosmological-argument/.

1. IvyPanda . "Rudolf Carnap's View on Cosmological Argument." April 29, 2020. https://ivypanda.com/essays/rudolf-carnaps-view-on-cosmological-argument/.

Bibliography

IvyPanda . "Rudolf Carnap's View on Cosmological Argument." April 29, 2020. https://ivypanda.com/essays/rudolf-carnaps-view-on-cosmological-argument/.

“Aquinas’ Cosmological Argument proves that God exists.” Critically evaluate this statement.  [40] 

Cosmological Arguments start with the existence of the universe (Greek = Kosmos) and conclude that God is the most logical explanation of it.  They are some of the oldest arguments for God’s existence and have an intuitive appeal.  As Richard Swinburne observed in “Is There a God?” (1996)

“ The human quest for explanation inevitably and rightly seeks for the ultimate explanation of everything observable.”

Cosmological Arguments can be found in the work of Plato (Laws Book X) and Aristotle (Physics Book II, Metaphysics Book IV) and make up the first four of Aquinas’ five ways to God in the Summa Theologica (1.2.3).   While Aquinas’ Cosmological Arguments are all framed as a  posteriori  arguments  – and so could never provide  proof  – they do provide strong support for the existence of a Prime Mover.  Nevertheless, Aquinas goes too far in his claim that this is what everybody calls God.

Aquinas’ first way draws on the Aristotelian concept of movement. In the Physics, Book V, Aristotle wrote, “ all things that are in motion must be moved by something. ”  Motion does not necessarily mean movement in the sense that things are is moving through space from location A to location B, but rather that they are moving from a state of potentiality to a state of actuality in multiple different respects. As Aristotle wrote, movement involves the… “ actualizing of some potency. It is because things have real potencies that they are able to change.”  Aquinas later wrote, “ for motion is nothing else than the reduction of something from potentiality to actuality .”  Whereas Plato’s argument, later refined by Muslim scholars of the 8 th  Century Kalam School and more recently by William Lane Craig, focuses on a temporal series of causes much like a domino-rally, pointing towards a beginning in time, an uncaused cause, which is what everybody calls God, for Aristotle and for Aquinas, even if the universe is as infinite as it appeared to be, there is still the need for a Prime Mover because everything depends on other things.  As Parmenides, Heraclitus and separately the Buddha observed, everything changes or moves and nothing stays the same, but nothing changes or moves without being moved by something else, even if that is just time itself.  As Aristotle wrote, “ potential, precisely because it is potential, cannot make itself actual ”.  Aristotle concluded that there must be a Prime Mover outside time and space, but stops short of claiming that this is God.  Aquinas went further, claiming that this Prime Mover is “ what everybody calls God ”, but in doing this he weakened the argument. It is true that the Prime Mover must be outside time and space and thus wholly simple and unchanging, pure actuality and with zero potentiality.  As Aquinas wrote,

“nothing can be reduced from potentiality to actuality, except by something in a state of actuality… it is impossible that in the same respect and in the same way a thing should be both mover and moved…” 

It is also true that everything ultimately depends on the Prime Mover for its existence.  As Aquinas wrote, “therefore it is necessary to arrive at a first mover, put in motion by no other”  However, it is a step too far to claim that the Prime Mover could be seen as the “creator”, let alone act in the world, speak to human beings or be crucified and rise again. When Aquinas writes “and this everyone understands to be God”  He goes beyond the evidence and possible knowledge.  In his book “The Nature of God” Gerard J. Hughes describes the Prime Mover changing potentiality to actuality in terms of a bowl of milk causing a cat to cross a room.  The bowl of milk does nothing, in the way that the Prime Mover – being timelessly unchanging and impassive – does nothing, because it has no potential and is pure act. It follows that Aquinas’ first framing of the Cosmological Argument provides strong support for the existence of a Prime Mover, but not for the existence of the God that Christians worship, because the Prime Mover would be unable to say “ let there be light ”, work miracles or judge individuals on the final day… all of these require in God potential and the ability to act in time, which the Prime Mover cannot have.

