The Design Argument for the Existence of God

by James R. Beebe

Dept. of Philosophy

University at Buffalo

Copyright ã 2002

Outline of Essay :

I. The Analogical Version of the Design Argument

II. Criticisms of the Analogical Version

A. No Experience of Cosmic Beginnings: A Disanalogy

B. Evil: A Harmful Analogy

C. Full-Blown Anthropomorphism

D. Begging the Question

III. The Inference to the Best Explanation Version of the Design Argument

A. Inference to the Best Explanation

B. The Data

C. Explaining the Data

D. Conditional Probability

IV. Objections and Replies

A. The Anthropic Principle

B. “Sometimes the Improbable Happens”

C. Non-deductive Inference

            The design argument is the simplest, most straightforward argument for the existence of God.  Unlike the cosmological argument, the design argument can be stated in a few, easy-to-understand steps.  In a nutshell, the design argument claims that the fact that everything in nature seems to be put together in just the right manner suggests that an intelligent designer was responsible for its creation.  Immanuel Kant (1724-1804)—a strident critic of the design argument—recognized both its simplicity and its importance.  He wrote, "This proof always deserves to be mentioned with respect.  It is the oldest, the clearest, and the most accordant with the common reason of mankind" (Kant 1781/1965, A 623, B 651).

            In the first section of this essay I will describe the most famous version of the design argument—William Paley’s argument by analogy.  Analogical arguments are perhaps the weakest sort of arguments one can offer without committing an outright fallacy.  As we will see in section II, the analogical version of the design argument has come in for some heavy fire over the years.  A contemporary reformulation of the argument, which I will call the ‘Inference to the Best Explanation’ (IBE) version of the design argument, claims to be able to escape the criticisms that are leveled against the analogical version.  The IBE version will be explained in section III.  It eschews the analogical form of the first version and uses evidence from contemporary science to back up its claims.

            William Paley (1743-1805), an Anglican priest whose textbooks were required reading at Cambridge until the twentieth-century, put forward the most famous version of the design argument in his book Natural Theology: or Evidences of the Existence and Attributes of the Deity Collected from the Appearances of Nature .  In his autobiography, Charles Darwin (1876/1958, p. 19) cites Paley’s book as one of his favorite undergraduate texts:

In order to pass the B.A. examination, it was also necessary to get up Paley’s Evidences of Christianity , and his Moral Philosophy .  This was done in a thorough manner, and I am convinced that I could have written out the whole of the Evidences with perfect correctness, but not of course in the clear language of Paley.  The logic of this book and, as I may add, of his Natural Theology , gave me as much delight as did Euclid.  The careful study of these works, without attempting to learn any part by rote, was the only part of the academical course which, as I then felt, and as I still believe, was of the least use to me in the education of my mind.  I did not at that time trouble myself about Paley’s premises; and taking these on trust, I was charmed and convinced by the long line of argumentation.

The only discussion of the design argument that might be more famous than Paley’s is David Hume’s (1711-1776) in his Dialogues Concerning Natural Religion .  In this work Hume subjects the argument to severe criticism. 

            Paley famously begins his version of the argument by comparing the universe to a watch.  Suppose, he says, that we come upon a watch while walking through the forest.

[W]hen we come to inspect the watch, we perceive... that its several parts are framed and put together for a purpose, e.g. that they are so formed and adjusted as to produce motion, and that motion so regulated as to point out the hour of the day; that, if the different parts had been differently shaped from what they are, if a different size from what they are, or placed after any other manner, or in any other order than that in which they are placed, either no motion at all would have been carried on in the machine, or none which would have answered the use that is now served by it....  This mechanism being observed,... the inference, we think, is inevitable, that the watch must have had a maker; that there must have existed, at some time, and at some place or other, an artificer or artificers who formed it for the purpose which we find it actually to answer; who comprehended its construction, and designed its use.  (cited in Hick 1964, pp. 99-100)

Paley claims that the same can be said for the universe as a whole.  It seems to show evidence of an intelligent designer as well.  The parts of the universe have an order, complexity and simplicity that resemble the parts of a finely crafted, well-oiled machine.  It seems, then, that the universe was fashioned by some kind of Divine Watchmaker.

            To support the analogy between a finely crafted watch and the universe, defenders of the design argument typically put forward the following kinds of considerations.  Consider the fact that the universe is constructed in a way that is conducive to life.  There is just enough oxygen to support life on earth.  If there were even a little less, the Earth’s atmosphere would not be able to support life as we know it.  But if there were just a little bit more oxygen in the atmosphere, combustion would occur too easily and often and it would once again be difficult to sustain life in such conditions.  Moreover, the Earth is just the right distance from the sun.  If we were a little bit closer, the atmosphere would be too hot to sustain life; but if we were a little further away, plants would not receive enough energy from the sun to carry on photosynthesis—the primary process by which the sun’s energy is converted into life on Earth.  These and other “fine-tuning” aspects of the universe suggest that there was an intelligent mind that intentionally brought these features into being.  As Sir Isaac Newton put it, the “most beautiful system of the sun, planets and comets could only proceed from the counsel and dominion of an intelligent and powerful Being” (from the General Scholium of Principia Mathematica ; cited in Leslie 1989, p. 25). 

            Defenders of the design argument need not rest content with pointing to large-scale features of the universe that suggest design.  They can also point to the apparent design of many kinds of objects in the world.  Take, for example, mammalian organs, such as the heart, kidney, brain or eye.  Each of these has been given a certain function to perform and each has an amazing capacity to carry out that function.

            We can summarize the analogical version of the design argument as follows: 

1) Human artifacts are the products of intelligent design.

2) The universe resembles human artifacts.

3) Therefore, the universe is probably a product of intelligent design.

4) Therefore, the author of the universe is probably an intelligent being.  (adapted from Plantinga 1990, p. 97)

We know that (1) is true on the basis of personal experience and testimony from reliable sources.  We know that machines are always put together to serve certain purposes and that it takes careful planning and construction to make sure that each of the parts of a complicated machine work properly.  So, when we see that mammalian organs—e.g., hearts, lungs, eyes, brains, etc.—have certain very specific functions that they perform by carrying rather complicated series of interactions, it is plausible to think that they, too, are the products of intelligent design. 

            The most famous criticisms of the analogical version of the design argument appear in David Hume’s Dialogues Concerning Natural Religion .  According to Norman Kemp Smith (“Introduction” to Hume’s Dialogues, p. 3), “Hume’s destructive criticism of the argument was final and complete.”  Smith’s sentiments are shared by many.  As we will see, however, defenders of the IBE version of the design argument claim that Hume’s criticisms apply only to the analogical version of the argument and not to their own version.

            Most of Hume’s criticisms center on the analogical aspect of the traditional design argument.  The problem, according to Hume, is that the analogy in question is not as strong as it needs to be in order to succeed.  The more the universe resembles an artifact, the stronger the argument will be; but the less the universe is like an artifact, the weaker it will be. 

  A. No Experience of Cosmic Beginnings: A Disanalogy

            Hume points out there are many dissimilarities or disanologies that harm the theist’s case.  One of the first things he notes is that there is a disanalogy between the kind of experience we have with respect to artifacts and the kind of experience we have with respect to the universe.  We have a lot of experience with a wide range of artifacts, which includes a general idea about what kinds of artisans or craftspeople make artifacts.  We know that artifacts are made by intelligent designers because we have observed designers making a variety of things on many occasions.  However, we don’t know what usually makes universes.  We simply have no experience with this kind of thing.  As a result, Hume claims, we can’t be too confident that whatever was responsible for making the universe is going to be much like the designers we are familiar with. 

  B. Evil: A Harmful Analogy

            Hume also argues that there are analogies that are detrimental to belief in a designer.  He thinks the analogical design argument correctly notes that we generally infer properties about an artisan or manufacturer from properties we observe in their products.  But he claims the argument ignores important facts about our world.  For example, from the solid 24 karat gold, diamonds and precision timing of a Rolex watch, one can infer that the manufacturer has the highest commitment to quality.  When one is faced with a defective product, one draws analogous but opposite conclusions.  For example, I own a Soviet-era military watch with the KGB insignia on its face.  At noon and midnight, the two hands of the watch should both be pointing straight up, but there are five degrees of separation between them.  Moreover, it gains about eight minutes every day.  I have been led to form rather negative conclusions about Soviet-era craftsmanship from the properties of this watch. 

            Hume notes that there seem to be imperfections in nature: cancer, AIDS, heart disease, famines, plagues, floods, and countless other tragedies.  If God is a Divine Watchmaker, as Paley claimed, the world looks to be more like a Soviet watch than a Rolex.  In other words, if we are going to infer by analogy characteristics of the Creator from characteristics of the creation, it doesn’t seem we can conclude that the Creator had to be perfect because the world is anything but perfect.  In jest, Hume suggests that maybe the universe was created by a junior deity who is just learning the ropes of universe creation and didn’t get things quite right this time.

  C. Full-Blown Anthropomorphism

            Hume also asks, “While we’re in the business of arguing by analogy, what is to keep us from pursuing the analogy all the way to a full-blown anthropomorphism?”  The theist wants to argue that, since every highly ordered, complex contrivance we encounter has an intelligent designer behind it, we can conclude that the world also has a designer behind it.  Well, says Hume, every artifact we encounter also has a designer with toenails, a bellybutton, 46 chromosomes, teeth made out of calcium composites, a spleen, and bad breath in the morning.  What’s stopping us from concluding that the creator of the universe has all of these features as well?  Hume’s point is that the analogy upon which the traditional design argument is based supports other conclusions than the one the theist is seeking to support. 

  D. Begging the Question

            Kelly James Clark (1990, p. 30) echoes some of Hume’s worries when he claims,

The connection between the first premise, that the world appears designed, and belief in God is so tight that some contend that the argument from design simply assumes what it is trying to prove: to countenance apparent cosmic design is already to be committed to a design -er. 

Clark thinks that one must already believe in God before one can accept the first premise of the traditional design argument—viz., that the world appears to be designed.  If you don’t think there is a cosmic designer, then you’re probably not going to look at the world and think “This world appears to be designed.”  An argument that assumes from the start the very thing that is up for debate is said to ‘beg the question.’  Since Clark thinks the analogical design argument begs the question, he concludes that the argument fails to have any persuasive force.

            While I think Hume’s criticisms for the most part succeed in hitting their mark, I think that Clark’s does not.  Most (if not all) evolutionary theorists admit that biological organs appear to be designed, but they remain committed to the project of explaining apparent design without appealing to a designer.  In other words, they do not deny the first premise that Clark finds so worrisome; they merely deny the inferences theists wish to draw from it. 

            In any case, the objections put forward in this section have convinced many people that the design argument fails miserably.  However, in recent years there has been a renewal of interest in the design argument—from both philosophers and scientists—and this has led to an updated formulation of the argument.

            The design argument can be reformulated so that it is not an analogical argument.  Instead, it can be understood as an inference to the best explanation.  This form of inference is common to both science and ordinary life.  We start with a set of data that is initially surprising or unexplained, and we wonder what could explain it.  As possible explanations pop into our minds, we evaluate the initial plausibility and simplicity of each explanation and see how well the explanations make sense of the data.  J. P. Moreland (1994, p. 26) offers the following example of an ordinary inference to the best explanation.

Suppose I get a terrible stomachache.  Then it dawns on me that I just ate a gallon and a half of ice cream, two bags of popcorn and a lot of candy on an empty stomach.  A hypothesis suggests itself as the best explanation of the stomachache—it arose because of what I had just eaten.  Other hypotheses may also suggest themselves, but I should adopt the explanation that best solves the problems for which it was postulated. 

In what follows I will present an array of scientific data that, according to defenders of the IBE design argument, cries out for explanation.  The facts described are extraordinarily improbable and unlikely to be the result of chance.  After presenting the data, we will examine the suggestion that the best explanation for these unlikely occurrences is that a personal, transcendent Being of tremendous power and intelligence is directly responsible for purposefully bringing them about. 

1.         We can begin by considering the fact that the energy the Earth receives from the Sun is precisely the amount required to nurture life.  According to Richard Brennan (1997, pp. 244-245),

The term used in science for this energy is the solar constant , which is defined as 1.99 calories of energy per minute per square centimeter.  If Earth received much more or less than 2 calories per minute per square centimeter, the water of the oceans would be vapor or ice, leaving the planet with no liquid water or reasonable substitute in which life could evolve.  It is only because Earth is 93 million miles away from a Sun that produces 5,600 million, million, million, million calories per minute that life is possible.

2.         Brennan (1997, p. 245) continues,

For another example, it has been calculated that if Earth were just 5 million miles closer to the Sun, the intensity of the Sun’s rays would have broken apart water molecules in the atmosphere and eventually turned the planet into a dry and dusty wasteland.  If Earth were only 1 million miles farther from the Sun, the cold would have frozen the ocean solid. 

3.         For there to be enough carbon around to support life, the strong nuclear force (the force that holds quarks together to form protons and neutrons and holds protons and neutrons together to form the nuclei of atoms) can be no more than 1% stronger or weaker than it is.  Increasing its strength by 2% would block the formation of protons, so that there would either be no atoms at all or else stars would burn a billion billion times faster than our sun, thereby making it difficult to have an environment friendly to living organisms (Leslie 1989, p. 4).  Increasing its strength by only 1% would result in all carbon being burned into oxygen (Leslie 1989, p. 35). 

4.         Decreasing the strong nuclear force by 5% would make it impossible for stars to burn (Leslie 1989, p.4).

5.         If the force of electromagnetism were somewhat stronger, the amount of light given off by stars would be significantly lower.  Main sequence stars (i.e., stars like our sun in the stable phase during which they spend most of their lifetimes and have their interior heat and radiation provided by nuclear fusion reactions near their centers) would be too cold to support life and would not contain any elements heavier than iron.  It would also make protons repel one another strongly enough to prevent the existence of atoms (Leslie 1989, p. 4).

6.         If the force of electromagnetism were slightly weaker, all main sequence stars would be very hot and short-lived blue stars.  According to the physicist, P. C. W. Davies, changes in either electromagnetism or gravity by only one part in 10 40 would spell catastrophe for stars like the sun (Leslie 1989, p. 37).

7.         Some of the basic forces of the universe also need to be finely-tuned to each other.  Gravity is roughly 10 39 times weaker than electromagnetism.  If it had been only 10 33 times weaker, stars would be a billion times less massive and would burn a million times faster (Leslie 1989, p. 5).  If gravity were ten times less strong, stars and planets could probably not form at all (Leslie 1989, p. 39).

8.         The opposite charges of electrons and protons perfectly balance each other.  They are identical magnitudes.  If there had been a difference between their charges even as small as one part in ten billion, scientists have calculated that no solid bodies could weigh more than one gram (Leslie 1989, p. 45).

9.         The difference in mass between protons and neutrons is twice the mass of the electron, which is itself a very small quantity.  If this were not so, then

all neutrons would have decayed into protons or else all protons would have changed irreversibly into neutrons.  Either way, there would not be the couple of hundred stable types of atom on which chemistry and biology are based. (Leslie 1989, p. 5) 

If all protons were changed irreversibly into neutrons, the universe would consist of nothing but neutron stars and black holes.  (A neutron star is a kind of collapsed star that is immensely dense and is made mostly of neutrons.  It is not the sort of star that could support life as we know it.)

10.       According to William Lane Craig (Strobel 2000, p. 77), P. C. W. Davies concluded that the odds against the initial conditions being suitable for the formation of stars is a one followed by at least a thousand billion billion zeroes.

11.       Davies also estimated that if the strength of gravity or of the weak force were changed by only one part in a ten followed by a hundred zeroes, life could never have developed (ibid.).

12.       Craig (1990, p. 143) writes,

[Astronomer Fred] Hoyle and his colleague Wickramasinghe calculated the odds of the random formation of a single enzyme from amino acids anywhere on the earth’s surface as one in 10 20 .  But that is only the beginning: “The trouble is that there are about two thousand enzymes, and the chance of obtaining them all in a random trial is only one part in (10 20 ) 20,000 = 10 40,000 , an outrageously small probability that could not be faced even if the whole universe consisted of organic soup.”  And of course, the formation of enzymes is but one step in the formation of life.  “Nothing has been said of the origin of DNA itself, nothing of DNA transcription to RNA, nothing of the origin of the program whereby cells organize themselves, nothing of mitosis and meiosis.  These issues are too complex to set numbers to.”  In the end, they conclude that the chances of life originating by random ordering of organic molecules is not sensibly different from zero.

13.       J. P. Moreland (1987, p. 53) claims, “If the mass of a proton were increased by 0.2 percent, hydrogen would be unstable and life would not have formed.” 

14.       Brennan (1997, p. 246) writes,

If something called the fine structure constant (the square of the charge of the electron divided by the speed of light multiplied by Planck’s constant) were slightly different, atoms would not exist.

15.       The fact that all of this fine tuning is distributed across enormous ranges makes it even more amazing that they should be found in just the right proportions.  The strong nuclear force is roughly 100 times stronger than electromagnetism.  Electromagnetism is itself some 10,000 billion billion billion times stronger than gravity (Leslie 1989, p. 6).

            None of the foregoing evidence of the “fine-tuning” of the universe depends upon acceptance of the Big Bang theory of the origin of (the present state of) the universe and the cosmic timeline (spanning 15 billion years) that goes with it.  Theists who think that the Big Bang is identical to the event of divine creation, however, can avail themselves of further evidence of the fine-tuning of the universe, some of which is described below.  According to defenders of the IBE design argument, this evidence shows that even the theory of cosmic origin most widely accepted by atheist scientists strongly suggests that there was and is an Intelligent Designer behind the controls of the universe. 

16.       The rate of expansion of the universe immediately after the Big Bang had to be finely tuned.  According to William Lane Craig (Strobel 2000, p. 77), Stephen Hawking, the world’s most famous living physicist, has calculated that if the rate of the universe’s expansion one second after the Big Bang had been smaller by even one part in a hundred thousand million million, the universe would have collapsed into a fireball.

17.       If the rate of expansion were decreased by only one part in a million when the Big Bang was a second old, the universe would have recollapsed before temperatures fell below 10,000 degrees (i.e., before it could cool off enough for life to be able to form) (Leslie 1989, p. 29).

18.       An increase of only one part in a million in the rate of the early universe’s expansion would have meant that the kinetic energy of expansion would have so dominated gravity that stars could not form (Leslie 1989, p. 29). 

19.       Had the weak nuclear force been slightly stronger, the Big Bang would have burned all hydrogen to helium.  There would then be neither water nor long-lived stable stars, which are hydrogen-burning (Leslie 1989, p. 4). 

20.       The weakness of the weak force results in our sun burning its hydrogen slowly and gently for billions of years instead of blowing up like a bomb (Leslie 1989, p. 34). 

21.       Making the weak nuclear force slightly weaker would have destroyed all of the hydrogen, and the neutrons formed during the earliest stages of the universe would not have decayed into protons (Leslie 1989, p. 4).  Without this neutron-decay, the universe would be made up of nothing but helium (Leslie 1989, p. 34). 

22.       P. C. W. Davies ( Other Worlds , London, 1980, pp. 168-169; cited in Leslie 1989, p. 28) claims that, because of all the parameters that had to be perfectly set before the Big Bang, the odds against a universe filled with stars is “one followed by a thousand billion billion zeros, at least.” 

            What should we make of all these facts?  John Leslie (1989, p. 25) responds, “Our universe does seem remarkably tuned to Life’s needs.”  Craig (1990, p. 143) writes,

The point is that within the wide range of universes permitted by the actual laws of physics, scarcely any are life-permitting, and those that are require incredible fine-tuning of the physical constants and quantities.  In fact, Donald Page of Princeton’s Institute for Advanced Study has calculated the odds against the formation of our universe as one out of 10,000,000,000 124 , a number that exceeds all imagination.

To get a handle on how large this number is, consider the fact that there are estimated to be only 10 80 elementary particles in the universe (Craig 1990, p. 159).  A universe that is inhospitable to life is extraordinarily more likely to have arisen than the one that we, in fact, find ourselves in.  Craig (1990, p. 143) claims that the fine-tuning of the universe “cries out for explanation.”  And an explanation immediately suggests itself: maybe this improbable “cosmic accident” wasn’t an accident after all.  Keith Parsons (1990, p. 181), who is quite skeptical of the design argument, summarizes the conclusion of the argument nicely as follows.

[A] “finely tuned” universe is much more likely if there is a God than if there is not.  In other words, it is implied that the cosmic “coincidences” that make possible a universe such as ours are extremely improbable unless they are the product of conscious design.  Presumably, the conclusion is that since a “finely tuned” universe does in fact exist, its existence strongly confirms the existence of a conscious Designer—that is, God. 

In other words, scientific discoveries of the infinitesimally small margin of error allowed in creating a universe capable of sustaining life support the central claim of theism: the universe was purposefully constructed by a personal, transcendent Being of tremendous power and intelligence. 

  D. Conditional Probability

            Let me introduce some ideas from probability theory that can make clearer how the IBE version of the design argument is supposed to work. 

1. Let ‘P(A ½ B)’ mean “the probability of A, given B.” 

2. Let A = “You will die of cancer in the next ten years.”

3. Let B = “You are 20 years old, do not smoke, have no family history of cancer, and are very healthy.”

The value of ‘P(A ½ B)’ is called a ‘ conditional probability ’ value because we are asking what the probability is that A is true, on the condition that B is true.  We are not simply asking what the probability of A is.  We are asking about A’s likelihood in light of certain background assumptions. 

            According to the stipulations above, P(A ½ B) is the probability that you will die of cancer in the next ten years, given that you are 20 years old, do not smoke, have no family history of cancer, and are very healthy.  The probability of that happening should be very low.  Let’s replace B with the following conditions and see how the resulting probability values differ.

4. Let C = “You are a chain-smoking, 55-year old male.” 

5. Let D = “You are a chain-smoking, 55-year old male who has been working in an asbestos factory for 35 years.”

Consider the value of P(A ½ C).  It will obviously be a lot higher than P(A ½ B).  And it’s a good bet that P(A ½ D) will be even higher.

            In each of these cases, A remained the same.  The only thing that changed was the set of background assumptions we used to determine the conditional probability value in question.  We are now in a position to use the idea of conditional probability to achieve a better understanding of the IBE version of the design argument.

6. Let F = “There exists a finely-tuned, life-permitting universe.”

7. Let T = “There exists an all-powerful, all-knowing, perfectly good God who created the universe.”

8. Let Not-T = “There does not exist an all-powerful, all-knowing, perfectly good God who created the universe.”

Now consider the following probabilities.

9. P(F ½ T)

10. P(F ½ Not-T)

According to the scientists cited above, a conservative estimate of P(F ½ Not-T) is 1/10,000,000,000 124 .  In other words, it is extraordinarily unlikely that the fine-tuning of the universe could have been brought about without the conscious planning of an intelligent designer. 

            Now think about the value of P(F ½ T).  You can make an estimate of this probability, regardless of whether you are a theist or an atheist (or neither).  I am simply asking what you think the probability of there being a a finely-tuned, life-permitting universe would be IF there were an all-powerful, all-knowing, perfectly good God who created the universe.  Although I can’t give a precise number, it seems that the probability of P(F ½ T) would be extremely high.  If there were a supremely powerful and intelligent God, that being could easily create a finely-tuned universe if he so desired.  So, the difference between P(F ½ T) and P(F ½ Not-T) is enormous. 

            Why is this fact significant?  Parsons (1990, pp. 193-194) writes,

It is the consequence of Bayes’ theorem [a theorem of probability theory that undergirds the currently accepted view about the confirmation of scientific theories]… that a given piece of evidence e confirms a hypothesis h if and only if e is more probable on h than on not- h .  Hence, where h is theism and not- h is atheism and e comprises all of the “finely tuned” features of the universe, the “finely tuned” features of the universe confirm theism if and only if those features are more likely if God exists than if God does not exist.

In other words, since P(F ½ T) is tremendously higher than P(F ½ Not-T), facts about the fine-tuning the universe provide confirmation of the existence of God.  The defender of the IBE version of the design argument concludes that it is more reasonable to believe that the universe was created by an Intelligent Designer than to believe that it spontaneously arose through chance. 

            The Inference to the Best Explanation version of the design argument sometimes encounters the following objection.

We should not be surprised that the universe is life-permitting.  If it weren’t life-permitting, we wouldn’t be here to contemplate it.  The fact that we are indeed here to contemplate it shows that it obviously must be life-permitting.  Consequently, expressions of surprise at the fact that our universe is well suited for living things are inappropriate. 

Part of this objection is obviously true, but another part of it is mistaken.  The trivially true part is the claim that if our universe were not life-permitting, we living human beings would not be around to contemplate it.  But it is a mistake to think that this fact neutralizes the need to explain why the universe is life-permitting. 

            Let me use the following example to make the point.  Suppose I were brought before a firing squad made up of one hundred professional marksmen and that each of them was instructed to shoot one dozen rounds of ammunition at me.  Now suppose that, after the smoke clears, it becomes evident that all 1200 bullets fired at me have missed their intended target.  After a brief moment of elation, I will begin to wonder why I am still alive.  Suppose I said to myself, “If they hadn’t all missed me then I shouldn’t be contemplating the matter so I mustn’t be surprised that they missed” (Leslie 1989, p. 108).  Would that thought thoroughly satisfy my curiosity?  Not by a long shot [sic].  I would begin to wonder whether they really intended to harm me.  Were they instructed to miss me on purpose?  Did someone load all of their rifles with blanks?  Was this just a cruel birthday joke perpetrated by my wife?  I might start looking around for the cameras from Spy TV or Candid Camera. 

            Similarly, merely pointing out that if the universe were not life-permitting, we would not be around to contemplate it does not satisfy our curiosity about the fine-tuning of the universe.  We can still ask for an explanation of why these amazing and unlikely facts came to be. 

  B. “Sometimes the Improbable Happens”

            A second objection that is often raised against the Inference to the Best Explanation version of the design argument goes like this:

Sometimes the improbable happens.  For example, the fact that it is extraordinarily unlikely that any single person will win the Powerball lottery does not mean that no one will ever win.  In fact, people whose odds of winning are vanishingly small win the Powerball lottery on a regular basis.  Our reaction to the existence of an improbable, “finely-tuned,” life-permitting universe should be the same as our reaction to the news that somebody won the latest Powerball lottery: an uninterested yawn.

The defender of the IBE design argument will claim: a) that there is a confusion lurking behind these remarks; and b) once we clear up the confusion we will see that there are important disanalogies between the Powerball case and the case of a fine-tuned universe. 

            Suppose that in a certain lottery there are 100 million tickets sold and that one of these tickets will be chosen at random.  If it is a fair lottery, then every ticket has an equal chance of winning.  So, the probability that any particular ticket will bring riches to its bearer is 1/100,000,000.  Now consider the probability that at least one of the 100 million lottery tickets that were sold will win.  That probability is 1 (probabilities come in ranges of continuous values between zero and one).  In other words, there is a 100% chance that one of the 100 million tickets sold will win. 

            According to the defender of the IBE design argument, we need to distinguish between the following two kinds of probability judgments:

1) The probability that a particular ticket will win.

2) The probability that some (i.e., at least one) ticket will win.

The value of (1) is 1/100,000,000.  The value of (2) is 1.  The reason we are unsurprised that somebody (or other) won the latest Powerball lottery is that the probability of somebody (or other) winning is 1.  It’s a sure bet.  But that doesn’t mean that we would not not be surprised if we held the winning ticket.  We be very surprised because of the enormous odds against our winning. 

            Our winning, however, would not be completely mysterious to us.  It’s not as if we would have no idea about how to explain how we won.  Our knowledge of how lotteries work includes the knowledge that somebody has to win.  We also know that winners in a fair lottery are selected through some kind of random process that gives everybody a fair shot.  Knowledge of this process—even if it is vague and unspecific—keeps the fact of our winning from being an utterly mysterious, unexplainable fact.

            The defender of the IBE design argument will maintain that the central problem with the current objection is that we do not know that the following is true:

4) The initial conditions of the present universe—e.g., the strengths of the four fundamental forces (the strong and weak nuclear forces, electromagnetism, and gravity), the masses of the fundamental particles, etc.—were the result of some kind of cosmic lottery.  Out of the indefinitely large number of possible universes that could have been brought into being, ours was the one that just so happened—by pure chance—to be selected.  If other universes had been selected, they would have collapsed into fireballs just a few seconds after being formed, while others would have been composed of only neutron stars and black holes.  Still others would have consisted of nothing but electromagnetic radiation.  Fortunately for us, none of these cosmic options were selected.

If we knew: a) that each possible universe had an equal probability of being actualized; b) that the probability that any particular universe would be selected was extremely small; and c) that at least one of them had to be selected; then it seems that we should show the same lack of surprise at the existence of our improbable but life-permitting universe that we do at the news of the latest lottery winner.  The problem, however, is that we don’t know that our universe was the winner of a perfectly fair cosmic lottery.  Lack of surprise is appropriate only when we have this knowledge.  The fact that a life-permitting universe is extraordinarily improbable raises the suspicion that our universe wasn’t randomly selected after all. 

            Consider the following unlikely events and the “explanations” offered of these events. 

i) The stones in one garden are randomly strewn about.  In another garden the stones spell “Welcome to Wales by British Railways.”  Regarding this example William Dembski (1998, p. xi) writes, “In both instances the precise arrangement of stones is vastly improbable.  Indeed, any given arrangement of stones is but one of an almost infinite number of possible arrangements.”  When asked for the best explanation of why one set of stones spells out an English sentence, someone replies, “Sometimes the improbable happens.” 

ii) On several occasions during the last week, large pieces of scrap metal have fallen from above and nearly killed me.  After each “accident” I turn around and see hurrying away from the scene one of my colleagues who has only a temporary contract with LSU but whose chances of being permanently hired by LSU would be greatly increased if I were out of the way.  When detained and questioned by the police about why he always seemed to be present when pieces of scrap metal were falling near my head, my colleague simply replies “Sometimes the improbable happens.” 

iii) I am brought before a firing squad made up of one hundred professional marksmen, each of whom is instructed to shoot one dozen rounds of ammunition at me.  All 1200 bullets fired at me miss their intended target.  When I ask someone for an explanation of this unlikely phenomenon, someone replies “Sometimes the improbable happens.” 

iv) A silk merchant who, while trying to sell a silk gown, keeps his thumb over a hole in the silk the entire time his customer is looking at the gown.  When his ruse is found out and he is asked to account for his behavior, he replies, “Every thumb must be somewhere.  While it is improbable that my thumb should cover the hole the entire time, it is equally improbable that my thumb should be at any other location on the gown.  Sometimes the improbable happens.” 

v) The winner of January’s state lottery was the nephew of the Lottery Commissioner.  The winner of the February lottery was the niece of the Lottery Commissioner.  The winner of the state lottery in March was the Lottery Commissioner’s brother.  The winner in April was the Lottery Commissioner’s ex-wife who, it is well known, has been trying to sue him for everything he’s got.  When asked to account for this highly improbable string of events the Lottery Commissioner replies, “Sometimes the improbable happens.” 

In none of these cases is the offered explanation even remotely satisfying or convincing.  When we lack the positive knowledge that an event is the outcome of a fair lottery (or its probabilistic equivalent), we find ourselves unable to accept the answer that “Sometimes the improbable happens.”  Our minds immediately turn to more likely scenarios that would explain the events in question.  We automatically assume that British Railways intentionally arranged the set of stones to be a greeting.  We think it highly likely that my colleague wants to bump me off so he can take my position.  We think the firing squad must be a sham that serves some unseen purpose.  We believe beyond any reasonable doubt that the location of the silk merchant’s thumb is due to greed and dishonesty rather than chance.  And no one, I take it, would believe the Lottery Commissioner’s claim to innocence. 

            Recall the fine-tuned features of the universe cited above.  The fact that all of these life-permitting features have come together is exceedingly improbable.  The defender of the IBE design argument claims that this situation is more similar to the five cases listed above than to a fair lottery.  As in the five cases above, they think we should be led to seek an explanation that does not appeal to mere chance.  That explanation, they suggest, is that the universe was purposefully created by an Intelligent Designer. 

            A third objection to the Inference to the Best Explanation version of the design argument stems from the fact that the conclusion of the argument is not necessitated by its premises.  Neal Gillespie (1979, pp. 83-84) has stated the objection as follows.

It has been generally agreed (then and since) that Darwin’s doctrine of natural selection effectively demolished William Paley’s classical design argument for the existence of God.  By showing how blind and gradual adaptation could counterfeit the apparently purposeful design that Paley... and others had seen in the contrivances of nature, Darwin deprived their argument of the analogical inference that the evident purpose to be seen in the contrivances by which means and ends were related in nature was necessarily a function of mind.

Although Gillespie’s objection is aimed at Paley’s analogical version of the design argument, it can be modified  to apply to the IBE version as well.  Gillespie takes the design argument to task for thinking that the apparent design of the universe “was necessarily a function of [an intelligent, creative] mind.”  But neither version of the design argument claims that the fine-tuned features of the universe are necessarily the product of intelligent design. 

            The IBE version merely claims that the hypothesis of intelligent design provides the best explanation for those features.  In other words, the design argument does not purport to be a deductive argument, in which the truth of the premises necessitates the truth of the conclusion.  Instead, it claims to offer a strong non-deductive argument for the hypothesis of intelligent design.  Pointing out that the premises of a non-deductive argument do not necessitate its conclusion is like pointing out that Einstein’s general theory of relativity does not explain how to make a great Cabernet.  That was never its intended purpose. 

            When dealing with non-deductive inferences, such as inferences to the best explanation, we must ask ourselves how much likelihood or palusibility is conferred upon the conclusion by the premises.  If the IBE design argument is strong, then the facts about fine-tuning make the conclusion about an Intelligent Designer highly probable.  If the argument is weak, then these facts do not make the Intelligent Design conclusion very probable at all.  The key point is that, when dealing with non-deductive arguments, the issue is always one of probability rather than necessity .  Strong, inductive arguments purport to make their conclusions probable.  They do not claim to necessitate their conclusions.  So, pointing out that they do not necessitate their premises cannot count as an objection against them.  The IBE design argument is an inference to the best explanation; not an inference to the only possible explanation.

  References

Brennan, Richard. 1997. Heisenberg Probably Slept Here: The Lives, Times, and Ideas of the Great Physicists of the 20th Century . New York: John Wiley & Sons.

Clark, Kelly James. 1990. Return to Reason: A Critique of Enlightenment Evidentialism and a Defense of Reason and Belief in God . Grand Rapids, MI: Eerdmans.

Craig, William Lane. 1990. “In Defense of Rational Theism.” In J. P. Moreland & Kai Nielsen (Eds.), Does God Exist? The Great Debate . Nashville, TN: Thomas Nelson Publishers.

Darwin, Charles. 1876/1958. Autobiography . Francis Darwin (Ed.). New York: Dover.

Dembski, William A. 1998. The Design Inference: Eliminating Chance Through Small Probabilities . Cambridge: Cambridge University Press.

Gillespie, Neal. 1979. Charles Darwin and the Problem of Creation . Chicago: University of Chicago Press.

Hick, John. 1964. The Existence of God . New York: Macmillan.

Hume, David. 1779/1947. Dialogues Concerning Natural Religion , edited with an introduction by N. K. Smith. New York: Macmillan.

Kant, Immanuel. 1781/1965. Critique of Pure Reason , trans. N. K. Smith. New York: St. Martin’s Press.

Leslie, John. 1989. Universes . London: Routledge.

Moreland, J. P. 1987. Scaling the Secular City: A Defense of Christianity . Grand Rapids, MI: Baker Book House.

Moreland, J. P. 1990. “Yes! A Defense of Christianity.” In J. P. Moreland & Kai Nielsen (Eds.), Does God Exist? The Great Debate . Nashville, TN: Thomas Nelson Publishers.

Moreland, J. P. 1994. “Introduction.” In J. P. Moreland (Ed.), The Creation Hypothesis: Scientific Evidence for an Intelligent Designer . Downers Grove, IL: InterVarsity Press.

Parsons, Keith. 1990. “Is There a Case for Christian Theism?” In J. P. Moreland & Kai Nielsen (Eds.), Does God Exist? The Great Debate . Nashville, TN: Thomas Nelson Publishers.

Plantinga, Alvin. 1990. God and Other Minds: A Study of the Rational Justification of Belief in God , paperback edition. Ithaca, NY: Cornell University Press.

Strobel, Lee. 2000. The Case for Faith: A Journalist Investigates the Toughest Objections to Christianity . Grand Rapids, MI: Zondervan.

  • Evolutionary Concepts
  • Open access
  • Published: 24 October 2009

The Argument from Design: A Guided Tour of William Paley’s Natural Theology (1802)

  • T. Ryan Gregory 1  

Evolution: Education and Outreach volume  2 ,  pages 602–611 ( 2009 ) Cite this article

104k Accesses

4 Citations

7 Altmetric

Metrics details

According to the classic “argument from design,” observations of complex functionality in nature can be taken to imply the action of a supernatural designer, just as the purposeful construction of human artifacts reveals the hand of the artificer. The argument from design has been in use for millennia, but it is most commonly associated with the nineteenth century English theologian William Paley and his 1802 treatise Natural Theology , or Evidence of the Existence and Attributes of the Deity , Collected from the Appearances of Nature . The book remains relevant more than 200 years after it was written, in large part because arguments very similar to Paley’s underlie current challenges to the teaching of evolution (indeed, his name arises with considerable frequency in associated discussions). This paper provides an accessible overview of the arguments presented by Paley in Natural Theology and considers them both in their own terms and in the context of contemporary issues.

I did not at that time [as a Cambridge theology student, 1827–1831] trouble myself about Paley’s premises; and taking these on trust, I was charmed and convinced by the long line of argumentation. (p. 59)
The old argument from design in nature, as given by Paley, which formerly seemed to me so conclusive, fails, now that the law of natural selection has been discovered. (p. 87) Charles Darwin, Autobiography

Introduction

The “teleological argument,” better known as the “argument from design,” is the claim that the appearance of “design” in nature—such as the complexity, order, purposefulness, and functionality of living organisms—can only be explained by the existence of a “designer” (typically of the supernatural variety). In its most familiar manifestation, the argument from design involves drawing parallels between human-designed objects (e.g., telescopes, outboard motors) and biological counterparts with similar functional roles (e.g., eyes, bacterial flagella). The former are complex, often indivisibly so if they are to maintain their current function, clearly perform specific functions, and are known to have been the product of intentional design. The functional complexity of living organisms is far greater still, it is argued, and, therefore, must present even stronger evidence for the role of intelligent agency.

Though the basic premise of the teleological argument had been articulated by thinkers as far back as ancient Greece and Rome, today it is almost universally associated with the writings of one person: William Paley (Fig.  1 ). Paley was born in July 1743 in Peterborough, Cambridgeshire, England. He was educated at Christ’s College, Cambridge, and was ordained a deacon in 1766 and soon thereafter a priest in Cambridge. He was appointed Archdeacon of Carlisle Cathedral in 1782 and awarded a Doctor of Divinity degree at Cambridge in 1795. He authored several successful theological works, the best-known being Natural Theology , or Evidence of the Existence and Attributes of the Deity , Collected from the Appearances of Nature .

figure 1

William Paley (1743–1805). Portrait circa 1790 by George Romney (National Portrait Gallery, London)

Natural Theology was published in 1802, only three years before Paley’s death on May 25, 1805. It was very successful, going through ten editions in the first four years alone (see Fyfe 2002 ). Despite being written in labyrinthine prose (by modern standards), Natural Theology remains an especially lucid exposition of the classic argument from design. This undoubtedly is one of the reasons that Paley’s name is most commonly linked with the design argument even though it was by no means original to him. Footnote 1

Darwin was influenced by Paley’s work, and some modern authors have cited it as an important example of pre-Darwinian “adaptationist” thinking (e.g., Dawkins 1986 ; Williams 1992 ; but see Gliboff 2000 ; McLaughlin 2008 ). Whatever its significance in the past, it is clear that Paley’s contribution continues to be of direct relevance in the current educational and political climate. Notably, Natural Theology was exhibit P-751 in the landmark 2005 Kitzmiller v. Dover trial, which successfully challenged the constitutionality of promoting “intelligent design” in US public schools. Paley’s name appears more than 80 times in the trial testimony transcripts, and he is mentioned a further half a dozen times in Judge Jones’s decision. Footnote 2

In light of their continuing importance in current discourse, it is worth exploring the arguments presented in Paley’s classic treatise. This review is intended to provide a “guided tour” of Natural Theology , Footnote 3 giving the reader an abridged and annotated rendition of Paley’s widely referenced (but less often read) account of the argument from design.

Watches and Watchmakers

Natural Theology opens with the paragraph for which it is best (if not exclusively) known, in which Paley draws a contrast between a rock and a pocket watch: Footnote 4

In crossing a heath, suppose I pitched my foot against a stone , and were asked how the stone came to be there; I might possibly answer, that, for any thing I knew to the contrary, it had lain there for ever: nor would it perhaps be very easy to show the absurdity of this answer. But suppose I had found a watch upon the ground, and it should be inquired how the watch happened to be in that place; I should hardly think of the answer which I had before given, that, for any thing I knew, the watch might have always been there. Yet why should not this answer serve for the watch as well as for the stone? why is it not as admissible in the second case, as in the first? For this reason, and for no other, viz. that, when we come to inspect the watch, we perceive (what we could not discover in the stone) that its several parts are framed and put together for a purpose, e. g. that they are so formed and adjusted as to produce motion, and that motion so regulated as to point out the hour of the day; that, if the different parts had been differently shaped from what they are, of a different size from what they are, or placed after any other manner, or in any other order, than that in which they are placed, either no motion at all would have been carried on in the machine, or none which would have answered the use that is now served by it. (p.1–2)

Thus, in the very first passage of a book written more than two centuries ago, Paley encapsulates the core components of the argument from design—an argument that has been revived in much the same form by proponents of “intelligent design.” Specifically, Paley points out that the watch exhibits an irreducibly complex organization that was obviously constructed to perform a specific function. Remove or rearrange any of its intricate inner workings, and the watch becomes barely more effective at keeping time than the rock formerly dismissed with a kick. Footnote 5 From this, Paley concludes that

...the inference, we think, is inevitable, that the watch must have had a maker: that there must have existed, at some time, and at some place or other, an artificer or artificers who formed it for the purpose which we find it actually to answer; who comprehended its construction, and designed its use. (p.3)

Furthermore, like modern proponents of the argument from design, Paley argues that one need not know any details of the designer’s identity or methods to conclude that an intelligent agent was involved:

Nor would it, I apprehend, weaken the conclusion, that we had never seen a watch made; that we had never known an artist capable of making one; that we were altogether incapable of executing such a piece of workmanship ourselves, or of understanding in what manner it was performed; all this being no more than what is true of some exquisite remains of ancient art, of some lost arts, and, to the generality of mankind, of the more curious productions of modern manufacture. (p.3–4)

Today’s neo-Paleyans must also concur with Paley that the watch’s delicate functionality could not be the product of chance, inherent “principles of order,” or laws of matter, nor merely an illusion of design. For Paley, this conclusion supersedes all other considerations and renders additional details largely superfluous. As he wrote,

Neither...would our observer be driven out of his conclusion, or from his confidence in its truth, by being told that he knew nothing at all about the matter. He knows enough for his argument: he knows the utility of the end: he knows the subserviency and adaptation of the means to the end. These points being known, his ignorance of other points, his doubts concerning other points, affect not the certainty of his reasoning. The consciousness of knowing little, need not beget a distrust of that which he does know. Footnote 6 (p.7)

However, given their aggressive resistance to the notion of suboptimality or nonfunction of biological structures, Footnote 7 it appears that many modern design proponents disagree with Paley’s subsequent assertions that neither imperfections nor ambiguous—or even nonexistent—functions refute the thesis of design for the origin of complex, (mostly) functional objects or organs. In this sense, Paley could be said to adhere more closely to the analogy with human artifacts than do many of his present-day counterparts:

Neither, secondly, would it invalidate our conclusion, that the watch sometimes went wrong, or that it seldom went exactly right. The purpose of the machinery, the design, and the designer, might be evident, and in the case supposed would be evident, in whatever way we accounted for the irregularity of the movement, or whether we could account for it or not. It is not necessary that a machine be perfect, in order to show with what design it was made: still less necessary, where the only question is, whether it were made with any design at all. (p.4–5)
Nor, thirdly, would it bring any uncertainty into the argument, if there were a few parts of the watch, concerning which we could not discover, or had not yet discovered, in what manner they conduced to the general effect; or even some parts, concerning which we could not ascertain, whether they conduced to that effect in any manner whatever. For...if by the loss, or disorder, or decay of the parts in question, the movement of the watch were found in fact to be stopped, or disturbed, or retarded, no doubt would remain in our minds as to the utility or intention of these parts, although we should be unable to investigate the manner according to which, or the connexion by which, the ultimate effect depended upon their action or assistance; and the more complex is the machine, the more likely is this obscurity to arise. Then, as to the second thing supposed, namely, that there were parts which might be spared, without prejudice to the movement of the watch, and that we had proved this by experiment,—these superfluous parts, even if we were completely assured that they were such, would not vacate the reasoning which we had instituted concerning other parts. The indication of contrivance remained, with respect to them, nearly as it was before. (p.5–6)

Continuing with the analogy of the watch, Paley next argues that one could not explain away the evidence of design even if the watch in hand had, through some exceptional mechanics, been produced by the self-replication of a parental watch. It matters not, according to Paley, whether any particular entity had been born of similar entities, as this accounts only for its existence and not its complex functional characteristics. Indeed, discovering the watch’s capacity to reproduce would only increase an observer’s admiration for its remarkable complexity:

No answer is given to this question, by telling us that a preceding watch produced it. There cannot be design without a designer; contrivance without a contriver; order without choice; arrangement, without any thing capable of arranging; subserviency and relation to a purpose, without that which could intend a purpose; means suitable to an end, and executing their office, in accomplishing that end, without the end ever having been contemplated, or the means accommodated to it. Arrangement, disposition of parts, subserviency of means to an end, relation of instruments to a use, imply the presence of intelligence and mind. No one, therefore, can rationally believe, that the insensible, inanimate watch, from which the watch before us issued, was the proper cause of the mechanism we so much admire in it;—could be truly said to have constructed the instrument, disposed its parts, assigned their office, determined their order, action, and mutual dependency, combined their several motions into one result, and that also a result connected with the utilities of other beings. All these properties, therefore, are as much unaccounted for, as they were before. (p.11–12)

Thus, Paley argues, no matter how many generations of watches beget watches (or, by obvious implication, organisms produce offspring or cells generate daughter cells), the specific, irreducible, and purposeful arrangement of watches’ inner workings can only be attributed to the action of intelligent agency.

The Cosmic Optician

Having established the connection between watches and watchmakers, Paley begins his third chapter by arguing that the principle applies equally to living organisms and their components—or indeed, more so, given that their degree of adaptive complexity is vastly greater. As he might have argued, human hands more thoroughly evince design than anything crafted by them.

As did many of his predecessors Footnote 8 (and followers), Paley considered eyes to provide a particularly illuminating exemplar of organic design: “there is precisely the same proof that the eye was made for vision, as there is that the telescope was made for assisting it” (p. 18). Paley noted that telescopes and eyes rely on similar optical principles but that in fact vertebrate eyes are much more effective by virtue of their ability to adjust to different distances and brightness, in their well-developed protective features including eyelids and nictitating membranes, and by their capacity to correct for spherical aberration. In fact, he pointed out, telescope designers solved the problem of aberration by adopting features observed in biological lenses. In the absence of a natural explanation for their occurrence, eyes provided one of the best-known cases in support of the design argument. Darwin exposed the fallacy of this conclusion, and the efforts of countless scientists since then have resolved in increasingly fine detail how the various components of eyes are likely to have evolved. Footnote 9

In considering the eyes of different types of animals, Paley noted two critical facts: (1) that eyes differ according to the environment in which they are used to see and (2) that despite these differences, all vertebrate eyes are constructed according to the same basic physical plan. Eyes specialized for sight underwater, on land, or in the dark are not fundamentally different from each other; rather, they are modifications of a general theme: “Thus, in comparing the eyes of different kinds of animals, we see, in their resemblances and distinctions, one general plan laid down, and that plan varied with the varying exigencies to which it is to be applied” (p. 31). Today, this similarity amidst diversity is explained by the fact that all vertebrates share a common ancestor that was possessed of eyes and that specializations to different lifestyles have involved descent with modification of this ancestral organ.