Aquinas’ second way draws on the Aristotelian concept of efficient causation.  For Aristotle, all things have four causes – material, formal, efficient and final.  Efficient causes are agents which bring things into being, in the way that parents bring their children into being or the earth, sun and rain bring the oak tree out of the acorn.  If everything depends on efficient causes to bring them into being, again there is a chain of causation which requires explanation.  The chain cannot be infinite, because if there was no first efficient cause there would be no subsequent causes and the universe would not exist.  Something cannot come out of or be caused by nothing. Similarly, there cannot be a first efficient cause like other things in the universe, as if there were it would need efficient causes of its own and could not, therefore, be the first.  Aquinas concludes, “it is necessary to admit a first efficient cause, to which everyone gives the name of God.” Again, Aquinas’ second Cosmological Argument provides strong support for the existence of an uncaused efficient cause and again, it is a step too far to claim that this must be the God of Christian worship.   This not least because efficient causes do not need to be sustaining causes in esse (as Frederick Copleston later called them) but could be a cause in fieri (again, to use Copleston’s terminology).   An uncaused cause which began the universe but has no further role in it is not the God of Christian theism; at most it supports deism.  Further, Aquinas admits that

“ There is no case known (neither is it, indeed, possible) in which a thing is found to be the efficient cause of itself; for so it would be PRIOR to itself, which is impossible.”

  This implies that as efficient cause the uncaused cause must be PRIOR to the universe, something which would be difficult to reconcile with Big Bang Theory as this suggests that as time itself was created at the Big Bang, it makes no sense to speak of anything being PRIOR to it or indeed, as Stephen Hawking observed, causing it.  In these ways, Aquinas goes beyond the evidence in claiming that the uncaused cause is that to “ everyone gives the name of God ”.

Aquinas’ third way develops the idea of the contingency of things in the universe, pointing out that everthing has the potential to be or not to be; “We find in nature things that are possible to be and not to be”.  In an infinite universe, all potentials not to be might be expected to have been realized; as something can’t come out of nothing, nothing would then exist and I could not be writing this essay.  It follows, therefore, that EITHER the universe cannot be infinite – in which case there would have to be a first cause in time which would be what everybody calls God – or the universe is infinite and there exists a “ necessary being ”, a fully actual “ neither something nor nothing ” which contains its own explanation and has no potential not to exist.  This, Aquinas claims, is what “ all men speak of as God .”  Aquinas’ third Cosmological Argument is just as problematic as an argument for the existence of the God of Christian worship as the first and second.  Not only as an a posteriori argument does it stop short of providing proof, it also goes well beyond the observable evidence in concluding that the necessary being is God as Christians would define Him.  Leibniz later recast the third Cosmological Argument as an a priori  argument, writing:

“Why is there something rather than nothing? The sufficient reason […] is found in a substance which […] is a necessary being bearing the reason for its existence within itself.” 

For Leibniz, anything that exists has a cause for its existence, either in the necessity of its own nature or in an external cause.  If the universe has an external cause for its existence, this cause must be God.  As the universe exists, it must have a cause for its existence, either in the necessity of its own nature or in an external cause.  Because the universe exists contingently, not necessarily, the universe must have an external cause and this must be God.  Nevertheless, like Aquinas, Leibniz argument fails to prove the existence of the God of Christian worship.  Causing the universe is not enough to be called God; the Christian God does rather more than an abstract singularity or the Higgs Boson does.  Further, it doesn’t make sense to predicate much of what the Christian God does to the necessary cause of the universe supported by Aquinas and by Leibniz in his supposed improvement of Aquinas’ third cosmological argument.  Both the God of Aquinas and the God of Leibniz are timelessly impassive and it is inconceivable how such a being could act even once to create the world, given that this would involve a change in its being incompatible with being timeless and fully actual with no potential.  Further, as Immanuel Kant observed, we have no experience of necessary beings so it makes little sense to speculate about their possible existence.  Also, it is inconsistent to start an argument by claiming that all things are contingent and conclude by hypothesizing something that is not contingent.  Again, while Aquinas’ third Cosmological Argument strongly points towards the existence of a necessary being or beings in the universe, it is far from being conclusive proof of such, even when recast as an a priori argument and cannot in any case justifiably claim that the necessary being is what Christians worship as God.

While it is true that most of the classical criticisms of Aquinas presented by David Hume and Bertrand Russell fail to undermine his Cosmological Arguments, the point (which they all make) about the Cosmological Argument failing to support the God of Christian worship stands.

Firstly, Hume criticized a version of the Cosmological Argument presented by his character Demea in Dialogues Concerning Natural Religion (1779) Book IX.  Cleanthes points out that there is no support for the claim that everything in the universe is moves and is moved, caused and is caused or is contingent.  We have a limited view of the universe and no sensible reason to believe that the universe is homogenous or that we see things the way they actually are, a principle known in science as isotropy.  For all we know, argues Cleanthes, there could be unmoved movers, uncaused causes or necessary beings within the universe which could explain its continued existence.  Nevertheless, accepting these criticisms of the Cosmological Argument entails abandoning Natural Science altogether.  Leibniz coined the term “Cosmological Principle” to refer to the principles of homogeneity and isotropy which all scientists must assume in order to reason inductively towards natural laws.  Without the Cosmological Principle, we could not make many scientific knowledge claims; Cosmology and Quantum Science, Medicine and Biochemistry would all be a waste of time.  In practice, laws of nature supported by inductive reasoning enable mobile phones and space shuttles to work, so it doesn’t make sense to doubt the authority of our observations as Hume, through Cleanthes, does.  These criticisms of Hume’s fail to undermine Aquinas’ Cosmological Arguments insofar as they point to a necessary cause for the universe.

Secondly, Cleanthes continues by criticizing the claim that just because the parts of the universe have causes, so must the universe as a whole.  Demea (and Aquinas) rely on the so-called fallacy of composition.   Further, Cleanthes asks why the universe cannot be the explanation of itself, why there must be an external cause for the universe.  Later, Bertrand Russell asked why the universe cannot be a “brute fact”.  Yet neither of these criticisms is conclusive.  As Leibniz points out, it is difficult to see how a universe of contingencies can itself exist necessarily.  Contingencies involve potential which cannot, by definition, exist within a necessary being.  Further, while characteristics of the parts do not necessarily have to be characteristics of the whole and while (as Russell argued) just because all men have mothers it doesn’t mean that the human race has a mother, it can sometimes follow.  Each strand of spaghetti has two ends, something which also applies to the whole packet of spaghetti.  In a sense and because it is made up of material in the way that things in the universe are, the universe is a thing.  Things exist contingently and need to be moved and caused by things other than themselves.  These criticisms of Hume’s fail to undermine Aquinas’ Cosmological Arguments insofar as they point to a necessary cause for the universe as well.

Nevertheless, Cleanthes’ criticism that the cause of the universe could not be said to have the attributes of the Christian God is, for reasons previously explored, is persuasive .  Again it is clear that while Aquinas’ Cosmological Arguments do offer support to the hypothesis that there is an uncaused, necessary cause for the universe, they are far from proving that the God of Christian worship exists. 

In conclusion, Aquinas’ Cosmological fail to prove that God exists, both because as a posteriori arguments they stop short of proving their conclusions and because even if they are reframed into a format which could provide proof, as Leibniz attempted, they demonstrate only the existence of an abstract necessary being far short of having the attributes of the God Christians worship. 

Bibliography

  • Class notes on the Cosmological Argument
  • Aristotle, Physics Book V
  • Aquinas, Summa Theologica 1, 2, 3
  • Hume’s Dialogues Concerning Natural Religion Book IX
  • Vardy & Vardy “God Matters” Chapters 4 & 5

Share this:

Leave a comment cancel reply.