The fact that all eyes have evolved through the modification of prior form and the cooption of preexisting components also explains some otherwise puzzling structural complications (not to mention features that are downright maladaptive; see Novella 2008 for several examples). However, through his teleological lens, Paley viewed complexities as further evidence of good design, as with his example of muscles in the eyes of cassowaries:

In the configuration of the muscle which, though placed behind the eye, draws the nictitating membrane over the eye, there is...a marvellous mechanism....The muscle is passed through a loop formed by another muscle: and is there inflected, as if it were round a pulley. This is a peculiarity; and observe the advantage of it. A single muscle with a straight tendon, which is the common muscular form, would have been sufficient, if it had had power to draw far enough. But the contraction, necessary to draw the membrane over the whole eye, required a longer muscle than could lie straight at the bottom of the eye. Therefore, in order to have a greater length in a less compass, the cord of the main muscle makes an angle. This, so far, answers the end; but, still further, it makes an angle, not round a fixed pivot, but round a loop formed by another muscle; which second muscle, whenever it contracts, of course twitches the first muscle at the point of inflection, and thereby assists the action designed by both. (p.37–38; italics in original)

In mammals, the recurrent laryngeal nerve provides a connection between the brain and the larynx, though not a direct one. Instead of taking a direct route, it passes down into the chest, circles under the aorta, and ascends back up to the neck (in giraffes, this nerve is more than 2 meters long; Harrison 1995 ). Similarly, the mammalian vas deferens connects the testes to the urethra, but not before passing into the pelvic cavity, looping around the urinary bladder and then descending back to complete its circuitous path. Meanwhile, the urethra itself passes directly through the prostate gland, an arrangement that readily engenders urinary difficulties if the prostate becomes swollen. It is only with great effort that arrangements such as these might be characterized as optimizations rather than as simple quirks of evolutionary history Footnote 10 (for additional examples, see Williams 1997 ; Shubin 2008 ; Coyne 2009 ).

But why bother with eyes at all? If the designer is omnipotent, why does the detection of visual information require such a complex arrangement of lenses, receptors, nerves, muscles, and neurons? In Paley’s words,

Why make the difficulty in order to surmount it? If to perceive objects by some other mode than that of touch, or objects which lay out of the reach of that sense, were the thing proposed; could not a simple volition of the Creator have communicated the capacity? Why resort to contrivance, where power is omnipotent? Contrivance, by its very definition and nature, is the refuge of imperfection. To have recourse to expedients, implies difficulty, impediment, restraint, defect of power. (p.39)

Thus answers Paley his own rhetorical queries:

...beside reasons of which probably we are ignorant, one answer is this: It is only by the display of contrivance, that the existence, the agency, the wisdom of the Deity, could be testified to his rational creatures. This is the scale by which we ascend to all the knowledge of our Creator which we possess, so far as it depends upon the phenomena, or the works of nature. Take away this, and you take away from us every subject of observation, and ground of reasoning; I mean as our rational faculties are formed at present. Whatever is done, God could have done without the intervention of instruments or means: but it is in the construction of instruments, in the choice and adaptation of means, that a creative intelligence is seen. It is this which constitutes the order and beauty of the universe. God, therefore, has been pleased to prescribe limits to his own power, and to work his end within those limits. (p.40)

According to Paley, the exquisite function of eyes bespeaks the great power of a designer, but the very decision to create eyes reflects an intentional limitation of this power so that humans might understand how eyes came to be. This logic may strike the modern reader as rather tortuous, but it serves to illustrate a very important point: that any explanation for complex organs must account not only for their adaptive characteristics but also their imperfections. Today, this duality can be accounted for by the countervailing influences of adaptive modification and the constraints of genetics, anatomy, and history, but for Paley, the intent of a designer provided the only conceivable answer. As noted, Paley did not consider imperfections to challenge the conclusion that design indicates the work of a designer. It is only if one wishes to defend the infallibility of the designer that one must assume all features of an object to be functional, and optimally so (see pp.56–57). Paley considered nonfunctional aspects of organisms to be “extremely rare,” and as is clear from later chapters, he viewed most aspects of the world to be optimally designed, but he was careful not to base his argument for the existence of a designer on these secondary considerations. Again, this represents something of a more sophisticated application of the argument from design than is often encountered in contemporary discourse.

Not a Chance

One of the most obstinate misconceptions about evolutionary theory is that it hypothesizes that eyes and other complex organs arise “by chance.” Even under the most charitable assessment, such a view of adaptive evolution must be considered deeply misguided. Whereas genetic mutation is both integral to the process and indeed is random with respect to its effects, natural selection is, by definition, the nonrandom survival and reproduction of individuals. Variation is generated at random, but whether or not it is preserved depends on its effects on survival and reproduction within a given environment Footnote 11 (for reviews, see Gregory 2008 , 2009 ). No serious evolutionary biologist of the past 150 years has suggested that the emergence of complex organs is merely the result of chance.

Writing as he did before Darwin and Wallace proposed the theory of natural selection, it was not possible for Paley to make this error (modern neo-Paleyans, by contrast, do so with remarkable proficiency). In the early 1800s, chance was not the only suggested alternative to conscious design (McLaughlin 2008 ), but Paley viewed its refutation as an important part of his argument. Paley (and Darwin) understood that chance plays a role in nature, but that it is incapable of producing adaptive complexity:

What does chance ever do for us? In the human body, for instance, chance, i.e. the operation of causes without design, may produce a wen, a wart, a mole, a pimple, but never an eye. Amongst inanimate substances, a clod, a pebble, a liquid drop might be; but never was a watch, a telescope, an organized body of any kind, answering a valuable purpose by a complicated mechanism, the effect of chance. (p.63)

Modern evolutionary biologists do not part company with Paley on the claim that complex organs must arise through a mechanism other than pure chance. The disagreement is only with his subsequent assertion, that “in no assignable instance hath such a thing existed without intention somewhere” (p.63).

Paley takes his argument against the role of chance a step farther, in the process raising—and summarily rejecting—a possible explanation that exhibits shades of the principle of natural selection. However, his description is of an ancient version of the idea that, as Paley rightly notes, is unworkable in practice. Footnote 12

There is another answer which has the same effect as the resolving of things into chance; which answer would persuade us to believe, that the eye, the animal to which it belongs, every other animal, every plant, indeed every organized body which we see, are only so many out of the possible varieties and combinations of being, which the lapse of infinite ages has brought into existence; that the present world is the relict of that variety: millions of other bodily forms and other species having perished, being by the defect of their constitution incapable of preservation, or of continuance by generation. Now there is no foundation whatever for this conjecture in any thing which we observe in the works of nature; no such experiments are going on at present: no such energy operates, as that which is here supposed, and which should be constantly pushing into existence new varieties of beings. Nor are there any appearances to support an opinion, that every possible combination of vegetable or animal structure has formerly been tried. Multitudes of conformations, both of vegetables and animals, may be conceived capable of existence and succession, which yet do not exist. Perhaps almost as many forms of plants might have been found in the fields, as figures of plants can be delineated upon paper. A countless variety of animals might have existed, which do not exist. Upon the supposition here stated, we should see unicorns and mermaids, sylphs and centaurs, the fancies of painters, and the fables of poets, realized by examples. Or, if it be alleged that these may transgress the limits of possible life and propagation, we might, at least, have nations of human beings without nails upon their fingers, with more or fewer fingers and toes than ten, some with one eye, others with one ear, with one nostril, or without the sense of smelling at all. All these, and a thousand other imaginable varieties, might live and propagate. We may modify any one species many different ways, all consistent with life, and with the actions necessary to preservation, although affording different degrees of conveniency and enjoyment to the animal. And if we carry these modifications through the different species which are known to subsist, their number would be incalculable. No reason can be given why, if these deperdits ever existed, they have now disappeared. Yet, if all possible existences have been tried, they must have formed part of the catalogue. (p.63–65)

The problem, of course, is not the notion that great variety may arise by chance and be narrowed by differential survival—this is the basis of natural selection as it is now understood. Rather, the implausibility of Paley’s scenario is the scale at which he considered the process. Specifically, he envisioned an unconstrained morphospace in which drastically divergent species continually pop into existence. This lies in stark contrast to Darwin’s later emphasis on small-scale variation arising within species and then being sorted generation by generation.

In this context, it is also worth noting Paley’s view on extinction—namely that it does not happen. According to Paley, the classification of species into larger taxa would be rendered impossible by widespread extinction. In contrast, extinction was established as a common process in the history of life by Darwin’s time, and today it is acknowledged that the overwhelming majority of species that have existed no longer grace the Earth. In fact, the major divisions among extant lineages are now understood to exist precisely because so many ancestors and intermediate forms have perished.

Design and Diversity

The preceding arguments occupy only the first six of the 27 chapters in Natural Theology . In Chapters 7 through 20, Paley leads the reader on an expedition through the annals of early nineteenth century biological knowledge as he understands it, pausing along the way to admire the elegance of the bones (Chapter 8), muscles (Chapter 9), blood vessels (Chapter 10), and digestive systems (Chapters 7 and 10) of vertebrates, as well as features of insects (Chapter 19) and plants (Chapter 20) which, though less well understood, he also described as bearing the hallmarks of design.

Once again, modern biology does not contradict Paley’s enthusiastic exposition of features well-suited to specific functions, only the way in which their origin is explained. Similarly, Paley takes a broad comparative approach that would be at home in modern evolutionary biology were it interpreted from a different perspective. He recognizes that specializations for particular lifestyles reflect modifications of traits shared by many animals. He grasps the unity of underlying body plans. And he notes that though it dissipates among groups living in widely divergent habitats, the similarity does not disappear. Consider the following passages from Chapter 12 on “Comparative Anatomy”:

Whenever we find a general plan pursued, yet with such variations in it as are, in each case, required by the particular exigency of the subject to which it is applied, we possess, in such plan and such adaptation, the strongest evidence that can be afforded of intelligence and design; an evidence which most completely excludes every other hypothesis. If the general plan proceeded from any fixed necessity in the nature of things, how could it accommodate itself to the various wants and uses which it had to serve under different circumstances, and on different occasions? (p.211)
Very much of this reasoning is applicable to what has been called Comparative Anatomy . In their general economy, in the outlines of the plan, in the construction as well as offices of their principal parts, there exists between all large terrestrial animals a close resemblance. In all, life is sustained, and the body nourished by nearly the same apparatus. The heart, the lungs, the stomach, the liver, the kidneys, are much alike in all. The same fluid (for no distinction of blood has been observed) circulates through their vessels, and nearly in the same order. The same cause, therefore, whatever that cause was, has been concerned in the origin, has governed the production of these different animal forms.
When we pass on to smaller animals, or to the inhabitants of a different element, the resemblance becomes more distant and more obscure; but still the plan accompanies us. (pp.212–213)

Chapter 13 deals with the opposite subject, namely adaptations (“Peculiar Organizations”) that are unique to particular groups: features of the neck of large mammals, the swim bladder of fishes, the fangs of snakes, the pouches of marsupials, the claws of birds, the stomach of camels, the tongue of woodpeckers, and the curved tusks of wild boars. These, like functional traits shared more broadly, also reflect the remarkable fit of species to their environments which Paley takes as strong evidence for their origin by design.

In Chapter 14, Paley lends particular credence to examples of adaptive features that emerge ontogenetically before they are needed, in preparation for use later in life (“Prospective Contrivances”):

I can hardly imagine to myself a more distinguishing mark, and, consequently, a more certain proof of design, than preparation , i.e. the providing of things beforehand, which are not to be used until a considerable time afterwards; for this implies a contemplation of the future, which belongs only to intelligence. (p.252)

The teeth of mammals, the milk that nourishes their young, their eyes that develop while still in the darkness of the womb, and their lungs that form before encountering any opportunity to draw a breath—in the absence of knowledge about developmental genetics, these struck Paley as especially weighty examples of foresightful design.

Fitting Together

Paley does not only rely on individual examples of function to support his position. In addition, he expounds upon the close interaction of parts in service of a specific function. He returns to the analogy of the watch in this capacity at the opening of Chapter 15:

When several different parts contribute to one effect; or, which is the same thing, when an effect is produced by the joint action of different instruments; the fitness of such parts or instruments to one another, for the purpose of producing, by their united action the effect, is what I call relation: and wherever this is observed in the works of nature or of man, it appears to me to carry along with it decisive evidence of understanding, intention, art. In examining, for instance, the several parts of a watch , the spring, the barrel, the chain, the fusee, the balance, the wheels of various sizes, forms, and positions, what is it which would take an observer's attention, as most plainly evincing a construction, directed by thought, deliberation, and contrivance? It is the suitableness of these parts to one another; first, in the succession and order in which they act; and, secondly, with a view to the effect finally produced. (pp. 261–262)

These “relations” are common in nature, and they form a central part of the modern incarnation of the argument from design just as they did two centuries before (albeit recast in terms of the “specified,” “irreducibly complex,” or “purposeful” arrangements of parts in service of a particular function). Contemporary embodiments of the teleological argument typically appeal to examples from microbiology and biochemistry, but the basic approach is the same as in Paley’s discussion of systems involved in feeding, digestion, and excretion. Paley is especially taken by close interactions among independent components (which today are explained as the product of co-evolution):

But relation perhaps is never so striking as when it subsists, not between different parts of the same thing, but between different things. The relation between a lock and a key is more obvious, than it is between different parts of the lock. A bow was designed for an arrow, and an arrow for a bow: and the design is more evident for their being separate implements.
Nor do the works of the Deity want this clearest species of relation. The sexes are manifestly made for each other. They form the grand relation of animated nature; universal, organic, mechanical; subsisting like the clearest relations of art, in different individuals; unequivocal, inexplicable without design.
So much so, that, were every other proof of contrivance in nature dubious or obscure, this alone would be sufficient. The example is complete. Nothing is wanting to the argument. I see no way whatever of getting over it. (pp. 268–270)

According to Paley, the fit of parts is also especially significant “when the defects of one part, or of one organ, are supplied by the structure of another part or of another organ” (p.275), a special case of “relation” that he dubs “compensation.” For example, the neck of an elephant is inflexible, but this is compensated for by its dexterous trunk. Paley’s ideas of “inconveniency” and “compensation” are recognizable in the modern concept of tradeoffs:

When I speak of an inconveniency, I have a view to a dilemma which frequently occurs in the works of nature, viz. that the peculiarity of structure by which an organ is made to answer one purpose, necessarily unfits it for some other purpose. (pp.278–279)

Not surprisingly, the fit of organisms to their physical environments (e.g., wings for movement in air, fins in water) also provides evidence for design under Paley’s view:

We have already considered relation , and under different views; but it was the relation of parts to parts, of the parts of an animal to other parts of the same animal, or of another individual of the same species.
But the bodies of animals hold, in their constitution and properties, a close and important relation to natures altogether external to their own; to inanimate substances, and to the specific qualities of these, e.g. they hold a strict relation to the elements by which they are surrounded. (p.291; italics in original)

Paley also contemplates instincts as an example of “relations” indicative of design. In so doing, he confronts what are now labeled as proximate versus ultimate causes (and, in a long-outdated dichotomy, “nature vs. nurture”). Citing examples of reproductive behaviors—the incubation of eggs by birds, the host specificity of egg laying by butterflies, the spawning journeys of salmon—he argues that many behaviors are neither learned nor simple reactions to stimuli but are hardwired by design.

Beyond Biology

For Paley, evidence of design is not restricted to examples from the living world but also comes from the conditions of the world that make life possible. Chapter 21, dedicated to “The Elements,” considers the beneficial properties of air, water, fire, and light. The extent of Paley’s teleological reasoning is illustrated with particular clarity in the case of light which, he argues, travels so fast that it might obliterate both eyes and beholders had the designer not provided a safeguard: Footnote 13

Urged by such a velocity, with what force must its particles drive against (I will not say the eye, the tenderest of animal substances, but) every substance, animate or inanimate, which stands in its way! It might seem to be a force sufficient to shatter to atoms the hardest bodies.
How then is this effect, the consequence of such prodigious velocity, guarded against? By a proportionable minuteness of the particles of which light is composed. (p.376)

The properties of the sun, the planets, and the physical laws that govern their motions also point to the action of a designer according to Paley’s interpretation. For example, the sun, which provides heat and light, is situated in a convenient central location; the Earth’s axis of rotation provides climatological stability; and the laws of gravity are restricted to the miniscule subset of all possible configurations that is compatible with life. Paley held to a particularly strong version of the anthropic principle (not unlike one still invoked by many modern creationists):

That the subsisting law of attraction falls within the limits which utility requires, when these limits bear so small a proportion to the range of possibilities upon which chance might equally have cast it, is not, with any appearance of reason, to be accounted for, by any other cause than a regulation proceeding from a designing mind. But our next proposition carries the matter somewhat further. We say, in the third place, that, out of the different laws which lie within the limits of admissible laws, the best is made choice of; that there are advantages in this particular law which cannot be demonstrated to belong to any other law; and, concerning some of which, it can be demonstrated that they do not belong to any other. (pp. 395–396)

The Nature of the Maker

Unlike modern proponents of “intelligent design,” Paley is totally forthright in his view on the identity of the designer. As he concludes at the close of Chapter 23, “The marks of design are too strong to be gotten over. Design must have had a designer. That designer must have been a person. That person is God.” (p.441). From this starting position, Paley proceeds in Chapters 24 through 26 to rally evidence from nature for the omnipotence, omniscience, and benevolence of the Designer. He must, therefore, contend with examples not only confirmatory (e.g., the good fit of organisms to their needs, the capacity for pleasure) but also seemingly contradictory (e.g., death, disease, predation, pain) to such a theological stance. Thus, predation is presented as a form of population control that prevents the potentially disastrous effects of superfecundity (and the application of venom is a particularly merciful method of dispatching prey). Other apparent ills are depicted as being similarly beneficial upon closer inspection. As Paley argues, “It is a happy world after all” (p. 456).

Design Since Darwin

As Darwin noted, Paley’s thesis that the appearance of design must in fact be the outcome of design was refuted by the advent of a workable theory of evolutionary change. Nevertheless, 150 years later, biologists are still awed—but are no longer stunned—by complexity in natural systems. Evolutionary theory provides a means of exploring the origin of complex adaptations using a variety of analytical approaches (e.g., fossil record, genetics, comparative anatomy and physiology, phylogenetics, developmental biology), rather than drawing a conclusion based on the observation of complexity alone. Evolutionary theory, which includes more than adaptive mechanisms, also provides a straightforward explanation for suboptimality, vestigial traits, and wastefulness, including excessive complexity and redundancy where simpler solutions could easily be envisioned.

Interestingly, evolutionary biologists who strongly emphasize adaptive evolution by natural selection (e.g., Dawkins 1986 ; Williams 1992 )—the theory that undermined Paley’s central conclusion—more often express positive opinions of Natural Theology than do modern proponents of the argument from design. It is not difficult to see why, as Paley passionately addressed what they see as the most important question in evolutionary biology: the origins of complex adaptations. Contemporary design proponents, by contrast, generally attempt to distance themselves from Paley so as not to be seen as simply repackaging a 200-year-old theological argument. The examples may have changed (e.g., from eyes to flagella), but the premise has not.

Concluding Remarks

Like many of the early works that have had a lasting impact on biological thought, Paley’s Natural Theology remains worthy of reading by the current generation of biologists, educators, and students. It presents the argument from design in a clear fashion, unencumbered by the mathematical and biochemical accoutrements with which it recently has been festooned. That arguments very similar to Paley’s are now being used in an attempt to undermine the teaching and acceptance of evolutionary science gives Natural Theology continued relevance. From a much more positive point of view, his wonder at the complexity of the natural world, expressed with sincerity and style, is something to which all evolutionary biologists can relate. It is to be hoped that, in time, it will be for this reason alone that Paley’s classic treatise retains its significance.

John Ray’s 1691 treatise The Wisdom of God Manifested in the Works of the Creation was a notable forerunner (available online: http://books.google.ca/books?id=HvQ4AAAAMAAJ ).

Online copies are available for both the transcripts ( http://ncseweb.org/creationism/legal/kitzmiller-trial-transcripts ) and the decision ( http://www.pamd.uscourts.gov/kitzmiller/kitzmiller_342.pdf ).

The first edition of Natural Theology of 1802 is available in the Oxford World’s Classics series from Oxford University Press. Quotes and associated page numbers given in this review come from the 12th edition, as provided by The Complete Work of Charles Darwin Online ( http://darwin-online.org.uk/content/frameset?itemID=A142 ). A later edition is also available at Google Books ( http://books.google.com/books?id=-XFHAAAAIAAJ ). The differences between editions are minor.

Like many of the arguments most famously attributed to Paley, an analogy to a timepiece was used by various thinkers before him. For example, Marcus Tullius Cicero wrote in his De Natura Deorum in 45 BCE: “When you see a sundial or a waterclock, you see that it tells the time by design and not by chance. How then can you imagine that the universe as a whole is devoid of purpose and intelligence, when it embraces everything, including these artifacts themselves and their artificers?” Numerous other examples are reviewed by Moore ( 2009 ).

But, as the old cliché goes, even a broken watch gives the correct time twice per day.

For Paley, biochemistry represented a particularly impenetrable frontier that reinforced the existence of a designer with a more thorough knowledge of chemistry (pp.83–91). For their part, modern design proponents tend to focus on subcellular and molecular features whose origins have only recently become amenable to investigation.

A prime example is provided by so-called junk DNA. A common myth, repeatedly invoked by anti-evolutionists, science writers, and misinformed biologists alike, asserts that potential functions for non-genic components of the genome were long dismissed by mainstream science. Reference to the scientific literature from the relevant time period thoroughly undermines this claim ( http://genomicron.blogspot.com/2008/02/junk-dna-quotes-of-interest-series.html ).

For example, Paley cites the work of Johann Christophorous Sturm from a century before: “Sturmius held, that the examination of the eye was a cure for atheism” (p.33). Paley himself wrote that “Were there no example in the world, of contrivance, except that of the eye , it would be alone sufficient to support the conclusion which we draw from it, as to the necessity of an intelligent Creator” (p.75).

For two recent special issues of peer-reviewed journals dedicated to eye evolution, see Evolution: Education and Outreach volume 1, issue 4, Oct. 2008 ( http://www.springerlink.com/content/m3k441k67q3n/ ) and Philosophical Transactions of the Royal Society B volume 364, issue 1531, Oct. 19, 2009 ( http://rstb.royalsocietypublishing.org/content/364/1531.toc ).

Paley does make an effort to portray the tiny vestigial eyes of moles, which are permanently covered with skin, as an example of good design (pp.273–274).

This refers specifically to evolution by natural selection. Variation can also be sorted by chance through genetic drift, especially when it is neutral with respect to reproductive success and in small populations, but this mechanism is not responsible for the evolution of complex adaptations.

Similar notions of extreme trial-and-error date back to the ancient Greeks, including Lucretius and Empedocles (Gliboff 2000 ).

Indeed, he later states that “The eye is made for light , and light for the eye. The eye would be of no use without light, and light perhaps of little without eyes” (p.424).

Coyne JA. Why evolution is true. New York: Viking; 2009.

Google Scholar  

Dawkins, R. The Blind Watchmaker. New York: Penguin Books; 1986.

Fyfe A. Publishing and the classics: Paley's Natural theology and the nineteenth-century scientific canon. Stud Hist Phil Sci. 2002;33:729–51.

Article   Google Scholar  

Gliboff S. Paley's design argument as an inference to the best explanation, or, Dawkins' dilemma. Stud Hist Phil Biol Biomed Sci. 2000;31:579–97.

Gregory TR. The evolution of complex organs. Evolution: Education and Outreach. 2008;1:358–89.

Gregory TR. Understanding natural selection: essential concepts and common misconceptions. Evolution: Education and Outreach. 2009;2:156–75.

Harrison OFN. The anatomy and physiology of the mammalian larynx. Cambridge: Cambridge University Press; 1995.

Book   Google Scholar  

McLaughlin P. Reverend Paley's naturalist revival. Stud Hist Phil Biol Biomed Sci. 2008;39:25–37.

Moore R. William Paley, 1743–1805. Reports of the NCSE. 2009;29:26–8.

Novella S. Suboptimal optics: vision problems as scars of evolutionary history. Evolution: Education and Outreach. 2008;1:493–7.

Paley W. Natural Theology, or evidences of the existence and attributes of the deity collected from the appearances of nature. Oxford: Oxford University Press; 1802.

Shubin N. Your inner fish. New York: Pantheon; 2008.

Williams GC. Natural selection: domains, levels, and challenges. Oxford: Oxford University Press; 1992.

Williams GC. The pony fish’s glow. New York: Basic Books; 1997.

Download references

Author information

Authors and affiliations.

Department of Integrative Biology, University of Guelph, Guelph, ON, N1G 2W1, Canada

T. Ryan Gregory

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to T. Ryan Gregory .

Rights and permissions

Open Access This is an open access article distributed under the terms of the Creative Commons Attribution Noncommercial License ( https://creativecommons.org/licenses/by-nc/2.0 ), which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.

Reprints and permissions

About this article

Cite this article.

Gregory, T.R. The Argument from Design: A Guided Tour of William Paley’s Natural Theology (1802). Evo Edu Outreach 2 , 602–611 (2009). https://doi.org/10.1007/s12052-009-0184-6

Download citation

Published : 24 October 2009

Issue Date : December 2009

DOI : https://doi.org/10.1007/s12052-009-0184-6

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Teleological argument

Evolution: Education and Outreach

ISSN: 1936-6434

the design argument essay

SEP home page

  • Table of Contents
  • Random Entry
  • Chronological
  • Editorial Information
  • About the SEP
  • Editorial Board
  • How to Cite the SEP
  • Special Characters
  • Advanced Tools
  • 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

Teleological Arguments for God’s Existence

Some phenomena within nature exhibit such exquisiteness of structure, function or interconnectedness that many people have found it natural to see a deliberative and directive mind behind those phenomena. The mind in question is typically taken to be supernatural. Philosophically inclined thinkers have both historically and at present labored to shape the relevant intuition into a more formal, logically rigorous inference. The resultant theistic arguments, in their various logical forms, share a focus on plan, purpose, intention, and design, and are thus classified as teleological arguments (or, frequently, as arguments from or to design).

Although enjoying some prominent defenders over the centuries, such arguments have also attracted serious criticisms from major historical and contemporary thinkers. Critics and advocates include not only philosophers but also scientists and thinkers from other disciplines as well. In the following discussion, major variant forms of teleological arguments will be distinguished and explored, traditional philosophical and other criticisms will be discussed, and the most prominent contemporary turns (cosmic fine tuning arguments, many-worlds theories, and the Intelligent Design debate) will be tracked. Discussion will conclude with a brief look at one historically important non-inferential approach to the issue.

1. Introduction

2.1 analogical design arguments: schema 1, 2.2 deductive design arguments: schema 2, 2.3 inferences to the best explanation/abductive design arguments: schema 3.

  • 3.1 Explaining Away

3.2 Indirect Causation, Design and Evidences

4.1.1 no explanation needed, 4.1.2 rival explanations.

  • 4.2 Biological: The “Intelligent Design” Movement

5. The Persistence of Design Thinking

6. conclusion, other internet resources, related entries.

It is not uncommon for humans to find themselves with the intuition that random, unplanned, unexplained accident just couldn’t produce the order, beauty, elegance, and seeming purpose that we experience in the natural world around us. As Hume’s interlocutor Cleanthes put it, we seem to see “the image of mind reflected on us from innumerable objects” in nature (Hume 1779 [1998], 35). And many people find themselves convinced that no explanation for that mind-resonance which fails to acknowledge a causal role for intelligence, intent and purpose in nature can be seriously plausible.

Cosmological arguments often begin with the bare fact that there are contingently existing things and end with conclusions concerning the existence of a cause with the power to account for the existence of those contingent things. Others reason from the premise that the universe has not always existed to a cause that brought it into being. Teleological arguments (or arguments from design ) by contrast begin with a much more specialized catalogue of properties and end with a conclusion concerning the existence of a designer with the intellectual properties (knowledge, purpose, understanding, foresight, wisdom, intention) necessary to design the things exhibiting the special properties in question. In broad outline, then, teleological arguments focus upon finding and identifying various traces of the operation of a mind in nature’s temporal and physical structures, behaviors and paths. Order of some significant type is usually the starting point of design arguments.

Design-type arguments are largely unproblematic when based upon things nature clearly could not or would not produce (e.g., most human artifacts), or when the intelligent agency is itself ‘natural’ (human, alien, etc.). Identifying designed traces of ‘lost’ human civilizations or even non-human civilizations (via SETI) could in principle be uncontroversial. Objections to design inferences typically arise only when the posited designer is something more exotic or perhaps supernatural.

But despite the variety of spirited critical attacks they have elicited, design arguments have historically had and continue to have widespread intuitive appeal—indeed, it is sometimes claimed that design arguments are the most persuasive of all purely philosophical theistic arguments. Note that while design arguments have traditionally been employed to support theism over metaphysical naturalism, some might also be relevant for panentheism, panpsychism, and other views involving irreducible teleology.

2. Design Inference Patterns

The historical arguments of interest are precisely the potentially problematic ones—inferences beginning with some empirical features of nature and concluding with the existence of a designer. A standard but separable second step—the natural theology step—involves identifying the designer as God, often via particular properties and powers required by the designing in question. Although the argument wielded its greatest intellectual influence during the 18th and early 19th centuries, it goes back at least to the Greeks and in extremely clipped form comprises one of Aquinas’s Five Ways. It was given a fuller and quite nice early statement by Hume’s Cleanthes (1779 [1998], 15).

The question remains, however, about the formal structure of such arguments. What sort of logic is being employed? As it turns out, that question does not have just a single answer. Several distinct answers are canvassed in the following sections.

Design arguments are routinely classed as analogical arguments—various parallels between human artifacts and certain natural entities being taken as supporting parallel conclusions concerning operative causation in each case. The standardly ascribed schema is roughly thus:

Schema 1 : Entity e within nature (or the cosmos, or nature itself) is like specified human artifact a (e.g., a machine) in relevant respects R . a has R precisely because it is a product of deliberate design by intelligent human agency. Like effects typically have like causes (or like explanations, like existence requirements, etc.) Therefore It is (highly) probable that e has R precisely because it too is a product of deliberate design by intelligent, relevantly human- like agency.

2.1.1 Humean objections

This general argument form was criticized quite vigorously by Hume, at several key steps. Against (1), Hume argued that the analogy is not very good—that nature and the various things in it are not very like human artifacts and exhibit substantial differences from them—e.g., living vs. not, self-sustaining vs. not. Indeed, whereas advocates of design arguments frequently cited similarities between the cosmos on the one hand and human machines on the other, Hume suggested that the cosmos much more closely resembled a living organism than a machine. (Contemporary cosmology strengthens Hume’s point, given the radical changes that have taken place since the Big Bang. The universe looks less stable and “machine-like” than when the design argument was at the height of its popularity.) But if the alleged resemblance is in relevant respects distant, then the inference in question will be logically fragile. And while (2) may be true in specific cases of human artifacts a , that fact is only made relevant to natural phenomena e via (3), which underpins the transfer of the key attribution. Against (3), Hume argued that any number of alternative possible explanations could be given of allegedly designed entities in nature—chance, for instance. Thus, even were (1) true and even were there important resemblances, the argument might confer little probabilistic force onto the conclusion.

More generally, Hume also argued that even if something like the stated conclusion (4) were established, that left the arguer far from anything like a traditional conception of God. For instance, natural evils or apparently suboptimal designs might suggest e.g., an amateur designer or a committee of designers. And if phenomena instrumental to the production of natural evils (e.g., disease microorganisms) exhibited various of the R s, then they would presumably have to be laid at the designer’s door, further eroding the designer’s resemblance to the wholly good deity of tradition. And even the most impressive empirical data could properly establish only finite (although perhaps enormous) power and wisdom, rather than the infinite power and wisdom usually associated with divinity. But even were one to concede some substance to the design argument’s conclusion, that would, Hume suggested, merely set up a regress. The designing agent would itself demand explanation, requiring ultimately a sequence of prior analogous intelligences producing intelligences. And even were the existence of a designer of material things established, that did not yet automatically establish the existence of a creator of the matter so shaped. And since analogical arguments are a type of induction (see the entry on analogy and analogical reasoning ), the conclusion even if established would be established only to some, perhaps insignificant, degree of probability. Furthermore, we could not ground any induction concerning the cosmos itself upon a requisite fund of experiences of other cosmoi found to be both deliberately designed and very like ours in relevant respects—for the simple reason that this universe is our only sample. And finally the fraction of this one cosmos (both spatially and temporally) available to our inspection is extraordinarily small—not a promising basis for a cosmically general conclusion. Hume concluded that while the argument might constitute some limited grounds for thinking that “ the cause or causes of order in the universe probably bear some remote analogy to human intelligence ” (1779 [1998], 88) Hume’s emphasis)—and that is not a trivial implication—it established nothing else whatever.

Historically, not everyone agreed that Hume had fatally damaged the argument. It is simply not true that explanatory inferences cannot properly extend beyond merely what is required for known effects. As a very general example, based on the few observations which humans had made during a cosmically brief period in a spatially tiny part of the cosmos, Newton theorized that all bits of matter at all times and in all places attracted all other bits of matter. There was nothing logically suspect here. Indeed, simplicity and uniformity considerations—which have considerable well-earned scientific clout—push in the direction of such generalizations.

But Hume certainly identified important places within the argument to probe. Whether his suggestions are correct concerning the uncertain character of any designer inferred will depend upon the specific R s and upon what can or cannot be definitively said concerning requirements for their production.

2.1.2 R Concerns: Round 1

Key questions, then, include: what are the relevant R s typically cited? do those R s genuinely signal purpose and design? how does one show that either way? are there viable alternative accounts of the R s requiring no reference to minds? how does one show that either way? The specific R s in question are obviously central to design argument efforts. Although the underlying general category is, again, some special type of orderliness, the specifics have ranged rather widely historically. Among the more straightforwardly empirical are inter alia uniformity, contrivance, adjustment of means to ends, particularly exquisite complexity, particular types of functionality, delicacy, integration of natural laws, improbability, and the fitness (fine-tuning) of the inorganic realm for supporting life. Several problematic proposals that are empirically further removed and have axiological overtones have also been advanced, including the intelligibility of nature, the directionality of evolutionary processes, aesthetic characteristics (beauty, elegance, and the like), apparent purpose and value (including the aptness of our world for the existence of moral value and practice) and just the sheer niftiness of many of the things we find in nature.

But some advocates of design arguments had been reaching for a deeper intuition. The intuition they were attempting to capture involved properties that constituted some degree of evidence for design, not just because such properties happened to be often or even only produced by designing agents. Advocates were convinced that the appropriate R s in question were in their own right directly reflective of and redolent of cognition, that this directly suggested mind , that we could see nearly directly that they were the general sort of thing that a mind might or even would generate, and that consequently they did not depend for their evidential force upon previously established known instances of design. When we see a text version of the Gettysburg Address, that text says mind to us in a way totally unrelated to any induction or analogy from past encounters with written texts. It was that type of testimony to mind, to design, that some historical advocates of design arguments believed that they found in some R s observed in nature—a testimony having no dependency on induction or analogy. Beauty, purpose and in general value especially when conjoined with delicate complexity were popular underlying intuitive marks. Intricate, dynamic, stable, functioning order of the sort we encounter in nature was frequently placed in this category. Such order was taken to be suggestive of minds in that it seemed nearly self-evidently the sort of thing minds were prone to produce. It was a property whose mind-resonating character we could unhesitatingly attribute to intent.

Despite Hume’s earlier demurs that things in nature are not really very like artifacts such as machines, most people who are most familiar with nature’s dazzling intricacies freely admit that nature abounds with things that look designed—that are intention- shaped . For instance, Francis Crick (no fan of design) issued a warning to his fellow biologists:

Biologists must constantly keep in mind that what they see was not designed, but rather evolved. (Crick 1988, 138).

Along with this perception of mind-suggestiveness went a further principle—that the mind-suggestive or design- like characteristics in question were too palpable to have been generated by non-intentional means.

That allows specification of a second design inference pattern:

Schema 2 : Some things in nature (or nature itself, the cosmos) are design-like (exhibit a cognition-resonating, intention-shaped character R ) Design-like properties ( R ) are not producible by (unguided) natural means—i.e., any phenomenon exhibiting such R s must be a product of intentional design. Therefore Some things in nature (or nature itself, the cosmos) are products of intentional design, which in turn requires agency of some type.

Notice that explicit reference to human artifacts has dropped out of the argument, and that the argument is no longer comparative but has become essentially deductive. Some arguments were historically intended as arguments of that type. William Paley famously sees no design-like properties in a stone, but finding a watch on the ground would be another matter. Unlike the stone, it could not always have been there. Why not? [ 1 ]

For this reason, and for no other, namely, that when we come to inspect the watch, we perceive—what we could not discover in the stone—that its several parts are framed and put together for a purpose … [The requisite] mechanism being observed … the inference we think is inevitable, that the watch must have had a maker. ... Every observation which was made in our first chapter concerning the watch may be repeated with strict propriety concerning the eye, concerning animals, concerning plants, concerning, indeed, all the organized parts of the works of nature. … [T]he eye … would be alone sufficient to support the conclusion which we draw from it, as to the necessity of an intelligent Creator. …

Although Paley’s argument is routinely construed as analogical, it contains an informal statement of the above variant argument type. Paley goes on for two chapters discussing the watch, discussing the properties in it which evince design, and destroying potential objections to concluding design in the watch. It is only then that entities in nature—e.g., the eye—come onto the horizon. Paley isn’t trying to persuade his readers that the watch is designed and has a designer. He is teasing out the bases and procedures from and by which we should and should not reason about design and designers. Thus Paley’s use of the term ‘inference’ in connection with the watch’s designer. [ 2 ]

Once having acquired the relevant principles, then in Chapter 3 of Natural Theology —“Application of the Argument”—Paley applies the same argument (vs. presenting us with the other half of the analogical argument) to things in nature. The cases of human artifacts and nature represent two separate inference instances:

up to the limit, the reasoning is as clear and certain in the one case as in the other. (Paley 1802 [1963], 14)

But the instances are instances of the same inferential move:

there is precisely the same proof that the eye was made for vision as there is that the telescope was made for assisting it. (Paley 1802 [1963], 13)

The watch does play an obvious and crucial role—but as a paradigmatic instance of design inferences rather than as the analogical foundation for an inferential comparison.

Schema 2, not being analogically structured, would not be vulnerable to the ills of analogy, [ 3 ] and not being inductive would claim more than mere probability for its conclusion. That is not accidental. Indeed, it has been argued that Paley was aware of Hume’s earlier attacks on analogical design arguments, and deliberately structured his argument to avoid the relevant pitfalls (Gillispie 1990, 214–229).

2.2.1 Assessing the Schema 2 argument

First, how are we to assess the premises required by this schema? Premise (5), at least, is not particularly controversial even now. Crick’s earlier warning to biologists would have been pointless were there no temptation toward design attributions, and even as implacable a contemporary opponent of design arguments as Richard Dawkins characterized biology as:

the study of complicated things that give the appearance of having been designed for a purpose. (Dawkins 1987, 1)

Day-to-day contemporary biology is rife with terms like ‘design’, ‘machine’, ‘purpose’, and allied terms. As historian of science Timothy Lenoir has remarked:

Teleological thinking has been steadfastly resisted by modern biology. And yet, in nearly every area of research biologists are hard pressed to find language that does not impute purposiveness to living forms. (Lenoir 1982, ix)

Whether or not particular biological phenomena are designed, they are frequently enough design- like to make design language not only fit living systems extraordinarily well, but to undergird the generation of fruitful theoretical conceptions. [ 4 ] Advocates of design arguments claim that the reason why theorizing as if organisms are designed meets with such success is that organisms are in fact designed. Those opposed would say that all teleological concepts in biology must, in one way or another, be reduced to natural selection.

However, principle (6) (that the relevant design-like properties are not producible by unguided natural means) will be more problematic in evolutionary biology. What might be the rational justification for (6)? There are two broad possibilities.

1. Empirical: induction . Induction essentially involves establishing that some principle holds within the realm of our experience (the sample cases), and then generalizing the principle to encompass relevant areas beyond that realm (the test cases). The attempt to establish the universality of a connection between having relevant R s and being a product of mind on the basis of an observed consistent connection between having relevant R s and being a product of mind within all (most) of the cases where both R was exhibited and we knew whether or not the phenomenon in question was a product of mind, would constitute an inductive generalization.

This approach would suffer from a variety of weaknesses. The R -exhibiting things concerning which we knew whether they were designed would be almost without exception human artifacts, whereas the phenomena to which the generalization was being extended would be things in nature. And, of course, the generalization in question could establish at best a probability, and a fairly modest one at that.

2. Conceptual . It might be held that (6) is known in the same conceptual, nearly a priori way in which we know that textbooks are not producible by natural processes unaided by mind. And our conviction here is not based on any mere induction from prior experiences of texts. Texts carry with them essential marks of mind, and indeed in understanding a text we see at least partway into the mind(s) involved. Various alien artifacts (if any)—of which we have had no prior experience whatever—could fall into this category as well. Similarly, it has been held that we sometimes immediately recognize that order of the requisite sort just is a sign of mind and intent.

Alternatively, it could be argued that although there is a genuine conceptual link between appropriate R s and mind, design, intent, etc., that typically our recognition of that link is triggered by specific experiences with artifacts. On this view, once the truth of (6) became manifest to us through those experiences, the appropriateness of its more general application would be clear. That might explain why so many advocates of design arguments seem to believe that they must only display a few cases and raise their eyebrows to gain assent to design.

Either way, principle (6), or something like it, would be something with which relevant design inferences would begin. Further investigation of (6) requires taking a closer look at the R s which (6) involves.

2.2.2 R Concerns: Round 2

One thing complicating general assessments of design arguments is that the evidential force of specific R s is affected by the context of their occurrence. Specifically, properties which seem to constitute marks of design in known artifacts often seem to have significantly less evidential import outside that context. For instance, we typically construe enormous complexity in something known to be a manufactured artifact as a deliberately intended and produced characteristic. But mere complexity in contexts not taken to involve artifacts (the precise arrangement of pine needles on a forest floor, for instance) does not seem to have that same force. In the case of natural objects, it is less clear that such complexity—as well as the other traditional empirical R s—bespeaks intention, plan, and purpose.

Furthermore, even within those two contexts—artifact and nature—the various R s exhibit varying degrees of evidential force. For instance, even in an artifact, mere complexity of whatever degree speaks less clearly of intent than does an engraved sentence. As most critics of design arguments point out, the examples found in nature are not of the “engraved sentence” sort.

2.2.3 Gaps and Their Discontents

Evidential ambiguity would virtually disappear if it became clear that there is no plausible means of producing some R independent of deliberate intent. Part of the persuasiveness of (6) historically came from absence of any known plausible non-intentional alternative causal account of the traditional R s. Such cases are often linked to alleged gaps in nature—phenomena for which, it is claimed, there can be no purely natural explanation, there being a gap between nature’s production capabilities and the phenomenon in question. (For example, nature’s unaided capabilities fall short of producing a radio.) Design cases resting upon nature’s alleged inability to produce some relevant ‘natural’ phenomenon are generally assumed to explicitly or implicitly appeal to supernatural agency, and are typically described as “God-of-the-gaps” arguments—a description usually intended to be pejorative.

The position that there are causal gaps in nature is not inherently irrational—and would seem to be a legitimate empirical question. But although gaps would profoundly strengthen design arguments, they have their suite of difficulties. Gaps are usually easy to spot in cases of artifactuality (take the radio example, again), but although they may be present in nature, establishing their existence there can usually be done (by science, at least) only indirectly—via probability considerations, purported limitations on nature’s abilities, etc.

Several possible snags lurk. Gaps in nature would, again, suggest supernatural agency, and some take science to operate under an obligatory exclusion of such. This prohibition—commonly known as methodological naturalism —is often claimed (mistakenly, some argue) to be definitive of genuine science. [ 5 ] While such “established” limitations on science have been overturned in the past, the spotty track record of alleged gaps provides at least a cautionary note. Purported gaps have been closed by new scientific theories postulating means of natural production of phenomena previously thought to be beyond nature’s capabilities. The most obvious example of that is, of course, Darwin’s evolutionary theory and its descendants.

Some philosophers of science claim that in a wide variety of scientific cases we employ an “inference to the best explanation” (IBE). [ 6 ] The basic idea is that if one among a number of competing candidate explanations is overall superior to others in significant respects—enhanced likelihood, explanatory power and scope, causal adequacy, plausibility, evidential support, fit with already-accepted theories, predictiveness, fruitfulness, precision, unifying power, and the like—then we are warranted in (provisionally) accepting that candidate as the right explanation given the evidence in question (Lipton 1991, 58). Some advocates see design arguments as inferences to the best explanation, taking design explanations—whatever their weaknesses—as prima facie superior to chance, necessity, chance-driven evolution, or whatever.

A general schema deployed in the current case would give us the following:

Schema 3 : Some things in nature (or nature itself, the cosmos) exhibit exquisite complexity, delicate adjustment of means to ends (and other relevant R characteristics). The hypothesis that those characteristics are products of deliberate, intentional design (Design Hypothesis) would adequately explain them. In fact, the hypothesis that those characteristics are products of deliberate, intentional design (Design Hypothesis) is the best available overall explanation of them. Therefore (probably) Some things in nature (or nature itself, the cosmos) are products of deliberate, intentional design (i.e., the Design Hypothesis is likely true).

In arguments of this type, superior explanatory virtues of a theory are taken as constituting epistemic support for the acceptability of the theory or for the likely truth of the theory.

There are, of course, multitudes of purported explanatory, epistemic virtues, including the incomplete list a couple of paragraphs back. Assessing hypotheses in terms of such virtues is often contentious. The assessment of ‘best’ is not only a value-tinged judgment, but is notoriously tricky (especially given the ambiguous and hard to pinpoint import of the R s in the present case). There is also the very deep question of why we should think that features which we humans find attractive in proposed explanations should be thought to be truth-tracking. What sort of justification might be available here?

2.3.1 IBE, Likelihood and Bayes

One key underlying structure in this context is typically traced to Peirce’s notion of abduction . Suppose that some otherwise surprising fact e would be a reasonably expectable occurrence were hypothesis h true. That, Peirce argued, would constitute at least some provisional reason for thinking that h might be true. Peirce’s own characterization was as follows (Peirce 1955, 151):

Schema 3: The surprising fact, C , is observed. But if A were true, C would be a matter of course. Hence, There is reason to suspect that A is true.

The measure of C being a ‘matter of course’ given A is frequently described as the degree to which C could be expected were A true. This intuition is sometimes—though explicitly not by Peirce himself—formalized in terms of likelihood, defined as follows:

The likelihood of hypothesis \(h\) (given evidence \(e\)) = \(P(e | h)\)

The likelihood of h is the probability of finding evidence e given that the hypothesis h is true. In cases of competing explanatory hypotheses \(h_1\) and \(h_2\) the comparative likelihoods on specified evidence can be taken to indicate which of the competitors the evidence better supports, i.e.:

Likelihood Principle : \( [P(e| h_1) > P(e|h_2)] \leftrightarrow\) (\(e\) supports \(h_1\) more than it does \(h_2\))

Higher likelihood of \(h_1\) than \(h_2\) on specific evidence does not automatically imply that \(h_1\) is likely to be true, or is better in some overall sense than is \(h_2\). \(h_1\) might, in fact, be a completely lunatic theory which nonetheless entails e , giving \(h_1\) as high a likelihood as possible. Such maximal likelihood relative to e would not necessarily alter \(h_1\)’s lunacy. Likelihood thus does not automatically translate into a measure of how strongly some specific evidence e supports the hypothesis \(h_1\) in question (Jantzen 2014a, Chap. 11). Given this limitation, some argue that only a full-blown Bayesian analysis should be used to assess competing hypotheses. For a contrast between IBE and Bayesianism, see the entry on abduction . For an important recent critique of theistic design arguments in Bayesian terms, see Sober 2009, the reply in Kotzen 2012, and the response in Jantzen 2014b.

There are other potential issues here as well. Sober argues that without additional very specific assumptions about the putative designer we could specify no particular value for \(P(e | h)\) — e.g., the likelihood that a designer would produce vertebrate eyes with the specific features we observe. Depending on the specific assumptions made we could come up with any value from 0 to 1 (e.g., Sober 2003, 38). Jantzen argues that the same problem applies to any likelihood involving chance as an alternative hypothesis to design. Chance is equally vague. If we allow the naturalist to fill it out in ways that reduce this vagueness, the design theorist should be accorded the same benefit (Jantzen 2014a, 184).

There is also the potential problem of new, previously unconsidered hypotheses all lumped together in the catch-all basket. Without knowing the details of what specific unconsidered hypotheses might look like, there is simply no plausible way to anticipate the apparent likelihood of a novel new hypothesis. This, on some views, is essentially what happened with traditional design arguments. Such arguments were the most reasonable available until Darwinian evolution provided a plausible (or better) alternative the details and likelihood of which were not previously anticipatable. We should note that the problem of unconsidered hypotheses is an issue for all likelihood arguments, not merely those involving design.

3. Alternative Explanation

Without going into the familiar details, Darwinian processes fueled by undesigned, unplanned, chance variations that are in turn conserved or eliminated by way of natural selection would, it is argued, over time produce organisms exquisitely adapted to their environmental niches. [ 7 ] And since many of the characteristics traditionally cited as evidences of design just were various adaptations, evolution would thus produce entities exactly fitting traditional criteria of design. Natural selection, then, unaided by intention or intervention could account for the existence of many (perhaps all) of the R s which we find in biology. (A parallel debate can be found between those who believe that life itself requires a design explanation (Meyer 2009) and those proposing naturalistic explanations; see the entry on life .)