' src=

  • Already have a WordPress.com account? Log in now.
  • Subscribe Subscribed
  • Copy shortlink
  • Report this content
  • View post in Reader
  • Manage subscriptions
  • Collapse this bar

cosmological argument essay conclusion

Skip to content

Get Revising

Join get revising, already a member.

Ai Tutor Bot Advert

Cosmological arguments

Teacher recommended

  • Created by: Ellie Charlish
  • Created on: 22-02-16 20:13
  • Basis for arguments is that the universe cannot account for its own existence.
  • Why do things exist at all - why is there something, rather than nothing.
  • Plato argued that everything must have been created by some cause. Aristotle argued that behind the series of cause and effect in the world there must be an unmoved mover. 

Aquinas and the cosmological argument in five ways

  • Ways to describe God through inductive argument, based on observation and evidence. 
  • A key goal for Aquinas was to show how faith and reason could work together. He was a well known man in the Church working as an advisor to the Pope 
  • One through revelation - God choses to the truth to people, e.g. through the Bible. 
  • Other is through human reason, he thought that if we applied reason to the evidence around us, we can reach valuable truths. 
  • Aquinas presented five ways in showing that God exists., because he was convinced that although the existence of God was not self evident, it could be demonstrated with logical thought.
  • The universe exists
  • There must be a reason why
  • However, some people such as Bertrand Russel and Richard Dawkins, are happy to accept the universe as just is, without moving to the conclusion that there should be some reason for it. 

First way - The Unmoved Mover

  • Aquinas concentrated on the existence of change, or motion, in the world. 
  • Something must of been set in motion by something
  • Aquinas thought that this sequence must have been an unmoved mover to set the whole thing off.
  • If change is brought around about in thing A by thing B then B must have the characteristics of the change it brings about. 
  • Agument was on dependency, rather than going back in time until a beginning was found; he was using the idea that God sustains the universe, and trying to show that we could not have  a universe of change, vitality and motion without a first mover. 

Second way - The Uncaused Mover

  • The argument is similar, replaces the idea with change and motion with the concept of cause 
  • Every 'effect' has a cause, argues infinite regress is impossible. therefore there must be a First cause 'which we call God'
  • He focuses on the 'efficient cause' - took up Aristotle's ideas and understanding of causes to argue that things do not cause themselves in this way - they cannot be thier own agents. Therefore he said, there must be a first efficient cause aka God. 

Third Way - Contingency

  • The world consists of contingent beings, which are beings that begin and end, and which are dependent on something else for their existence. 
  • Everything in the physical world is contingent, depending on external factors for its existence. 
  • Things are contingent in two ways: they…
  • Cosmological Argument

Report Tue 15th March, 2016 @ 18:28

10/10 best resource there is

Report Wed 10th May, 2017 @ 13:03

AMAZING THANK YOU SO MUCH

Report Fri 8th September, 2017 @ 17:20

helpful thankyouu

Report Thu 23rd November, 2017 @ 16:07

very helpful, thanks very much!!!

cosmological argument essay conclusion

Report Tue 13th December, 2022 @ 11:45

Similar Philosophy resources:

The Cosmological Argument 1.0 / 5 based on 1 rating

Cosmological Argument 0.0 / 5

Quotes 0.0 / 5

Philosophers arguing for God and opposition 4.5 / 5 based on 3 ratings

Review of the cosmological argument 4.0 / 5 based on 6 ratings

Cosmological argument 0.0 / 5

The Cosmological Argument 3.0 / 5 based on 4 ratings

The Cosmological Argument 1.5 / 5 based on 8 ratings

The cosmological argument Introduction 0.0 / 5

cosmological argument essay conclusion

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

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. 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 ...

  3. 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 ...

  4. 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 ...