That was—and is—widely taken as meaning that design arguments depending upon specific biological gaps would be weakened—perhaps fatally.

Premise (10)—not to mention the earlier (6)—would thus look to simply be false. What had earlier appeared to be purpose (requiring intent) was now apparently revealed as mere unintended but successful and preserved function .

Of course, relevant premises being false merely undercuts the relevant schemas in present form—it does not necessarily refute either the basic design intuition or other forms of design arguments. But some critics take a much stronger line here. Richard Dawkins, for instance, subtitles one of his books: “Why the evidence of evolution reveals a universe without design” (Dawkins, 1987). Typically underlying claims of this sort is the belief that Darwinian evolution, by providing a relevant account of the origin and development of adaptation, diversity, and the like, has explained away the alleged design in the biological realm—and an attendant designer—in much the same way that kinetic theory has explained away caloric. Indeed, this is a dominant idea underlying current responses to design arguments. However, undercutting and explaining away are not necessarily the same thing, and exactly what explaining away might mean, and what a successful explaining away might require are typically not clearly specified. So before continuing, we need clarity concerning some relevant conceptual landscape.

3.1 Explaining Away [ 8 ]

That an alleged explanatory factor α is provisionally explained away requires that there be an alternative explanation Σ meeting these conditions:

  • Σ is explanatorily adequate to the relevant phenomenon (structure, property, entity, event)
  • Σ can be rationally supported in terms of available (or likely) evidence
  • Σ is relevantly superior to the original in terms either of adequacy or support
  • Σ requires no essential reference to α

However, (a) – (d) are incomplete in a way directly relevant to the present discussion. Here is a very simple case. Suppose that an elderly uncle dies in suspicious circumstances, and a number of the relatives believe that the correct explanation is the direct agency of a niece who is primary heir, via deliberately and directly administering poison. However, forensic investigation establishes that the cause of death was a mix-up among medications the uncle was taking—an unfortunate confusion. The suspicious relatives, however, without missing an explanatory beat shift the niece’s agency back one level, proposing that the mix-up itself was orchestrated by the niece—switching contents of prescription bottles, no doubt. And that might very well turn out to be the truth.

In that sort of case, the α in question (e.g., niecely agency) is no longer directly appealed to in the relevant initial explanatory level, but is not removed from all explanatory relevance to the phenomenon in question. In general, then, for α to be explained away in the sense of banished from all explanatory relevance the following condition must also be met:

  • no reference to α is required at any explanatory level underlying Σ

Roughly this means that Σ does not depend essentially on any part β of any prior explanation where α is essential to β. There are some additional possible technical qualifications required, but the general intuition should be clear.

Thus, e.g., whereas there was no need to appeal to caloric at some prior or deeper level, with design, according to various design advocates, there is still an explanatory lacuna (or implicit promissory note) requiring reference to design at some explanatory level prior to Darwinian evolution. Indeed, as some see it (and as Paley himself suggested), there are phenomena requiring explanation in design terms which cannot be explained away at any prior explanatory level (short of the ultimate level).

That some phenomenon α has been explained away can be taken to mean two very different things—either as

  • showing that it is no longer rational to believe that α exists
  • showing that α does not exist

(And often, of course, both.)

For instance, few would assert that there is still an extant rational case for belief in phlogiston—any explanatory work it did at the proximate level seems to have ceased, and deeper explanatory uses for it have never subsequently materialized. Perhaps its non-existence was not positively established immediately, but removal of rational justification for belief in some entity can morph into a case for non-existence as the evidence for a rival hypothesis increases over time.

3.1.1 Level-shifting

Purported explanations can be informally divided into two broad categories—those involving agents, agency, intention, and the like; and those involving mechanism, physical causality, natural processes, and the like. The distinction is not clean (functioning artifacts typically involve both), but is useful enough in a rough and ready way, and in what follows agent explanations and mechanical explanations respectively will be used as convenient handles. Nothing pernicious is built into either the broad distinction or the specified terminology.

There are some instructive patterns that emerge in explanatory level-shifting attempts, and in what immediately follows some of the more basic patterns will be identified.

(a) Agent explanations

Intention, intervention, and other agency components of explanations can very frequently be pushed back to prior levels—much as many defenders of teleological arguments claim. The earlier case of the alleged poisoning of the rich uncle by the niece is a simple example of this.

But in some cases, the specifics of the agent explanation in question may appeal to some prior level less plausible or sensible. For example, suppose that one held the view that crop circles were to be explained in terms of direct alien activity. One could, upon getting irrefutable video proof of human production of crop circles, still maintain that aliens were from a distance controlling the brains of the humans in question, and that thus the responsibility for crop circles did still lie with alien activity. While this retreat of levels preserves the basic explanation, it of course comes with a significant cost in inherent implausibility.

Still the level-changing possibility is as a general rule available with proposed agent explanations.

(b) Mechanical explanations

Pushing specific explanatory factors back to a prior level often works less smoothly in cases of purely mechanical/physical explanations than in intentional/agency explanations. In many attempted mechanistic relocation cases, it is difficult to see how the specific relocated explanatory factor is even supposed to work, much less generate any new explanatory traction. Exactly what would caloric do if pushed back one level, for instance?

Although level shifting of specific explanatory factors seems to work less easily within purely physical explanations, relocation attempts involving broad physical principles can sometimes avoid such difficulties. For instance, for centuries determinism was a basic background component of scientific explanations (apparently stochastic processes being explained away epistemically). Then, early in the 20th century physics was largely converted to a quantum mechanical picture of nature as involving an irreducible indeterminism at a fundamental level—apparently deterministic phenomena now being what was explained away. However, DeBroglie, Bohm and others (even for a time Einstein) tried to reinstate determinism by moving it back to an even deeper fundamental level via hidden variable theories. Although the hidden variable attempt is generally thought not to be successful, its failure is not a failure of principle.

3.1.2 Possible disputes

How one assesses the legitimacy, plausibility, or likelihood of the specific counter-explanation will bear substantial weight here, and that in turn will depend significantly on among other things background beliefs, commitments, metaphysical dispositions, and the like. If one has a prior commitment to some key α (e.g., to theism, atheism, naturalism, determinism, materialism, or teleology), or assigns a high prior to that α, the plausibility of taking the proposed (new) explanation as undercutting, defeating, or refuting α (and/or Σ) will be deeply affected, at least initially.

Tilting the conceptual landscape via prior commitments is both an equal opportunity epistemic necessity and a potential pitfall here. Insisting on pushing an explanatory factor back a level is often an indication of a strong prior commitment of some sort. Disagreement over deeper philosophical or other principles will frequently generate divergence over when something has or has not been explained away. One side, committed to the principle, will accept a level change as embodying a deeper insight into the relevant phenomenon. The other, rejecting the principle, will see an ad hoc retreat to defend an α which has in fact been explained away.

Returning to the present issue, design argument advocates will of course reject the claim that design, teleology, agency and the like have been explained away either by science generally or by Darwinian evolution in particular. Reasons will vary. Some will see any science—Darwinian evolution included—as incompetent to say anything of ultimate design relevance, pro or con. (Many on both sides of the design issue fit here.) Some will see Darwinian evolution as failing condition (a), (b) and/or (c), claiming that Darwinian evolution is not explanatorily adequate to selected α’s, is inadequately supported by the evidence, and is far from superior to agency explanations of relevant phenomena. (Creationists and some—not all—‘intelligent design’ advocates fit here.) Some will argue that a Darwinian failure occurs at (d), citing e.g., a concept of information claimed to be both essential to evolution and freighted with agency. (Some intelligent design advocates fit here; see Dembski 2002 and Meyer 1998.) However, the major contention of present interest involves (e).

Historically, design cases were in fact widely understood to allow for indirect intelligent agent design and causation, the very causal structures producing the relevant phenomena being themselves deliberately designed for the purpose of producing those phenomena. [ 9 ] For instance, it was typically believed that God could have initiated special conditions and processes at the instant of creation which operating entirely on their own could produce organisms and other intended (and designed) results with no subsequent agent intervention required. Paley himself, the authors of the Bridgewater Treatises and others were explicitly clear that whether or not something was designed was an issue largely separable from the means of production in question. Historically it was insisted that design in nature did track back eventually to intelligent agency somewhere and that any design we find in nature would not—and could not—have been there had there ultimately been no mind involved. But commentators (including many scientists) at least from the early 17th century on (e.g., Francis Bacon and Robert Boyle) very clearly distinguished the creative initiating of nature itself from interventions within the path of nature once initiated. For instance, over two centuries before Darwin, Bacon wrote:

God … doth accomplish and fulfill his divine will [by ways] not immediate and direct, but by compass; not violating Nature, which is his own law upon the creation. (Quoted in Whewell 1834, 358)

Indeed, if the R s in question did directly indicate the influence of a mind, then means of production—whether unbroken causation or gappy—would be of minimal evidential importance. Thus, the frequent contemporary claim that design arguments all involve appeal to special divine intervention during the course of nature’s history—that in short design arguments are “God-of-the-gaps” arguments—represents serious historical (and present) inaccuracy.

However, if R s result from gapless chains of natural causal processes, the evidential impact of those R s again threatens to become problematic and ambiguous, since there will a fortiori be at the immediate level a full natural causal account for them. [ 10 ] Design will, in such cases, play no immediate mechanistic explanatory role, suggesting its superfluousness. But even if such conceptions were explanatorily and scientifically superfluous at that level, that does not entail that they are conceptually, alethically, inferential, or otherwise superfluous in general. The role of mind might be indirect, deeply buried, or at several levels of remove from the immediate production mechanism but would still have to be present at some level. In short, on the above picture Darwinian evolution will not meet condition (e) for explaining away design, which is not itself a shortcoming of Darwinian evolution.

But any gap-free argument will depend crucially upon the R s in question being ultimately dependent for their eventual occurrence upon agent activity. That issue could be integrated back into an altered Schema 2 by replacing (6) with:

(6a) Design-like properties ( R ) are (most probably) not producible by means ultimately devoid of mind/intention—i.e., any phenomena exhibiting such R s must be a product (at least indirectly) of intentional design.

The focus must now become whether or not the laws and conditions required for the indirect production of life, intelligent life, etc., could themselves be independent of intention, design, and mind at some deep (perhaps primordial, pre-cosmic) point. In recent decades, exactly that question has arisen increasingly insistently from within the scientific community.

4. Further Contemporary Design Discussions

4.1 cosmological: fine-tuning.

Intuitively, if the laws of physics were different, the evolution of life would not have taken the same path. If gravity were stronger, for example, then flying insects and giraffes would most likely not exist. Contemporary physics provides significantly more drama to the story. Even an extraordinarily small change in one of many key parameters in the laws of physics would have made life impossible anywhere in the universe. Consider two examples:

The expansion rate of the universe is represented by the cosmological constant Λ. If Λ were slighter greater, there would be no energy sources, such as stars. If it were slightly less, the Big Bang would have quickly led to a Big Crunch in which the universe collapsed back onto itself. For life to be possible, Λ cannot vary more than one part in 10 53 (Collins 2003)

Life depends on, among other things, a balance of carbon and oxygen in the universe. If the strong nuclear force were different by 0.4%, there would not be enough of one or the other for life to exist (Oberhummer, Csótó, and Schlattl 2000). Varying this constant either way “would destroy almost all carbon or almost all oxygen in every star” (Barrow 2002, 155).

Many examples of fine-tuning have to do with star formation. Stars are important since life requires a variety of elements: oxygen, carbon, hydrogen, nitrogen, calcium, and phosphorus. Stars contain the only known mechanism for producing large quantities of these elements and are therefore necessary for life. Lee Smolin estimates that when all of the fine-tuning examples are considered, the chance of stars existing in the universe is 1 in 10 229 . “In my opinion, a probability this tiny is not something we can let go unexplained. Luck will certainly not do here; we need some rational explanation of how something this unlikely turned out to be the case” (Smolin 1999, 45). Smolin is not merely claiming that all improbable events require an explanation, but some improbable events are special. (In poker, every set of five cards dealt to the dealer has the same probability, assuming that the cards are shuffled sufficiently. If the dealer is dealt a pair on three successive hands, no special explanation is required. If the dealer is dealt a royal flush on three successive hands, an explanation would rightly be demanded, and the improbability of this case isn’t even close to the magnitude of the improbability that Smolin mentioned.) Physicists who have written on fine-tuning agree with Smolin that it cries out for an explanation. (What physicists in general think is, of course, harder to determine.) One explanation is that the universe appears to be fine-tuned for the existence of life because it literally has been constructed for life by an intelligent agent.

There are two other types of responses to fine-tuning: (i) it does not, in fact, require a special explanation, and (ii) there are alternative explanations to theistic design. Let’s briefly consider these (also see the entry on fine-tuning ).

Three approaches have been taken to undermine the demand for explanation presented by fine-tuning.

4.1.1.1 Weak anthropic principle

In a sense, it is necessary for the fine-tuned constants to have values in the life-permitting range: If those values were not within that range, people would not exist. The fine-tuned constants must take on the values that they have in order for scientists to be surprised by their discovery in the first place. They could not have discovered anything else. According to the weak anthropic principle, we ought not to be surprised by having made such a discovery, since no other observation was possible. But if we should not have been surprised to have made such a discovery, then there is nothing unusual here that requires a special explanation. The demand for explanation is simply misplaced.

4.1.1.2 Observational selection effect

Sober gives a related but stronger argument based on observational selection effects (Sober 2009, 77–80). Say that Jones nets a large number of fish from a local lake, all of which are over 10 inches long. Let \(h_{\textrm{all}}\) = ‘all of the fish in the lake are over 10 inches long’ and \(h_{1/2}\) = ‘Half of the fish in the lake are over 10 inches long’. The evidence \(e\) is such that \(P(e|h_{\textrm{all}}) > P(e | h_{1/2})\). Now say that Jones discovers that his net is covered with 10 inch holes, preventing him from capturing any smaller fish. In that case, \(e\) does not favor one hypothesis over the other. The evidence \(e\) is an artifact of the net itself, not a random sample of the fish in the lake.

When it comes to fine-tuning, Sober considers \(h_{\textrm{design}}\) = ‘the constants have been set in place by an intelligence, specifically God’, and \(h_{\textrm{chance}}\) = ‘the constants are what they are as a matter of mindless random chance’. While intuitively

\(P(\textrm{constants are just right for life} | h_{\textrm{design}}) >\) \(\;\; P(\textrm{constants are just right for life} | h_{\textrm{chance}}) \)

one has to consider the role of the observer, who is analogous to the net in the fishing example. Since human observers could only detect constants in the life-permitting range, Sober argues, the correct probabilities are

\( P(\textrm{you observe that the constants are right} | h_{\textrm{design}} \amp \textrm{you exist}) =\) \(\;\; P(\textrm{you observe that the constants are right} | h_{\textrm{chance}} \amp \textrm{you exist}) \)

Given this equality, fine-tuning does not favor \(h_{\textrm{design}}\) over \(h_{\textrm{chance}}\). The selection effect prevents any confirmation of design.

Sober’s analysis is critiqued in Monton 2006 and Kotzen 2012. Also see Jantzen 2014a (sec. 18.4). We should note that if Sober is correct, then the naturalistic explanations for fine-tuning considered below (4.1.2) are likewise misguided.

4.1.1.3 Probabilities do not apply

Let C stand for a fine-tuned parameter with physically possible values in the range [0, ∞). If we assume that nature is not biased toward one value of C rather than another, then each unit subinterval in this range should be assigned equal probability. Fine-tuning is surprising insofar as the life-permitting range of C is tiny compared to the full interval, which corresponds to a very small probability.

As McGrew, McGrew & Vestrup argue (2001), there is a problem here in that, strictly speaking, mathematical probabilities do not apply in these circumstances. When a probability distribution is defined over a space of possible outcomes, it must add up to exactly 1. But for any uniform distribution over an infinitely large space, the sum of the probabilities will grow arbitrarily large as each unit interval is added up. Since the range of C is infinite, McGrew et al . conclude that there is no sense in which life-friendly universes are improbable; the probabilities are mathematically undefined.

One solution to this problem is to truncate the interval of possible values. Instead of allowing C to range from [0, ∞), one could form a finite interval [0, N ], where N is very large relative to the life-permitting range of C . A probability distribution could then be defined over the truncated range.

A more rigorous account employs measure theory. Measure is sometimes used in physics as a surrogate for probability. For example, there are many more irrational numbers than rational ones. In measure theoretic terms, almost all real numbers are irrational, where “almost all” means all but a set of zero measure. In physics, a property found for almost all of the solutions to an equation requires no explanation; it’s what one should expect. It’s not unusual, for instance, for a pin balancing on its tip to fall over. Falling over is to be expected. In contrast, if a property that has zero measure in the relevant space were actually observed to be the case, like the pin continuing to balance on its tip, that would demand a special explanation (Earman 1987, 315). Assuming one’s model for the system is correct, nature appears to be strongly biased against such behavior (Gibbons, Hawking & Stewart 1987, 736). The argument for fine-tuning can thus be recast such that almost all values of C are outside of the life-permitting range. The fact that our universe is life-permitting is therefore in need of explanation.

The question of whether probabilities either do not apply or have been improperly applied to cosmological fine-tuning continues to draw interest. For more, see Davies 1992, Callender 2004, Holder 2004, Koperski 2005, Manson 2009, Jantzen 2014a (sec. 18.3), and Sober 2019 (sec. 5.1). Manson (2018) argues that neither theism nor naturalism provides a better explanation for fine-tuning.

Assuming that fine-tuning does require an explanation, there are several approaches one might take (Koperski 2015, section 2.4).

4.1.2.1 Scientific progress

That the universe is fine-tuned for life is based on current science. Just as current science explained or explained away many past anomalies, future science may explain or explain away fine-tuning. Science may one day find a naturalistic answer, eliminating the need for design. For suggestions along these lines, see Harnik, Kribs & Perez 2006, and Loeb 2014.

While this is a popular stance, it is a promissory note rather than an explanation. The appeal to what might yet be discovered is not itself a rival hypothesis.

4.1.2.2 Exotic life

It’s conceivable that life could exist in a universe with parameter values that we do not typically believe are life-permitting. In other words, there may be exotic forms of life that could survive in a very different sort of universe. If so, then perhaps the parameter intervals that are in fact life-permitting are not fine-tuned after all.

The main difficulty with this suggestion is that all life requires a means for overcoming the second law of thermodynamics. Life requires the extraction of energy from the environment. Any life-form imaginable must therefore have systems that allow for something like metabolism and respiration, which in turn require a minimal amount of complexity (e.g., there can be no single-molecule life forms). Many examples of fine-tuning do not allow for such complexity, however. If there were no stars, for example, then there would be no stable sources of energy and no mechanism for producing the heavier elements in the periodic table. Such a universe would lack the chemical building blocks needed for a living entity to extract energy from the environment and thereby resist the pull of entropy.

4.1.2.3 Multiverse

While the odds of winning a national lottery are low, your odds would obviously increase if you were to buy several million tickets. The same idea applies to the most popular explanation for fine-tuning: a multiverse. Perhaps physical reality consists of a massive array of universes each with a different set of values for the relevant constants. If there are many—perhaps infinitely many—universes, then the odds of a life-permitting universe being produced would seem to be much greater. While most of the universes in the multiverse would be unfit for life, so the argument goes, ours is one of the few where all of the constants have the required values.

While the philosophical literature on the multiverse continues to grow (see Collins 2009, 2012, and Kraay 2014), many of the arguments against it share a common premise: a multiverse would not, by itself, be a sufficient explanation of fine-tuning. More would have to be known about the way in which universes are produced. By analogy, just because a roulette wheel has 38 spaces does not guarantee that the probability of Red 25 is 1/38. If the wheel is rigged in some way—by using magnets for example—to prevent that outcome, then the probability might be extremely small. If the table were rigged and yet Red 25 was the actual winner, that would require a special explanation. Likewise, one would have to know whether universe production within the multiverse is biased for, against, or is indifferent to the life-permitting values of the relevant constants. Depending on the outcome, the existence of a multiverse might or might not explain fine-tuning.

A more pressing problem would be the proliferation of “Boltzmann brains” (BB) in a multiverse (Carroll 2020). BBs are intelligent beings produced by quantum fluctuations. The beliefs of such nonstandard observers are formed by random processes and therefore lack rational support. Since there would be many orders of magnitude more BBs than standard observers in a multiverse, it is statistically likely that you are not, in fact, a standard observer after all. This fact serves as a defeater for most of your beliefs, including your belief in a multiverse. Belief in a multiverse is therefore self-defeating.

4.2 Biological: Intelligent Design

A high-profile development in design arguments over the past 20 years or so involves what has come to be known as Intelligent Design (ID). Although there are variants, it generally involves efforts to construct design arguments taking cognizance of various contemporary scientific developments (primarily in biology, biochemistry, and cosmology)—developments which, as most ID advocates see it, both reveal the inadequacy of mainstream explanatory accounts (condition (a)) and offer compelling evidence for design in nature at some level (condition (e) again).

ID advocates propose two specialized R s— irreducible complexity (Behe 1996) and specified complex information (Dembski 1998, 2002). [ 11 ] Although distinctions are sometimes blurred and while ID arguments involving each of those R s tend to be gap arguments, an additional focus on mind-reflective aspects of nature is typically more visible in ID arguments citing specified complexity than in arguments citing irreducible complexity.

The movement has elicited vociferous criticism and opposition (Pennock 2000). Opponents have pressed a number of objections against ID including, inter alia contentions that ID advocates have simply gotten the relevant science wrong, that even where the science is right the empirical evidences cited by design advocates do not constitute substantive grounds for design conclusions, that the existence of demonstrably superior alternative explanations for the phenomena cited undercuts the cogency of ID cases, and that design theories are not legitimate science , but are just disguised creationism, God-of-the-gaps arguments, religiously motivated, etc.

We will not pursue that dispute here except to note that even if the case is made that ID could not count as proper science, which is controversial, [ 12 ] that would not in itself demonstrate a defect in design arguments as such. Science need not be seen as exhausting the space of legitimate conclusions from empirical data. In any case, the floods of vitriol flowing from both sides in the current ID discussion suggest that much more than the propriety of selected inferences from particular empirical evidences is at issue.

That question is: why do design arguments remain so durable if empirical evidence is inferentially ambiguous, the arguments logically controversial, and the conclusions vociferously disputed? One possibility is that they really are better arguments than most philosophical critics concede. Another possibility is that design intuitions do not rest upon inferences at all. The situation may parallel that of the existence of an external world, the existence of other minds, and a number of other familiar matters. The 18th century Scottish Common Sense philosopher Thomas Reid (and his contemporary followers) argued that we are simply so constructed that in certain normally-realized experiential circumstances we simply find that we in fact have involuntary convictions about such a world, about other minds, and so forth. That would explain why historical philosophical attempts to reconstruct the arguments by which such beliefs either arose or were justified were such notorious failures—failures in the face of which ordinary belief nonetheless proceeded happily and helplessly onward. If a similar involuntary belief-producing mechanism operated with respect to intuitions of design, that would similarly explain why argumentative attempts have been less than universally compelling but yet why design ideas fail to disappear despite the purported failure of such arguments.

A number of prominent figures historically in fact held that we could determine more or less perceptually that various things in nature were candidates for design attributions—that they were in the requisite respects design- like . Some, like William Whewell, held that we could perceptually identify some things as more than mere candidates for design (1834, 344). Thomas Reid also held a view in this region, [ 13 ] as did Hume’s Cleanthes, and, more recently, Alvin Plantinga (2011, 263–264).

If something like that were the operative process, then ID, in trying to forge a scientific link to design in the sense of inferences from empirically determined evidences would be misconstructing the actual basis for design belief, as would be design arguments more generally. It is perhaps telling, in this regard, that scientific theorizing typically involves substantial creativity and that the resultant theories are typically novel and unexpected. Design intuitions, however, do not seem to emerge as novel construals from creative grappling with data, but are embedded in our thinking nearly naturally—so much so that, again, Crick thinks that biologists have to be immunized against it.

Perception and appreciation of the incredible intricacy and the beauty of things in nature—whether biological or cosmic—has certainly inclined many toward thoughts of purpose and design in nature, and has constituted important moments of affirmation for those who already accept design positions. Regardless of what one thinks of the arguments at this point, so long as nature has the power to move us (as even Kant admitted that the ‘starry heavens above’ did), design convictions and arguments are unlikely to disappear quietly.

  • Babbage, Charles, 1838. Ninth Bridgewater Treatise: A Fragment , London: J. Murray.
  • Barrow, John D. 2002. The Constants of Nature , New York: Pantheon Books.
  • Behe, Michael, 1996. Darwin’s Black Box , New York: Free Press.
  • Boyle, Robert, 1685–6. Free Inquiry into the Vulgarly Receiv’d Notions of Nature , in Hall 1965, pp. 150–153.
  • –––, 1688. A Disquisition about the Final Causes of Natural Things , London: John Taylor.
  • Broad, C.D., 1925. The Mind and its Place in Nature , London: Kegan Paul.
  • Callender, Craig, 2004. “Measures, Explanations and the Past: Should ‘Special’ Initial Conditions Be Explained,” British Journal for the Philosophy of Science , 55(2): 195–217.
  • Carroll, Sean M., 2020. “Why Boltzmann Brains Are Bad,” in Current Controversies in Philosophy of Science , edited by Shamik Dasgupta, Brad Weslake, and Ravit Dotan, 7–20. New York: Routledge.
  • Chesterton, G.K., 1908. “Ethics of Elfland,” in Orthodoxy , New York: John Lane, pp. 106–7.
  • Collins, Robin, 2003. “Evidence for Fine-Tuning,” in God and Design: The Teleological Argument and Modern Science , edited by Neil A. Manson, 178–99. New York: Routledge.
  • –––, 2009. “The Teleological Argument: An Exploration of the Fine-Tuning of the Universe,” in The Blackwell Companion to Natural Theology , edited by William Lane Craig and J.P. Moreland, 202–81. Chichester: Wiley-Blackwell.
  • –––, 2012. “Modern Cosmology and Anthropic Fine-Tuning: Three Approaches,” in Georges Lemaître: Life, Science and Legacy , edited by Rodney D. Holder and Simon Mitton. Berlin; Heidelberg: Springer.
  • Crick, Francis, 1988. What Mad Pursuit , New York: Basic.
  • Darwin, Charles, 1859 [1966]. On the Origin of Species , Facsimile first edition, Cambridge, MA: Harvard University Press.
  • Darwin, Charles, 1887. Life and Letters of Charles Darwin , Vol. 1, Francis Darwin (ed.), New York: D. Appleton.
  • –––, 1902 [1995]. The Life of Charles Darwin , Francis Darwin (ed.), London: Senate.
  • –––, 1987. Charles Darwin’s Notebooks, 1836–1844 . Transcribed and edited by Paul Barrett, Peter Gautrey, Sandra Herbert, Dave Kohn and Sydney Smith, Ithaca: Cornell.
  • Davies, Paul, 1992. The Mind of God , New York: Simon & Schuster.
  • –––, 1995. Are We Alone? , New York: Basic.
  • Dawkins, Richard, 1987. Blind Watchmaker , New York: Norton.
  • Dembski, William, 1998. The Design Inference , Cambridge: Cambridge University Press.
  • Earman, John. 1987. “The SAP Also Rises: A Critical Examination of the Anthropic Principle,” American Philosophical Quarterly , 24(4): 307–17.
  • Edwards, Jonathan, 1980. The Works of Jonathan Edwards Wallace Anderson (ed.), Volume 6: Scientific and Philosophical Writings , New Haven: Yale University Press.
  • Fitelson, Brandon, 2007. “Likelihood, Bayesianism, and Relational Confirmation,” Synthese , 156: 473–489.
  • Foster, John, 1982–3. “Induction, Explanation and Natural Necessity,” Proceedings of the Aristotelian Society , 83: 87–101.
  • Foster, John, 1985. A. J. Ayer , London: Routledge & Kegan Paul.
  • Gillispie, Neal C., 1990. “Divine Design and the Industrial Revolution: William Paley’s Abortive Reform of Natural Theology,” Isis , 81: 213–229.
  • Glass, Marvin and Julian Wolfe, 1986. “Paley’s Design Argument for God,” Sophia , 25(2): 17–19.
  • Gibbons, G. W., S. W. Hawking, and J. M. Stewart, 1987. “A Natural Measure on the Set of All Universes,” Nuclear Physics B , 281(3–4): 736–51.
  • Hall, Marie Boas, 1965. Robert Boyle on Natural Philosophy , Bloomington: Indiana University.
  • Harnik, Roni, Graham Kribs, and Gilad Perez, 2006. “A Universe without Weak Interactions,” Physical Review D , 74(3), doi:10.1103/PhysRevD.74.035006
  • Harrison, Edward, 1985. Masks of the Universe , New York: Macmillan.
  • Holder, Rodney D., 2004. God, the Multiverse, and Everything: Modern Cosmology and the Argument from Design , Aldershot: Ashgate.
  • Hoyle, Frederick, 1982. “The Universe: Past and Present Reflections,” Annual Review of Astronomy and Astrophysics , 20: 1–35.
  • Hume, David, 1779 [1998]. Dialogues Concerning Natural Religion , Richard Popkin (ed.), Indianapolis: Hackett.
  • Janet, Paul, 1884. Final Causes , 2nd edition, Robert Flint (trans.), New York: Charles Scribner’s Sons.
  • Jantzen, Benjamin C., 2014a. An Introduction to Design Arguments , New York: Cambridge University Press.
  • –––, 2014b. “Piecewise versus Total Support: How to Deal with Background Information in Likelihood Arguments,” Philosophy of Science , 81(3): 313–31.
  • Kingsley, Charles, 1890. Water Babies , London: Macmillan.
  • Koperski, Jeffrey. 2005. “Should We Care about Fine-Tuning?” British Journal for the Philosophy of Science , 56(2): 303–19.
  • –––, 2015. The Physics of Theism: God, Physics, and the Philosophy of Science . Malden, MA: Wiley-Blackwell.
  • Kotzen, Matthew, 2012. “Selection Biases in Likelihood Arguments,” British Journal for the Philosophy of Science , 63(4): 825–39.
  • Kraay, Klaas. 2014. God and the Multiverse: Scientific, Philosophical, and Theological Perspectives . Routledge.
  • Lenoir, Timothy, 1982. Strategy of Life , Chicago: University of Chicago Press.
  • Lipton, Peter, 1991. Inference to the Best Explanation . 1st Edition. London: Routledge.
  • Loeb, Abraham, 2014. “The Habitable Epoch of the Early Universe,” International Journal of Astrobiology , 13(4): 337–339.
  • Manson, Neil (ed.), 2003. God and Design: The teleological argument and modern science , New York: Routledge.
  • –––, 2009. “The Fine-Tuning Argument,” Philosophy Compass , 4(1): 271–86.
  • –––, 2018. “How Not to Be Generous to Fine-Tuning Sceptics,” Religious Studies , first online 21 September 2018, doi:10.1017/S0034412518000586
  • McGrew, Timothy, Lydia McGrew, and Eric Vestrup, 2001. “Probabilities and the Fine-Tuning Argument: A Sceptical View,” Mind , 110(440): 1027–38.
  • McPherson, Thomas, 1965. The Philosophy of Religion , London: Van Nostrand.
  • Meyer, Stephen, 1998. “DNA by Design: An inference to the best explanation for the origin of biological information,” Rhetoric & Public Affairs , 1: 519–555.
  • –––, 2009. Signature in the Cell: DNA and the Evidence for Intelligent Design , New York: HarperOne.
  • Monton, Bradley. 2006. “God, Fine-Tuning, and the Problem of Old Evidence,” British Journal for the Philosophy of Science , 57: 405–24.
  • Murray, Michael (ed.), 1999. Reason for the Hope Within , Grand Rapids: Eerdmans.
  • Oberhummer, H.H., A. Csótó, and H. Schlattl. 2000. “Fine-Tuning of Carbon Based Life in the Universe by Triple-Alpha Process in Red Giants,” Science , 289: 88–90.
  • Paley, William, 1802. Natural Theology , Indianapolis: Bobbs-Merrill, 1963.
  • Peirce, Charles S., 1955. Philosophical Writings of Peirce , Justus Buchler (ed.), New York: Dover.
  • Pennock, Robert T., 2000. Tower of Babel: The Evidence against the New Creationism , Cambridge: MIT Press.
  • Penrose, Roger, 1990. The Emperor’s New Mind , Oxford: Oxford University Press.
  • Plantinga, Alvin, 2011. Where the Conflict Really Lies: Science, Religion, and Naturalism , Oxford: Oxford University Press.
  • Ratzsch, Del, 2001. Nature, Design and Science , Albany: SUNY Press.
  • Rott, Hans, 2010. “Idealizations, Intertheory Explanations and Conditionals,” in Belief Revision Meets Philosophy of Science (Logic, Epistemology, and the Unity of Science, Volume 21), edited by E. J. Olsson and S. Enqvist, 59–76, New York: Springer.
  • –––, 2003. “Perceiving Design,” in Manson 2003, pp. 124–144.
  • Smolin, Lee. 1999. The Life of the Cosmos , New York: Oxford University Press.
  • Sober, Elliott, 1993. Philosophy of Biology , Boulder: Westview.
  • –––, 2003. “The Design Argument” in Manson 2003, pp. 27–54.
  • –––, 2009. “Absence of Evidence and Evidence of Absence: Evidential Transitivity in Connection with Fossils, Fishing, Fine-Tuning, and Firing Squads,” Philosophical Studies , 143(1): 63–90.
  • –––, 2019. The Design Argument. Elements in the Philosophy of Religion , Cambridge: Cambridge University Press. doi: 10.1017/9781108558068
  • Whewell, William, 1834. Astronomy and General Physics Considered with Reference to Natural Theology , London: William Pickering.
How to cite this entry . Preview the PDF version of this entry at the Friends of the SEP Society . Look up topics and thinkers related to this entry at the Internet Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers , with links to its database.
  • The Fine Tuning Argument: A Curated Bibliography , by Neil Manson (The University of Mississippi).
  • Natural Theology; or, Evidences of the Existence and Attributes of the Deity , by William Paley.
  • Design Arguments for the Existence of God , by Kenneth Einar Himma (Seattle Pacific University), hosted by the Internet Encyclopedia of Philosophy.
  • The Teleological Argument and the Anthropic Principle , by William Lane Craig (Talbot School of Theology), hosted by Leadership U.
  • Was the Universe Made For Us? , maintained by Nick Bostrom (Oxford University).

abduction | Bayes’ Theorem | creationism | Darwinism | fine-tuning | Hume, David | Hume, David: on religion | Kant, Immanuel | teleology: teleological notions in biology

Acknowledgments

Del Ratzsch would like to thank his colleagues in the Calvin College Philosophy Department, especially Ruth Groenhout, Kelly Clark and Terrence Cuneo, and to David van Baak.

Jeffrey Koperski would like to thank Hans Halvorson, Rodney Holder, and Thomas Tracy for helpful comments on source material for section 4. Special thanks to Benjamin Jantzen and an anonymous referee for several comments and corrections on the 2019 version.

Copyright © 2023 by Del Ratzsch Jeffrey Koperski < koperski @ svsu . edu >

  • Accessibility

Support SEP

Mirror sites.

View this site from another server:

  • Info about mirror sites

The Stanford Encyclopedia of Philosophy is copyright © 2023 by The Metaphysics Research Lab , Department of Philosophy, Stanford University

Library of Congress Catalog Data: ISSN 1095-5054

University of Notre Dame

Notre Dame Philosophical Reviews

  • Home ›
  • Reviews ›

God and Design: the Teleological Argument and Modern Science

Placeholder book cover

Manson, Neil (ed.), God and Design: the Teleological Argument and Modern Science , Routledge, 2003, 400pp, $25.95 (pbk), ISBN 0415263441.

Reviewed by Niall Shanks, East Tennessee State University

This volume consists of nineteen essays devoted to the merits and failings of the argument from design from the standpoint of modern science. The contributors include distinguished scientists, philosophers and an assortment of creationists – a veritable mix of the good the bad and the ugly. For the present I note that the volume is divided into four parts. Part 1 is concerned with the characteristics of the argument from design, part 2 with physical cosmology, part 3 with the multiple universes hypothesis, and part 4 with biological versions of the argument from design.

The volume opens with an essay by Elliot Sober that provides a clear characterization and critical analysis of the various forms of the argument from design. The argument has biological and cosmological variants, and these need to be sharply differentiated. Sober raises critical points against both the biological and the cosmological (“fine tuning”) versions of the argument from design. This essay is followed by John Leslie’s reflections on the meaning of design. Where Sober’s essay was rooted in rigorous analysis of the structure of arguments trading in probabilities, Leslie’s essay is rooted in more traditional philosophical concerns about the nature of God-the-designer. Robert O’Connor’s essay considers claims by contemporary intelligent design theorists that their latest versions of the argument from design go beyond traditional versions from the standpoint of scientific respectability and philosophical presupposition. He draws a useful distinction between loca- design arguments and globa- design arguments . The latter try to infer design from the very presence of life in the universe, whereas the former try to infer intelligent agency and design from facts in the domain of contemporary science (for example, the biochemical pathways that have bewitched creationists such as Michael Behe). O’Connor’s conclusion is that notwithstanding claims to the contrary, the new intelligent design arguments are nothing more than old creationist wine in new designer-label bottles.

The first part of the anthology also contains essays by Jan Narveson (critical of the design argument), Del Ratzsch (who explores a perceptual, non-inferential approach to the evidences for design rooted in the philosophy of Thomas Reid), and Richard Swinburne (who supports the design argument on the basis of (uncalculated and unsupported) claims concerning the improbability of atheism). To give you a taste for Swinburne’s style of argumentation: he comments of the multiple-universes hypothesis favored by some cosmologists, “Rational inference requires postulating one simple entity to explain why there are many complex entities. But to postulate many complex entities to explain why there is one no less complex entity is crazy” (p. 117). While there is no reason (certainly none provided by Swinburne) to suppose that the postulation of a supernatural designer is either simple or even coherent, it is far from clear that Swinburne is right about the requirements of rational inference. After all, we believe complex objects such as tables are made up of vast multiplicities of atoms. These are objects with complicated internal structures and incredibly complex mutual interactions, as any quantum chemist or quantum field theorist knows. Not since the eighteenth century have scientists tried to explain the complex manifestations of matter – water, for example – by postulating a simple, elemental substance!

I turn now to part 2 of the anthology. This section of the book is concerned primarily with cosmological versions of the argument from design. In the comic novel Right Ho, Jeeves by P. G. Wodehouse, the immortal Bertie Wooster, himself no great genius, described an acquaintance, Miss Madeline Bassett as follows, “her conversation, to my mind, was of a nature calculated to excite the liveliest suspicions. Well, I mean to say, when a girl suddenly asks you out of a blue sky if you don’t sometimes feel that the stars are God’s daisy chain, you begin to think a bit.” The disputes about the cosmological versions of the argument from design are essentially disputes between those who think that the stars (or the fundamental cosmological constants …) are God’s daisy chain, and those who are frankly skeptical.

Physicist Paul Davies opens this part of the anthology with a discussion of the appearance of design in physics and cosmology. The essay is largely speculative in nature, and Davies contends (though he does not argue) that the emergence of life and consciousness were somehow written into the basic laws of nature (he knows not how). Reading this essay one cannot help but feel that Davies should stick to his field of expertise – quantum gravity – and leave metaphysics to those who at least have some arguments up their sleeves.

This essay is followed by William Lane Craig’s ruminations on the anthropic fine-tuning of the universe. Put bluntly, many numbers appear in the standard cosmological model that are not predicted by basic theory but have to be fixed empirically. If these numbers were slightly different, then we would not be here for the universe would be inhospitable to our kind of life. Craig (and numerous others) see this fine-tuning of nature’s constants as evidence of providential design. On the face of it, this is a colossal argument from ignorance: we currently do not have a good physical theory that explains why nature’s constants take the values they do, so they must have been fixed by God.

In responding to the claim that the postulation of a providential designer only pushes the problem back one step (what then explains the designer), Craig argues, “It is widely recognized that in order for an explanation to be the best, one need not have an explanation of the explanation… If the best explanation of a disease is a previously unknown virus, doctors need not be able to explain the virus in order to know that it caused the disease” (p. 175). As Craig must surely be aware, the hypothetical virus only becomes the best explanation after independent evidential warrant for its postulation is forthcoming from empirical medical and virological inquiries (“It enters the body this way…”, “It parasitizes intracellular mechanisms this way…”, “It propagates this way….,” and so on). Until these inquiries are conducted, the viral hypothesis for the cause of a disease is an unsubstantiated hypothesis that explains nothing. Postulating a hitherto unknown supernatural being simply to explain the fine tuning of the universe, with no further evidence of mechanism and details, is mere idle speculation.

Craig’s essay is followed by essays by Robin Collins (supportive of claims about fine-tuning in the face of skeptics such as Steven Weinberg); and by Timothy McGrew, Lydia McGrew, and Eric Vestrup (skeptical about the role played by probabilities in fine-tuning arguments).

Part 3 of the anthology is concerned with the “multiple universes” hypothesis as a non-supernatural device to explain the anthropic coincidences. This section opens with an essay by astrophysicist Sir Martin Rees. Rejecting explanations of fine-tuning in terms of either chance or providential design (and, for that matter, ignoring that the apparent need for such explanations arises from the current incomplete state of cosmological science), Rees suggests that the fine-tuning could be explained by the postulation of many universes (in which the laws and physical constants differ from those obtaining in our universe). Our universe, far from being either designed or being a freak of nature, would then belong to a subset of the set of universes that offered habitats conducive to the emergence of complex life and consciousness. The basic idea is as follows: the probability of rolling a fair die five times and getting five 6s in a row is very small (0.00013). But if you have many millions of people rolling dice, there will be thousands, who, by chance alone, will roll five 6s in a row.

Personally I find the postulation of multiple, unobservable universes to be less than compelling. The probability that you get five 6s in a row in a given trial involving five rolls of the die may be 0.00013. But this same probability attaches to any other sequence outcomes from five rolls of the die (e.g., two 5s, a 4, and two 2s). This probability, moreover, tells us nothing about when, in a sequence of trials, each involving five rolls of the die, you will get five sixes. Perhaps it happens in the first run of five rolls of the die, perhaps in the twenty -seventh. That you anthropically care about five 6s in a row – perhaps you get a prize – hardly merits the postulation of either natural or supernatural entities in a realm invisible.

D. H. Mellor’s essay also expresses skepticism about the “multiple universes” hypothesis. Mellor questions whether it offers an explanation of our existence. In ostensibly dealing with the question “Why do we exist?” in terms of multiple universes, Rees seems to answer another, very different question (i.e., “Why do we exist where we exist?”). As Mellor observes, “to change the question in this way is like turning the question of why there are fish (say) into the question of why they live where they do, namely in water: to which the obvious answer is that water, unlike dry land, has what fish need” (p. 223).

Roger White’s essay continues the exploration of fine-tuning and the explanatory value of multiple universes. I find myself uneasy with White’s conclusion that “assuming there is just one universe, the fact that it is life-permitting is surprising. For this otherwise extremely improbable outcome of the Big Bang is more probable on the assumption that there is a cosmic designer who might adjust the physical parameters to allow for the evolution of life” (p. 243). Physicists know virtually nothing about the details of the instant of the Big Bang (as opposed to what happened a miniscule fraction of a second later). It is hard to say what is probable and what is not. Nothing follows about nature from our ignorance of it. That you can tell a theological science-fiction story about cosmic designers tweaking cosmological parameters hardly renders the existence of a single, life-permitting universe more probable than otherwise. It would be different if we had some independent evidence that there was such a being with the requisite wherewithal, but we do not.

White’s essay is followed by a piece by creation scientist William Dembski. Dembski’s work has been widely criticized elsewhere, and I will not rehearse those criticisms here.

While I share Dembski’s unease with the postulation of multiple universes, I find his counter-proposal to be deeply problematic: “We already have experience of human and animal intelligences generating specified complexity… Thus, when we find evidence of specified complexity in nature for which no embodied, reified or evolved intelligence could plausibly have been involved, it is a straightforward extrapolation to conclude that some unembodied intelligence must have been involved. Granted, this raises the question of how such an intelligence could coherently interact with the physical world. But to deny this extrapolation merely because of a prior commitment to naturalism is not defensible” (pp. 266-267). The problem here lies not with a prior commitment to naturalism, but with Dembski’s failure to adequately address the question of coherent interaction between the physical realm and the nonphysical realm, for unless this matter is properly dealt with, the extrapolation he envisions is in fact highly problematic. That evolved physical beings such as ourselves can design complex artifacts is evidentially irrelevant to the issue as to whether there are nonphysical supernatural beings who can design universes – objects that are, on the face of it, very different from the artifacts of human experience.

Finally we arrive at Part 4 of the anthology, which involves essays devoted to the issue of intelligent design in biology. Michael Behe opens this section with a discussion of the modern intelligent-design hypothesis. Behe, like Dembski, has been widely criticized in other forums, and Behe’s essay is followed here by an excellent critical essay by Kenneth Miller which does much to dismantle his uncritical speculations about biological intelligent design.

Behe’s strategy is to point to known features of biochemistry and to argue that since there is currently no obvious way (given our present state of knowledge) in which these systems could have evolved (we are ignorant of how they came to be – though not as ignorant as Behe would have you believe), then they must be fruits of intelligent design. This is essentially the same pattern of inference underlying the leap from our ignorance of the details of fine-tuning to the conclusion of supernatural design.

Michael Ruse’s essay on modern biology and the argument from design is perhaps one of the best essays in the entire volume from the standpoint of scientific insight and philosophical depth. Readers seeking a subtle account of the interface between science and religion need look no further than Ruse’s essay in this volume. Following Ruse, there is a speculative essay by Simon Conway Morris that purports to see evidence of a cosmic teleology in the facts of evolutionary convergence.

The final essay in the volume is by Peter van Inwagen, who explores the issue of the compatibility of Darwinism with the argument from design. Like the essay by Michael Ruse, this essay exhibits a considerable degree of philosophical and theological sophistication, and like Ruse’s essay, it too will repay serious study. For van Inwagen, if unaided natural selection could produce the ordered diversity we see in the biological world, why could not a God (or some other intelligent being) desiring such diversity have used this same elegant mechanism? “Anyone who thinks that the history of terrestrial life is inconsistent with its being the vehicle by which God’s purposes have unfolded in time really should have something to say about how the history of life would look if it were the vehicle of God’s unfolding purposes” (p. 362). Perhaps, after all, one can have one’s cake and eat it. Maybe so, but we are a long way here from the hostility to modern science implicit in the work of creationists, intelligent design theorists and Biblical literalists. If van Inwagen is right, it may be possible after all to rise above the din generated by the clash of naïve scientism with naïve theology.

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

The Oxford Handbook of Natural Theology

  • < Previous chapter
  • Next chapter >

The Oxford Handbook of Natural Theology

18 The Design Argument and Natural Theology

Neil A. Manson, Associate Professor of Philosophy at the University of Mississippi.

  • Published: 03 June 2013
  • Cite Icon Cite
  • Permissions Icon Permissions

In the broadest sense, natural theology is the effort to gain knowledge of God from non-revealed sources – that is, from sources other than scripture and religious experience – but there is also a much narrower sense of natural theology: the construction of arguments for the existence of God from empirical evidence. This narrower sense is most strongly associated with the argument for God's existence from the appearance that the natural world has been constructed for a purpose. This argument is referred to as ‘the Design Argument’. This chapter addresses a generic theological question confounding the Design Argument. Why would God design or create anything at all, much less a world like this one? Let us call this question ‘Why design?’ (WD). It is shown that answering WD entangles proponents of the Design Argument in age-old debates about divine freedom, divine moral perfection, and divine rationality. Despite the pretensions of some of its proponents, the Design Argument is not and never has been ‘strictly scientific’.

Introduction

In the broadest sense, natural theology is the effort to gain knowledge of God from non-revealed sources—that is, from sources other than scripture and religious experience. Natural theology in this sense encompasses an array of activities ranging from attempts to understand the divine attributes (omnipotence, omniscience, moral perfection, and so on) to establishing the existence of the soul. The essays in this volume attest to the vibrancy of natural theology understood in this sense. But there is also a much narrower sense of natural theology: the construction of arguments for the existence of God from empirical evidence. Although this narrower sense technically includes the argument for God's existence from the very existence of a contingent reality—‘the Cosmological Argument’—it is most strongly associated with the argument for God's existence from something much more specific: the appearance that the natural world, in whole or in part, has been constructed for a purpose. Philosophers call this ‘the Teleological Argument’, from the Greek word ‘ telos ’, meaning ‘goal’ or ‘end’. It is more commonly called ‘the Design Argument’.