  5. Cosmological argument

    A cosmological argument, in natural theology and the philosophy of religion, is an argument which asserts that the existence of God can be inferred from facts concerning causation, ... The conclusion of these arguments is that there exists a first cause, subsequently analysed to be God.

  6. 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. 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 scholars. Key points are summarized toward the end.

  7. Cosmological Argument: Everything Has a Cause Essay

    Exclusively available on IvyPanda®. The cosmological argument refers to a theoretical squabble for the subsistence of God which elucidates that everything has a cause. The cosmological argument attempts to prove that God exists by stating that there ought to be an ultimate origin of all things. This can be elaborated further by a simple ...

  8. 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.

  9. 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.

  10. 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 ...

  11. Essay Plan: Cosmological Argument

    The Cosmological argument fits in with the God of classical theism (omnipotent, omnibenevolent, omniscient). It makes sense to think that there is an initial cause to the universe: this fits with our experience of events within the universe. Most scientists would argue that the universe has a beginning, which fits in with the Cosmological argument.

  12. 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. Get a custom essay on The Cosmological Argument.

  13. An Analysis of Aquinas' Cosmological Argument

    The Cosmological argument is 'a posteriori' - it is reliant upon and fits with our experience of the world around us, with our own experience of causal chains provoking questions over how we, and the universe as a whole, came to be. The aim of the Cosmological argument is to attempt to prove God's existence by showing that an infinite regress of causal chains is logically impossible ...

  14. Rudolf Carnap's View on Cosmological Argument Essay

    The Cosmological Argument has two parts. The first part attempts to prove the existence of a being that is not like other beings, while the second part of the argument proves that this special being has some features, such as perfect goodness, omniscience, and omnipotence. Moreover, this being is the creator of the world, and it is separate and ...

  15. "Aquinas' Cosmological Argument proves that God exists." Critically

    The Cosmological Arguments, rooted in the existence of the universe, posit that God is the most logical explanation. ... as something can't come out of nothing, nothing would then exist and I could not be writing this essay. It follows, ... In conclusion, Aquinas' Cosmological fail to prove that God exists, both because as a posteriori ...

  16. Cosmological Argument For The Existence Of God

    Aquinas cosmological argument for the existence of God is known to be the most popular by philosopher and religious scholars. In his theological masterpiece, Summa Theologia, he proposed varies forms of cosmological arguments to explain "ways" that he thought would prove God exists. Many skeptics do not accept the existence of God because ...

  17. Samuel Clarkes Cosmological Argument

    Philosopher Samuel Clarke put forth a modern formulation of the cosmological argument taking a slightly different path than Aquinas's famous cosmological argument. But like Aquinas, Clarke adopts the premise that all beings that we encounter must have causes. Contrary to Aquinas, Clarke differentiates between contingent and necessary beings.

  18. Cosmological arguments

    Cosmological arguments. Basis for arguments is that the universe cannot account for its own existence. Why do things exist at all - why is there something, rather than nothing. Plato argued that everything must have been created by some cause. Aristotle argued that behind the series of cause and effect in the world there must be an unmoved ...

  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. Evaluation of the Kalam Cosmological Argument

    The Kalam Cosmological Argument is one of the variants of the cosmological argument and had been used to defend the philosophical position of theistic worldviews (AllAboutPhilosophy.org,2018). II. The Kalam Cosmological Argument. The Kalam Cosmological argument is an argument from the existence of the world or universe to the existence of God ...

  21. Ontological And Cosmological Arguments The Existence Of God ...

    God is the most perfect and flawless being, hence, he must exist. Similar to the ontological argument, the cosmological argument, also known as the first cause argument, is a classical argument for the existence of God. However, unlike the ontological argument, it derives the conclusion that God exists from a posterior premise (with evidence ...

  22. Cosmological Argument

    Cosmological Argument. 1. The three views of the relationship between faith and reason are inevitable, independent, and overlapping. Inevitable is conflict between science and religion one of them has to be right but they cannot both be right. Independent with no overlap means that faith and reason are separate.

  23. 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 ...