The Design Argument touches on an enormous variety of issues in science, logic, metaphysics, epistemology, and value theory. Furthermore, there is not one version of it, but many. These versions can be distinguished according to two factors: the sorts of scientific evidence to which they appeal and the logical nature of the sorts of inferences being made. Perhaps a dozen distinct fields of inquiry have at one time or another provided fodder for some version of the Design Argument. Additionally, a host of different argument forms have been employed (Manson 2003 : 5–8; Oppy 2006 : 202–16). Yet three versions stand out in particular.

The most influential version historically is that of William Paley ( 2006 [1802]), who appealed to functional biological parts (eyes, wings, and so on) whose structures seemed explicable only in terms of a supernatural designer. The version currently generating the most heat is ‘Intelligent Design Theory’. Like Paley's argument, this version appeals to functional biological parts, but this time at the cellular level (Behe 1996 ). The claim is that some of these parts exhibit a feature—‘irreducible complexity’—that cannot be explained by some combination of law and chance (i.e. by Darwinian natural selection), leaving design as the only remaining explanation. Finally there is ‘the Fine-tuning Argument’ (Manson 2009a ). According to it, the very laws of physics and constants of nature themselves involve numerical values such that, had they been the slightest bit different, life of any sort would not have been possible in the universe (Barrow and Tipler 1986 ; Leslie 1989 ; Collins 2003 ; Holder 2004 ). If there is no God to design the universe, we must believe something incredibly improbable happened, whereas if God does exist, we have an explanation of fine-tuning.

Instead of addressing the philosophical and scientific particularities of each version, this chapter addresses a generic theological question confounding the Design Argument. Why would God design or create anything at all, much less a world like this one? Let us call this question ‘Why design?’ (WD). Oddly, neither opponents nor proponents of the Design Argument have given WD much attention (Manson 2009b ). Yet answering WD is obviously crucial if the Design Argument is to succeed. If there is no reason to think God would design or create anything at all, then a fortiori there is no reason to think God would fine-tune the universe for life, make irreducibly complex cellular structures, fashion animals with eyes and wings, or bring about any of the other features of the world that supposedly have such a slim chance of being exhibited in a reality lacking God. As we will see, answering WD entangles proponents of the Design Argument in age-old debates about divine freedom, divine moral perfection, and divine rationality. Despite the pretensions of some of its proponents, the Design Argument is not and never has been ‘strictly scientific’. Natural theology in the narrow sense demands natural theology in the broad sense.

Framing the Issue: Possible Creations vs. Possible Worlds

At the outset, a clarification of WD is in order. Typically it is framed as the question of what God would ‘actualize’ and why. The picture here is of God contemplating all the possibilities, then picking one for realization (‘actualization’ in the jargon of analytic philosophy) according to some rationally defensible selection principle. On this picture possibilities are thought of as abstract entities on the ‘actualist’ model of Alvin Plantinga ( 1974 ) rather than as concrete entities on the ‘modal realist’ model of David Lewis ( 1986 ). The point of clarification concerns whether actualization-by-God is to be conceived as the actualization of a possible creation or the actualization of a possible world .

A possible creation is simply a proper part of a possible world—the part consisting of all and only the objects such that, were they actual, they would be contingent and non-abstract. On the actualist model of possible worlds, abstract objects—objects such as numbers, propositions, properties and other items that seem to be outside of space and time—are necessary and indestructible. As such, they are not subject to being created. Furthermore, though God is not an abstract object, His existence is logically prior to His actualizing anything and He exists independently of any other thing. Thus God, too, is not properly conceived as a part of a possible creation. A possible world, on the other hand, comprises abstract objects as well as ‘concrete’ (that is, non-abstract) objects, necessarily existing or independently existing objects as well as contingent objects. Possible worlds are maximal states of affairs, including necessary states of affairs as well as contingent ones (Plantinga 1974 : 44–5). Thus possible worlds are the sorts of things that might include God, whereas possible creations are not (Plantinga 1974 : 196–221).

Since a possible creation is merely a part of a possible world, it seems there could be a possible world without a created part. Two such possible worlds suggest themselves immediately. First is a possible world containing abstract objects and no other things. For reasons beyond the scope of this chapter, metaphysicians have typically held there can be only one such world. Metaphysicians who ask why there is something rather than nothing are asking why it is not the case that this world is actual. Second is the possible world containing abstract objects, God, and nothing else. Call this world ‘God Alone’ (GA). Thus the philosopher or theologian who asks why God would create anything at all is asking why it is not the case that GA is actual.

If actualization-by-God is conceived as the actualization of a possible creation , then, in asking what world God would actualize, the question does not arise of why GA is not actual. On this way of looking at actualization-by-God, God must actualize some possible creation, with the only issue being which one He picks. Now this way of conceiving actualization-by-God raises significant theological questions. One is why God would actualize a possible creation with any evil in it. This is the famous ‘Problem of Evil’. Another is how God could be both free and morally unsurpassable if, no matter what possible creation He actualizes, there is always another possible creation that is even better (Rowe 2004 ). Nonetheless, this way of conceiving actualization-by-God cloaks a question, WD, that comes to the fore if actualization-by-God is conceived as the actualization of a possible world , since on that conception, God's actualizing GA by not actualizing any possible creation at all is explicitly allowed as possible. This is a strong reason for conceiving of actualization-by-God as the actualization of a possible world rather than of a possible creation: doing so highlights a question of obvious interest. Another reason is simply that talk of actualizing creations is not standard within analytic philosophy whereas talk of actualizing possible worlds is (Swinburne 2003 : 121, n. 8).

WD thus breaks into two parts. First, why would God actualize some possible world other than GA? In other words, why would God not content Himself with a reality the concrete part of which was just Himself? Secondly, if God would actualize some possible world other than GA, what kind of world would that be? Adopting the terminology of Norman Kretzmann, we can refer to these two questions respectively as ‘the general problem of creation’ (Kretzmann 1991a ) and ‘the particular problem of creation’ (Kretzmann 1991b ). The focus of the remainder of this chapter is the general problem of creation. Before exploring the general problem of creation, however, we must be aware of some constraints imposed by the concept of the most perfect being.

Constraints From Traditional Theology

‘Perfect being theology’ (PBT) is the result of constructing a theology on the assumption that God is absolutely perfect—absolutely lacking in any deficiency. How, if at all, does PBT constrain any answer to WD?

Two constraints seem obvious. First, it cannot be that God actualizes some world other than GA out of need; otherwise God's absolute self-sufficiency would be threatened. This point is accepted by theologians as far back as Augustine (Teske 1988 : 248–9). Spinoza, controversially, generalized this latter point to argue that it is incoherent to think an absolutely perfect being would have desires, goals, or purposes, since he analysed desires, goals and purposes in terms of the fulfillment of needs. He claimed that ‘if God acts with an end in view, he must necessarily be seeking something that he lacks’ (Spinoza 1982 [1677]: 59). From Spinoza's perspective, any allegedly perfect being that acts to fulfill a desire, goal, or purpose would be incomplete and hence imperfect.

Secondly, it cannot be that God actualizes some world other than GA contrary to reason or for no reason at all, because that would contradict God's perfect rationality. It might be objected here that what is or is not reasonable is itself determined by the will of God. More modestly, it might be objected that the rationalist, Leibnizian conception of the divine will should not be built into PBT; it should not be presumed that there is a sufficient reason for everything God does. Instead, goes the objection, PBT should be open to a conception of the divine will according to which God could act for no reason whatsoever, so long as God does not act contrary to reason (Mawson 2005 : 149). Without dismissing this proposal, we may just note here that it marks a significant departure from traditional PBT. The goal of this chapter is to show that it is very difficult for the traditional theist to answer WD. Non-traditional theists (e.g. process theologians) may have resources to answer WD that traditional theists do not, but non-traditional theists are not our concern. For the remainder of the chapter we will thus presume that PBT forbids answers to WD whereby God creates for no reason.

A third constraint is more controversial than the first two: that PBT forbids answers to WD whereby God creates of necessity. The thought here is that any such answer would conflict with the idea that God is perfectly free. However, whether divine freedom should be conceived in the libertarian sense rather than the compatibilist sense is a matter of dispute. The libertarian says that if S did X freely, then it was possible that S did otherwise than X. The compatibilist says possibly S did X freely even though S could not have done otherwise than X if S's doing X was what S wanted, desired, or decided. Great minds can be found on either side of the libertarianism/compatibilism issue, with some (e.g. Augustine and Aquinas) vacillating (Kretzmann 1991a : 209–23; Teske 1988 ; Kretzmann 1999 : 101–41; Mann 1991 ). The vexing question for the compatibilist about divine freedom is why compatibilism should not likewise apply to human freedom (O’Connor 2008 : 121). That question is vexing because there being no answer to it threatens one very popular solution to the problem of evil—namely, that the possibility of evil is a precondition for the possession by humans of genuine freedom. At this point we can simply note that there are differing opinions on this issue—some think PBT requires a libertarian conception of divine freedom, others think PBT allows that God's actions are both necessitated and free—and that this difference will be relevant when it comes to assessing answers to WD.

Theologian Karl Barth acknowledges precisely these three constraints from PBT at the start of a passage in which he claims to answer WD:

Creation is the freely willed and executed positing of a reality distinct from God. The question thus arises: What was and is the will of God in doing this? We may reply that He does not will to be alone in His glory; that He desires something else beside Him. But this answer cannot mean that God either willed and did it for no purpose, or that He did so to satisfy a need. Nor does it mean that He did not will to be and remain alone because He could not do so. [Barth 1961 : 149–50]

These concessions seem to leave Barth with little room to manoeuvre. How then, does he answer WD?

The Occurrentist Fallacy

For Barth, the answer lies in the notion of divine love:

If, then, this positing is not an accident, if it corresponds to no divine necessity and does not in any sense signify a limitation of His own glory, there remains only the recollection that God is the One who is free in His love. In this case we can understand the positing of this reality—which otherwise is incomprehensible—only as the work of His love. He wills and posits the creature neither out of caprice nor necessity, but because He has loved it from eternity, because He wills to demonstrate His love for it, and because He wills, not to limit His glory by its existence and being, but to reveal and manifest it in His own co-existence with it. [Barth 1961 : 150]

So, for Barth, God created the world because He loved it. But this answer is befuddling. Doesn’t the world have to exist first for it to be loved? If so, how can love explain why the world exists in the first place?

Included within PBT is the idea that God is perfectly loving. Being perfectly loving, God will, of course, love any being that is an appropriate object of love (a human, say, rather than an electron). But that in no way explains why there exist objects of God's love. Barth's statement that God creates because ‘He has loved it [creation] from eternity [and] wills to demonstrate His love for it’ cannot mean that the creation exists independently of God. Otherwise its existence would run counter to the traditional doctrine of creation ex nihilo , another core element of PBT. So Barth still has not given us an answer to WD. Yes, God is perfectly loving, but why should that lead us to expect God to create objects of His love?

The property of being perfectly loving was just presented as a dispositional property—a property that need not be manifested to be possessed. A dispositional analysis of being perfectly loving might go something like this: S is perfectly loving if, and only if, if X is an appropriate object of love, then S loves X unconditionally. Perhaps Barth thinks being perfectly loving requires more than just the disposition to love unconditionally those things that are appropriate objects of love. Perhaps he thinks that, for a subject S to be perfectly loving, there must be an actual occurrence of unconditional love—there must exist at least one object (other than S) of S's unconditional love. Philosopher Leon Pearl maintains precisely this. He reminds us that Aristotle claimed ‘a poor person cannot be charitable. True, if the person was rich, he would be charitable; but not being rich he is in no position to exercise acts of charity; and is, therefore, not charitable’ (Pearl 1994 : 333).

Yet this position seems mistaken. The case Pearl gives of charity is not analogous to the case of love. To be analogous, the issue would have to be, not whether a person could be charitable while having no riches to give, but whether a person could be charitable while having no one to whom riches could be given. If being charitable requires that there be someone to whom riches could be given, however, then being charitable is no longer an intrinsic property of a person. On that view, a person would cease to be charitable if everyone around her suddenly became wealthy or suddenly died. Intuitively, however, properties such as being charitable, being just, being loving, and so on are intrinsic properties—if not in the human case, then certainly in the divine case (according to PBT). That means they must be dispositional properties rather than properties that must be manifested to be possessed. (Call these latter ‘occurrent properties’ to distinguish them from dispositional ones.) If we think the divine virtues are dispositional rather than occurrent properties, then Pearl's argument falls flat—as does Barth's, for the same reason.

A second criticism of Barth's suggestion is that, since the act of creation necessarily precedes any act of justice or of love (because such acts require that the appropriate objects exist), the act of creating such objects does not qualify as an act of justice or of love. Tim Mawson points this out when he says of created things that ‘their not existing prior to His creating them means that they themselves could hardly be said to have previous requirements met by their being created’ (Mawson 2005 : 149). Aquinas made the same point (Kretzmann 1999 : 130–1).

A third problem with Barth's suggestion is that occurrentist accounts of the divine virtues impinge on divine freedom by making God create of necessity, by making God create as a result of God's very nature as loving, benevolent, and just. Aquinas rejected the occurrentist account of justice for precisely this reason. As was noted earlier, there is a disagreement as to whether PBT forbids saying God necessarily creates. Indeed, according to Kretzmann ( 1997 : 223), Aquinas himself was of two minds on this issue. While Aquinas ‘explicitly endorses’ the libertarian explanation of creation, ‘his conceptions of God, goodness, creation, and choice entail a necessitarian explanation to which he was clearly drawn and which gets expressed, perhaps inadvertently, even in the context of a thoroughgoing presentation of his official libertarian line’. Given this controversy, I will merely note here that those with a libertarian conception of divine freedom have additional reason to reject the occurrentist account of the divine virtues, and hence to reject any answer to WD of the sort Barth gives.

It is worth noting that the sort of answer Barth gave has considerable precedent. Both Plato and Augustine suggested that the motive for creation was the lack in God of the vice of envy. Kretzmann calls attention to this passage from Plato's Timaeus :

Let us now give the reason why the maker made becoming, and the universe. He was good , and in him that is good no envy ever arises regarding anything. Being devoid of envy, he wanted everything to be like himself , as far as possible … God desired that everything should be good and nothing evil, as far as possible … For him who is most good it neither was nor is permissible to do anything other than what is most beautiful. [Plato 1902 : 29E–30B; cited in Kretzmann 1991a : 213–14]

Roland Teske cites several passages in which Augustine makes a similar suggestion. The following one is representative:

But if he could not make good things, there would be no power; if, however, he could and did not, there would be great enviousness ( invidentia ). Hence, because he is almighty and good, he made all things very good. [Augustine: 4.16.27 (PL 34.307; CSEL 28.1, p. 113, lines 10–13), cited in Teske 1988 : 249]

As with being loving, however, it seems the property of being envious is a dispositional one. A crude dispositional analysis of envy might go something like this: S1 is envious if, and only if, if there is another subject S2 such that (i) S2 possesses something S1 regards as good and (ii) S1 knows that (i), then S1 bears ill will towards S2. This analysis suggests that lacking envy is also a disposition. Let us help Plato and Augustine along here and suppose that they intend to speak, not of the absence of a vice, but the presence of a virtue: good will. Hence S1 is perfectly good-willed if, and only if, for any other subject S2, if (i) S2 possesses something S1 regards as good and (ii) S1 knows that (i), then S1 bears good will towards S2.

Given this (admittedly crude) analysis of what it is to be perfectly good-willed, can being perfectly good-willed explain why God created? No more so than being perfectly loving can. In all cases, the possession of the dispositional property is consistent with the non-existence of anything that might make the disposition manifest. We get no explanation of why there must exist objects of the disposition by being told God possesses the disposition. Since in this section we have seen that the mistake of thinking otherwise is made often enough, let us create a label for the mistake. Call it ‘the occurrentist fallacy’.

The Dionysian Answer: The Good As Essentially Diffusive

If all of the divine attributes are dispositional properties rather than occurrent ones, then it looks as though we simply cannot answer WD by reference to the divine attributes—at least, not without committing the occurrentist fallacy. According to Kretzmann, however, Aquinas held that at least one of the divine attributes should not be understood as a mere disposition. That attribute is goodness (Kretzmann 1991a : 215–23; 1997: 220–5; 1999: 120–36). According to Kretzmann ( 1999 : 134), Aquinas endorsed the Dionysian Principle that ‘goodness is by its very nature diffusive of itself and (thereby) of being’. Kretzmann ( 1991a : 217) himself agrees with this principle, saying it ‘expresses an important truth about goodness, most obviously about the goodness of agents’. Kretzmann admits that there are coherent dispositional accounts for some of the divine attributes (e.g. omniscience), but denies that goodness can be merely dispositional. ‘There is no obvious inconsistency in the notion of knowledge that is unexpressed, never shared by the agent who possesses it even if he is omnipotent’, he says (1991a: 217), ‘but there is inconsistency in the notion of goodness that is unmanifested, never shared, even though united with omnipotence’.

The Dionysian Principle is a beautiful thought, but we get no argument from Aquinas or Kretzmann that it is true. Perhaps it is simply a fundamental aesthetic/moral intuition for which no argument can be given. Appealing to the Dionysian Principle, however, seems to violate the third constraint imposed by PBT—the constraint that any answer to WD cannot entail that God creates necessarily. If God is essentially good, and the good is essentially creative, then God is essentially creative. It is part of God's very nature to be creative, so that God could not refrain from creating a contingent reality and still be God. Kretzmann documents how Aquinas struggled with this implication of his solution. According to Kretzmann ( 1991a : 215), there are both ‘libertarian and necessitarian strains in Aquinas’ that Aquinas never reconciled. ‘I see no way of avoiding the inconsistency (or, at least, ambivalence) in Aquinas's account as it stands’, he adds, tracing the ambivalence back to a fundamental conflict in Greek thought about God between ‘Platonist self-diffusiveness and Aristotelian self-sufficiency’ (Kretzmann 1991a : 222).

Kretzmann does think there is a way to tie together the competing strains, to reconcile the libertarian and necessitarian conceptions of creation:

what I think Aquinas should say about God's need to create something [is] that goodness does require things other than itself as a manifestation of itself, that God therefore necessarily though willingly wills the being of something other than himself, and that the free choice involved in creation is confined to the selection of which possibilities to actualize for the purpose of manifestation. [Kretzmann 1991a : 223]

Kretzmann is not the only one who finds this compromise plausible. Timothy O’Connor judges:

there is a strong case to be made that a perfect being would create something or other, though it is open to Him to create any of a number of contingent orders … Perhaps it is inevitable, then, that a perfect God would create. [O’Connor 2008 : 112–13]

In response, let me spell out an unwelcome consequence of the Dionysian answer from the perspective of the libertarian about divine freedom. According to the Dionysian Principle, it is metaphysically impossible that God sits on His hands. Doing nothing is not an option for Him. He is not free to actualize GA, according to Kretzmann, despite the seeming possibility that He does so. This is in tension with the idea that we should be grateful to God for not having actualized GA—that is, for having created something other than Himself. Creation can no longer be seen as a gift or an act of grace if the act of creating is necessitated (Rowe 2004 : 2).

The restriction on God seems to run further. Suppose, as is highly plausible, that there is no upper bound to the amount of goodness that could be created (this is the presupposition of Rowe's objection to the idea that God is both free and yet morally unsurpassable). Then no creation of finite value is compatible with the Dionysian Principle unless we arbitrarily limit the scope of that principle. Say, for example, that God diffuses His goodness to produce a creation possessing N units of value, even though it is possible for there to be a creation of N + M units of value. Suppose further that the explanation of God's creating just N units of value is that, although His goodness necessarily diffuses itself, it only does so to a limited extent—just enough to produce N units of value but not enough to produce N + M units of value. This would be an arbitrary, brute fact about divine goodness—a very puzzling fact. What force or factor would be confining the divine goodness to just N units of value? Nothing outside of God could restrain His goodness, and it seems nothing internal to God would do so. So if the Dionysian Principle is correct, not only is God not free to actualize GA, He is not free to actualize any possible world with a created component of finite value, on pain of acting arbitrarily. Some recent philosophers enamoured of the Dionysian Principle (Leslie 2001 ; O’Connor 2008 ) have embraced this consequence, proposing that God creates an ‘infinitely membered super-universe, whose members are ordered by value without upper bound’ (O’Connor 2008 : 118–19). Yet this consequence is unorthodox, to say the least. In short, serious limitations on God's choice of what world to actualize seem to follow from the Dionysian Principle.

Perhaps some easily accept these consequences, declaring that we have reconciled ourselves to the counterintuitive consequences of other ‘limitations of perfection’. I suspect the case of Aquinas is more representative of the actual situation. His appeal to the Dionysian Principle results in an account of creation marked by ‘inconsistency or ambivalence’ (Kretzmann 1991a : 223). This suggests adherents of PBT seeking an answer to WD should at least be cautious in their appeals to the Dionysian Principle. Not even Aquinas could figure out how to reconcile it with PBT.

Might GA Be Less than the Best of All Possible Worlds?

Earlier, three constraints PBT imposes on any answer to WD were proposed. Is there a fourth constraint—namely, that PBT forbids answers to WD whereby GA is something less than ‘the best of all possible worlds’ (BPW)? The idea that GA is BPW seems to follow directly from the idea that God is the best of all possible beings. Kretzmann states that ‘the existence of an absolutely perfect being and nothing else at all seems unquestionably the best of all possible worlds’ (Kretzmann 1997 : 221)—a claim which leads him to ask ‘what could motivate God to choose to create anything at all?’

Whether PBT imposes a fourth constraint on any answer to WD will depend on our metric for value and on whether we equate maximal possession of a value with possession of that value to an infinite degree. If there is only one fundamental kind of value—and so only one dimension along which to measure value—and if PBT entails that God possesses that value to an infinite degree, then it seems to follow directly that, of all the possible worlds, GA rates the highest. Given that God has infinite value, and given that there is only one kind of value, God will have that value to an infinite degree. Say, for example, that the only kind of value is happiness. In that case, PBT entails that God has happiness to an infinite degree, so that no possible individual or sum of individuals—including sums with God as a part—surpasses God in happiness. Timothy Chappell ( 1993 ) uses precisely these assumptions to prove that God is not a consequentialist—a maximizer of value—for if He were, then He would actualize GA, which He obviously did not. Robert Nozick made a similar point (1989: 225). Of course, there may be more than one ultimate kind of value. Indeed, some forms of ultimate value might even conflict with one another. But if there is only one ultimate value, then GA must be rated BPW on the metric for that value. As Kretzmann just pointed out, this seems to leave us with no way to answer WD.

As William Mann interprets medieval theologian Vital du Four, a startling consequence emerges from the idea that there is only one kind of value and God possesses it to an infinite degree: all possible worlds that God might actualize are of equal value.

‘Every creature has an accidental relation to the goodness of God, because nothing is added to his goodness from them, just as a point adds nothing to a line.’ … One can interpret du Four's remark in such a way that it allows that some creatures have intrinsic value, even that some creatures are intrinsically more valuable than others, but simply denies that the existence of any of them can add to God's goodness. This interpretation would stress that God's goodness is perfect and infinite, thus that the addition of any amount of created goodness cannot increase his goodness. In fact, one can maintain further that the addition of any amount of created goodness cannot increase the aggregate amount of goodness in existence, since God's goodness alone provides an infinite amount of goodness in existence. [Mann 1991 : 253–4, citing du Four 1891 : 307]

Here du Four seems to have in mind the paradoxical question beloved of beginning maths students: ‘What is infinity plus one?’ For du Four, the answer is ‘infinity’. If an infinite number of marbles is in existence, adding one does not increase the number of marbles, nor does taking one away (nor, for that matter, does adding another infinity of marbles, assuming that both infinities are of the same cardinality). Here du Four conceives of God's creating as adding a finite quantity to an infinite one. Thus the ‘goodness rating’ of any possible world with God in it is going to be the same: infinity. So on du Four's view, GA is tied for the title of BPW with every other possible world God might actualize—simply by virtue of the fact that all of those worlds include God. No world surpasses GA in goodness, but countless worlds equal GA in goodness.

If we adopt the picture of perfect goodness of du Four, the presumed constraint that God must actualize BPW turns out to be no constraint at all. If every possible world with God in it is equal in its value, then we cannot answer WD in terms of God's wanting to actualize as good a possible world as He can, because God's wanting to do so in no way favours actualization of some world other than GA. The world containing God, abstract objects, and cosmic dust is just as valuable as our world. So is the world containing God, abstract objects and five times as much cosmic dust. So is the world containing God, abstract objects, and one very grumpy cat. And so on. None of these worlds stand above GA in value, and so there is no reason to actualize one of them over GA. Thus to actualize one of them over GA would be for God to act without reason; according to the second constraint of PBT on any answer to WD, God cannot so act.

There are two points to jump off here. First, one might maintain that, while God possesses the maximum value any individual can have, it does not follow that God possesses infinite value. The idea here is that there is an intrinsic maximum to the value that can be possessed by any individual. But this idea is hard to square with the assumption that there is only one fundamental kind of value. If there was more than one kind of value, then we could imagine a function such that maximizing the combined value of the several values is inconsistent with maximizing any particular one of the several values. But how can there be an upper limit on the value of an individual being if there is only one kind of fundamental value? What could be imposing the upper limit? There seems to be no way to answer this question. If there is only one kind of value, it seems God having maximal value entails God having infinite value.

This suggests a second option. One might reject the idea that there is only one kind of fundamental value. Nothing about PBT commits us to this idea. The alternative idea, proposed by Mann ( 1991 : 268–76) among others, is that there are plural, incommensurable values—fundamental values with no deeper measure of value against which they might be compared. What might these values be? O’Connor ( 2008 : 117–18) proposes three different measures of the value of a possible creation: the intensive value of the parts of a creation (so that a Rembrandt painting rates higher than a tube of toothpaste); the extensive or aggregate value of the whole creation (so that a world with a thousand happy people rates higher than one with just five); and the organic value of the creation (the value attaching to the relational structure of the world).

The idea that intensive and extensive/aggregate value are fundamental yet incommensurable values should be familiar to observers of the debates amongst hedonistic utilitarians. Which is better, the hedonistic utilitarians ask themselves: a world with a small number of extremely happy people or an enormous number of marginally happy people? Which is more important: average utility or total utility? Hedonistic utilitarians have tied themselves in knots over this question, which suggests both to ideal utilitarians and to other outside observers that average utility and total utility are incommensurable values. In a similar vein, O’Connor is suggesting that intensive value, extensive value, and organic value are incommensurable. They are all fundamental values, but since they are fundamental there is no deeper measure of value against which we can compare them.

Schemes of incommensurable values other than O’Connor's can no doubt be proposed. The issue here is whether or not it is possible that these values are capable of all being maximized within a single individual. If one holds that they not only can be, but that they are, in fact, all maximized in God, then appealing to incommensurable values does nothing to solve the problem at hand. If God maximizes every incommensurable value, then GA rates the highest on every measure of value. For every fundamental measure of value there is a BPW with respect to that value, but in each case the BPW is the same: GA. But if GA is BPW on all measures of value, we still have no reason to expect God to actualize any world but GA. Surprisingly, O’Connor himself suggests that, while it may be that no possible creation has maximal value along all three dimensions, GA does, saying that ‘the limit case, if attainable at all, could be reflected only by that which is perfect goodness itself—God Himself’ (O’Connor 2008 : 118). Mann makes a similar suggestion, saying that BPW might ‘be so full of so many goods that it is a candidate—in fact, a winner—in every cluster of worlds of incommensurable goods’ (Mann 1991 : 272).

What if there are incommensurable values that are not all capable of being maximized within a single possible individual? For example, suppose that happiness and organic unity are incommensurable yet fundamental values, and suppose that no individual can have both maximum happiness and maximum organic unity. On this picture, GA is BPW with respect to a limited set of values but GA is not BPW with respect to all values, just because there is no possible world containing only one individual that is BPW with respect to all values.

This alone will not yield an answer to WD, for it would only establish that GA is not BPW with respect to all values, not that some world other than GA is BPW. Suppose, however, that there is a meta-value with respect to possible worlds and that God seeks to maximize the meta-value. The greater the variety of fundamental, incommensurable values displayed in a world, the better a world is with respect to the meta-value. On this view, actualizing GA does not maximize the meta-value. If actualizing some world other than GA does maximize the meta-value—or even if doing so only, in some sense, makes there be ‘more’ of the meta-value than there would be if God actualized GA—then we will have a general answer to WD.

For this proposal to make sense there must be values displayed maximally in worlds other than GA that are not displayed maximally in GA and that are incommensurable with the values displayed maximally in GA. Plausibly there are many such values (e.g. diversity, harmony, novelty). But if those values are incommensurable with the values displayed in GA (e.g. simplicity) then an unacceptable result follows. Suppose the created part (CP) of a world displays to the maximal degree a value V 1 that cannot be displayed to the maximal degree in GA, which instead displays to a maximal degree a different value V 2 . Let us now switch from rating possible worlds to rating concrete beings. If we treat CP as a concrete being—the mereological sum of all of the created beings in that world—then it seems right to say CP displays V 1 to the maximal degree. So which is better—God or CP? On this line of thinking, there can only be a qualified answer to this question. God is better than CP with respect to V 2 , while CP is better than God with respect to V 1 . It is not true that God is better than CP tout court . So if one says GA is not BPW because there are incommensurable values and no one being is capable of displaying all of those values to the maximal degree, one must deny that God is without qualification the greatest possible being, since CP is also a possible being and it is just as great with respect to the values it displays as God is with respect to the values He displays. Theists confront a powerful dilemma here: deny that God is the greatest possible being tout court or admit that GA is BPW. For the reasons already given in this chapter, grabbing the second horn is admitting that there is no good answer to WD.

Consequences for the Design Argument

Unlike its cousin the Cosmological Argument, the Design Argument does not present its premises as logically guaranteeing that God exists, but only as giving strong inductive support for the claim that God exists. According to it, phenomena improbable or inexplicable on the hypothesis that there is no God are probable or explicable on the hypothesis that God exists: ‘Of course God would make a universe that is hospitable to life in which there are magnificent creatures with marvellous features like eyes and wings.’ That is what the proponent of the Design Argument needs us to believe. I hope to have shown in this chapter, however, that it is really quite a puzzle for traditional theists why God would design or create anything at all if, as natural theologians ask us to pretend, reason and observation alone are sufficient to support the premises of the Design Argument. Just considering the bare concept of God, and setting aside alleged revelation, there seems to be no obvious reason why God would create a universe and design some or all of its parts.

The difficulty can be presented in the form of a series of questions that can be asked of anyone who advances the Design Argument.

‘Are you saying God made the world because He needed something?’

‘Are you saying God made the world contrary to reason, or for no reason at all?’

‘Are you saying it is necessary that God made the world? Are you saying that God is not free not to make a world like ours?’

‘Are you saying God could not have some of His essential properties unless He made the world?’

‘Are you saying that a world consisting of God alone is less than the best of all possible worlds? Are you saying that reality is somehow inadequate unless it contains something more than God?’

It seems as though traditional theists are pushed to answer ‘no’ to every one of these questions. That list of ‘no’ answers makes it hard for them to then go on and advance the Design Argument. At the very least these deliberations show that one must do considerable work in natural theology in the broad sense before one can proceed to doing natural theology in the narrow sense.

Augustine ( 1887 and 1894). De Genesi ad litteram ( The Literal Commentary on Genesis ) in Patrologia Latina (PL). Vol. XXXIV. Edited by Jacques-Paul Migne . Paris: Garnier and in Corpus Scriptorum Ecclesiasticorum Latinorum (CSEL). Vol. XXVIII.1: Sancti Aureli Augustini opera . Edited by Joseph Zycha. Vienna: Tempsky.

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

Google Scholar

Google Preview

Barth, Karl ( 1961 ). Church Dogmatics: A Selection . Translated and edited by G. W. Bromiley. New York: Harper and Row.

Behe, Michael ( 1996 ) Darwin's Black Box: The Biochemical Challenge to Evolution . New York: Simon & Schuster.

Chappell, Timothy ( 1993 ). ‘ Why God is Not a Consequentialist ’. Religious Studies 29: 239–43.

Collins, Robin ( 2003 ). ‘Evidence for Fine-tuning’ in God and Design: The Teleological Argument and Modern Science . Edited by Neil A. Manson . London: Routledge: pp. 178–99.

Four, Vital du ( 1891 ). De Rerum Principio , in Joannis Duns Scoti Opera Omnia . Vol. IV. Paris: Vives.

Holder, Rodney ( 2004 ). God, the Multiverse, and Everything: Modern Cosmology and the Argument from Design . Hampshire: Ashgate.

Kretzmann, Norman ( 1991 a). ‘A General Problem of Creation: Why Would God Create Anything at All?’ in Scott MacDonald , ed., Being and Goodness: The Concept of the Good in Metaphysics and Philosophical Theology . Edited by Scott MacDonald. Ithaca, New York: Cornell University Press: pp. 208–28.

—— ( 1991 b). ‘A Particular Problem of Creation: Why Would God Create This World?’ in Being and Goodness: The Concept of the Good in Metaphysics and Philosophical Theology . Edited by Scott MacDonald . Ithaca, New York: Cornell University Press: pp. 229–49.

—— ( 1997 ). The Metaphysics of Theism: Aquinas's Natural Theology in Summa Contra Gentiles I . Oxford: Clarendon Press.

—— ( 1999 ). The Metaphysics of Creation: Aquinas's Natural Theology in Summa Contra Gentiles II . Oxford: Clarendon Press.

Leslie, John ( 1989 ). Universes . London: Routledge.

—— ( 2001 ). Infinite Minds: A Philosophical Cosmology . Oxford: Clarendon Press.

Lewis, David ( 1986 ). On the Plurality of Worlds . Oxford: Blackwell.

Mann, William E. ( 1991 ). ‘The Best of All Possible Worlds’ in Being and Goodness: The Concept of the Good in Metaphysics and Philosophical Theology . Edited by Scott MacDonald . Ithaca, NY: Cornell University Press: pp. 250–77.

Manson, Neil A. , ed. ( 2003 ). God and Design: The Teleological Argument and Modern Science . London: Routledge.

—— ( 2009 a). ‘ The Fine-Tuning Argument ’. Philosophy Compass 4/1: 271–86.

—— ( 2009 b). ‘The “Why Design?” Question’ in New Waves in Philosophy of Religion . Edited by Yujin Nagasawa and Erik Wielenberg . New York: Palgrave-MacMillan.

Mawson, Timothy J. ( 2005 ). Belief in God: An Introduction to the Philosophy of Religion . Oxford: Clarendon Press.

Nozick, Robert ( 1989 ). The Examined Life . New York: Simon & Schuster.

O’Connor, Timothy ( 2008 ). Theism and Ultimate Explanation . Oxford: Blackwell.

Oppy, Graham ( 2006 ). Arguing About Gods . Cambridge: Cambridge University Press.

Paley, William ( 2006 [1802]). Natural Theology . Oxford: Oxford University Press.

Pearl, Leon ( 1994 ). ‘ God Had to Create the World ’. Religious Studies 30: 331–3.

Plantinga, Alvin ( 1974 ). The Nature of Necessity . Oxford: Oxford University Press.

Plato ( 1902 ). Timaeus . Edited by J. Burnet. Oxford Classical Texts, Oxford: Clarendon Press.

Rowe, William ( 2004 ). Can God Be Free? Oxford: Oxford University Press.

Spinoza, Baruch ( 1982 [1677]). The Ethics and Selected Letters . Edited by Seymour Feldman and translated by Samuel Shirley. Indianapolis: Hackett.

Swinburne, Richard ( 2003 ). ‘The Argument to God from Fine-tuning Reassessed’ in God and Design: The Teleological Argument and Modern Science . Edited by Neil A. Manson . London: Routledge: pp. 105–23.

Teske, Roland J. (S.J.) ( 1988 ). ‘ The Motive for Creation According to St. Augustine ’. The Modern Schoolman LXV (May):, 245–53.

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

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

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

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

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

Home — Essay Samples — Literature — Literature Review — William Paley and the “Argument from Design”

test_template

William Paley and The "Argument from Design"

  • Categories: Existence of God Literature Review

About this sample

close

Words: 989 |

Published: Jul 18, 2018

Words: 989 | Pages: 2 | 5 min read

Works Cited

  • Collins, F. S. (2006). The Language of God: A Scientist Presents Evidence for Belief. Free Press.
  • Dawkins, R. (1986). The Blind Watchmaker: Why the Evidence of Evolution Reveals a Universe Without Design. W. W. Norton & Company.
  • Haught, J. F. (2004). God and the New Atheism: A Critical Response to Dawkins, Harris, and Hitchens. Westminster John Knox Press.
  • McGrath, A. (2007). The Dawkins Delusion?: Atheist Fundamentalism and the Denial of the Divine. SPCK Publishing.
  • Paley, W. (1802). Natural Theology: or, Evidences of the Existence and Attributes of the Deity. J. Faulder.
  • Plantinga, A. (2011). Where the Conflict Really Lies: Science, Religion, and Naturalism. Oxford University Press.
  • Reichenbach, B. R. (2012). The Cosmological Argument: A Reassessment. Cambridge University Press.
  • Swinburne, R. (2004). The Existence of God. Oxford University Press.
  • Ward, K. (1996). God, Chance, and Necessity. Oxford University Press.
  • Weisberg, J. (2008). The Argument from Design. In The Stanford Encyclopedia of Philosophy (Spring 2019 Edition). Retrieved from https://plato.stanford.edu/archives/spr2019/entries/design-argument/

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

Prof Ernest (PhD)

Verified writer

  • Expert in: Religion Literature

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

5 pages / 2067 words

3 pages / 1271 words

3 pages / 1487 words

1.5 pages / 801 words

Remember! This is just a sample.

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

121 writers online

Still can’t find what you need?

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

Related Essays on Literature Review

Flynn, Eilionoir. Disabled Justice?: Access to Justice and the UN Convention on the Rights of Persons with Disabilities. New York: Ashgate Publishing, 2015. Flynn focuses on the entire justice system regarding the issues [...]

Addiction is a profoundly complex and insidious issue that affects not only the individual struggling with substance abuse but also the entire network of family and loved ones. Joyce Maynard's novel, "Under The Influence," [...]

"Across a Hundred Mountains" by Reyna Grande is a poignant and evocative novel that explores the harrowing journey of immigration, the profound sense of loss experienced by those who leave their homeland, and the enduring hope [...]

Doyle’s essay uses science and vivid descriptions to depict how powerful yet fragile the organ is. The author’s purpose was to show that amongst all animals, they have a common feature; the heart. To convey this message, he [...]

Among the various definitions of tragedy, the one most commonly proffered is: a play that treats - at the most uncompromising level - human suffering, or pathos, with death being the usual conclusion. According to Aristotle's [...]

Traditional qualities of a feminine women usually include a beautiful physique, a gentle, nurturing nature, and a degree of sexual reservation. Throughout literature and film, women that embrace typical ideas of femininity are [...]

Related Topics

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

Where do you want us to send this sample?

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

Be careful. This essay is not unique

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

Download this Sample

Free samples may contain mistakes and not unique parts

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

Please check your inbox.

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

Get Your Personalized Essay in 3 Hours or Less!

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

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

the design argument essay

Marked by Teachers

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

Examine the design argument for the existence of God.

Authors Avatar

(a) Examine the design argument for the existence of God.

In this essay, I will be focusing on the different variations of the design argument, as expressed by Paley, as well as the version expressed by Hume, through Cleanthes, and its modern version in the form of the Anthropic Principle. I will describe the ideas of design ‘qua regularity’ and design ‘qua order’; explaining what each means. I will also attempt to deconstruct the design argument and identify both its strengths and weaknesses, though this will form part (b) of my essay. I will then draw my own conclusion in response to the design argument.

 The design argument is also known as the teleological argument.  This argument attempts to prove the existence of God, using analogy. The teleological argument is that of an a posteriori  nature, as its basic form comes from the idea that we can deduce knowledge from what we know or have sense of, in this case, the existence of the universe. Even before the Greek philosophers, people have presumed that the natural order of the universe – and the way in which it works, i.e.the changing season, the intricacy of an eyeball, the complexity of the human brain – serves as evidence of a designer (who is often asserted to be God).

The design argument has been expressed in two ways – ‘design qua regularity ’, and ‘design qua purpose ’. Firstly, we will focus on design qua regularity. This aspect looks at design in relation to the order and regularity in the universe. Philosophers who support the argument would believe that it is evident that there is some sort of order in the universe - for example, the rotation of the planets – and that it cannot have occurred by random chance. This therefore suggests existence of a creator: God. St. Thomas Aquinas  argues from design qua regularity in the fifth of his five ways, identifying that the way in which ‘natural bodies’ act in a regular fashion provides the evidence for the existence of a creator: ‘…there must be some intelligent being who directs all things to the purpose for which they exist. This being we call god.’

 The other part of the argument – design qua purpose – looks at the evidence of design in relation to the complexity of everything in the universe and that things work because they have a purpose. This is used as evidence for a designer (God), as a purpose would imply that something or someone has given an object that purpose.

I will begin by outlining William Paley’s version of the design argument. Paley   puts forward the most famous form of the design argument, using analogy . Paley presented the following situation; in order to consolidate his argument: ‘…In crossing a heath, suppose I pitched my foot against a stone, and were asked how the stone came to be there, I might possibly answer, that for any thing I knew to the contrary it had lain there forever…But suppose I had found a watch upon the ground, and it should be inquired how the watch happened to be in that place, I should hardly think of the answer which I had before give, that for any thing I knew the watch might have always been there…when we come to inspect the watch, we perceive – what we could not discover in the stone – that its several parts are framed and put together for a purpose…’ .

The first part of Paley’s argument is design qua purpose – and is to show that there must be a purpose. Paley’s analogy attempts to compare the universe to a watch; and states that the watch is so complex – it has parts that all work together and it itself works and is in motion (much like the universe) and because the watch is so complex and is in motion; it is logical to ask who or what created this machine; and we can presume that there was a designer – or shall we say, a watchmaker. Paley used the idea that   like effects have like causes, and therefore came to the conclusion that the universe had a universe-maker – asserted, by Paley, to be God. But Paley also mentions a stone – and that questioning where the stone came from is different – as the stone has no complexity, does not change, is not in constant motion, and seems to have no purpose; therefore, it is acceptable to presume that the stone had existed forever.

Join now!

This is a preview of the whole essay

The second part of the argument for the existence of God is design qua regularity – Paley used evidence from astronomy and Newton’s laws of motion and gravity to prove that there is design in the universe. The universal laws, such as gravity and the rotation of the planets, were for Paley, evidence that the universe had not been an accident; as they provide the universe with order and regulation. Paley argues that this order and regulation is so complex that the universe can’t have come about by chance; and that it implies a universe-maker in the same way that the watch had a watch-maker. This universe-maker can be called God.

Consider this diagram.

        I will now rehearse the design argument as expressed by David Hume , through Cleanthes, using analogy. Hume follows Plato’s method and writes in the form of a debate between two main characters: Cleanthes and Philo. Cleanthes is used to put forward the argument; and Philo is then used to put forward the critique and demonstrate that Cleanthes is wrong.

        Hume’s argument is simply based from analogy and relies heavily on a posteriori deductions – assumptions that we are able to make using our knowledge of the universe and the way things work now. However, because Hume’s argument has no initial evidence to prove or disprove his theories and beliefs; it is obviously very easy to pick flaws within it. These flaws are identified by the character Philo (as mentioned above) and will be described further in part (b) of the essay.

Hume, or rather, Cleanthes , basically argued that we can observe material things in our world – for example, houses, paintings and machines – created by humans; and that these things have the common features of order and being produced by intelligent design. It is natural to conclude that orderly things are produced by intelligence. We know from scientific evidence that the universe is complex; and we know through our own experience that the universe exhibits the same kind of complex order that houses, paintings and machines do. Therefore; we can presume that the universe was produced by intelligent design (intelligence) also – which encourages us to believe in a creator – and this creator is asserted to be God.

The teleological argument has recently been developed and from this recent development we are now aware of the theory of the Anthropic  Principle. The Anthropic Principle  is basically an argument which states that the chances of the universe existing – and existing in the way that it does (i.e. with human life) – are so remote, that there must have been a design built into the universe in order for this production to take place; in other words: ‘The Universe seems like a put up job.’  The Anthropic Principle claims that if there had been a minute change in the values of the strong nuclear force, or perhaps the charge of an electron – any life form would have been unlikely to develop. In its simplest form; the Anthropic Principle opposes the idea that there is any chain of coincidences that led to the evolution of human life; and again, because of the complexity of our life, and the infinitesimal chances of the universe and our planet evolving as it has done, believers in the Anthropic Principle would conclude that this serves as evidence that a creator has been at work – and that this creator is/was God.

As aforementioned, the Anthropic Principle was developed by F. R. Tennant. Tennant believed that there were three types of evidence supporting the idea of a designer (God) (which he deduced from his knowledge of the world now  as opposed to factual evidence of events which occurred then ). These are:

  • The fact that the world can be analysed in a rational manner
  • The way in which the inorganic world has provided the basic necessities required to sustain life
  • The progress of evolution towards the emergence of intelligent human life

Basically, Tennant argues that because life is as it is – because we can see it for what it is and because we can survive in our universe as it is – there must have been a creator of this life and of the universe.

‘The fact is, we are here, and here by the grace of some pretty felicitous arrangements. Our existence cannot of itself explain these arrangements. One could shrug the matter aside with the comment that we are certainly very lucky that the universe just happens to possess the necessary conditions for life is flourish, but that this is a meaningless quirk of fate. Again, it is a question of personal judgement.’

Tennant also used the idea of aesthetic  appreciation to further consolidate his argument. He stated that humans have the ability to appreciate the beauty of our world – and things such as art, music and literature – and this aesthetic appreciation is not a necessity. Tennant argues that it is not vital to regard things as beautiful and certainly not necessary for evolutionary survival, and yet, humans do – so there must have been a ‘divine creator’ of the universe, who determined that humans would have specific attributes in addiction to those required for basic survival.

Although Tennant regarded his argument as valid, it was not particularly appreciated by people at that time. It was the developments in cosmology, theoretical physics and pure mathematics, which urged philosophers – such as Richard Swinburne – to re-examine it.

        Swinburne argues, using probability , that it is much more likely for the universe to have been created through design as opposed to having been created by random chance. Swinburne again reiterated that the complexity and the order of the universe support the theory that there must have been some form of design: and he concludes that this design’s simplest explanation is that of God.

        In conclusion, I have discovered the design argument is built up of many points and has been varied several times, however, each of the variations seem very similar – and all come down to the same one point: that the universe is so very complex that there must have been a creator – however, as the argument is basically based around probability, it seems easy to argue against it and/or disprove such a theory. These critiques, along with any strengths of the argument, will form part (b) of my essay.

(b) What are the strengths of the design argument?

Comment on some of the criticisms raised against the design argument.

After describing the Design Argument, and the Anthropic Principle, I now intend to focus on the strengths of the argument, and then I will comment on the criticisms of the argument, as put forward by various philosophers. I will consider various points of view and not hesitate to incorporate some of my own ideas.

I will begin by considering the strengths of the argument. We know that the design argument is an a posteriori argument (using evidence of the world now to create theories of past activity) and therefore the evidence provided is obvious to humankind – as we can witness it and understand it and deduce things from it. This is a strong point, I think, as we know that the world does exist and we know that our world is complex. The argument is strong in the way that it uses analogy to consolidate the basic idea – we can witness the complexity of our own beings, or of different material things – and can easily connect this back to the idea of our universe – observing the possible likenesses between the two. The argument itself does follow an order, and initially the argument seems valid – and the argument involving the analogy of a watch is thorough and easy to comprehend. In fact, Immanuel Kant has described the argument as the oldest, clearest and most reasonable argument for the existence of God, even though he himself found it unconvincing. Even science cannot be used to disprove the design argument – in fact; it can be used to support it. There is nothing to say that the process of evolution is not the way in which God decided to create his design.

However, there are many arguments against the design argument as well as arguments against the anthropic principle. I will begin by examining and commenting on the flaws and criticisms of the design argument. To begin with, the whole idea of the design argument and its basis mainly on probability weakens the argument. It is probable  that there was a God – this is because of the order and complexity of the universe. However, probability does not serve as substantial evidence for a God. The universe could in fact be the result of a committee of Gods, if you will: the comparison of the world to a building supports this idea – many buildings are designed by one person, and yet, constructed by many others.

Paley used analogy to make comparisons between a watch and the universe. However, this analogy may not appear valid – how can two things of perhaps different laws and time be compared in such a way? And even though the analogy stands, we can still not presume that the same laws or laws of our space and time can be applied to it: just because the watch had a creator, it does not necessarily follow that the universe had a creator. We cannot make assumptions using our knowledge, as we have no knowledge of anything before our existence.

David Hume  specifically emerged as a major opponent of the design argument, and found several flaws with Paley’s version. Hume argued that:

  • We do not have sufficient knowledge of the creation of the universe to conclude that there was only one creator; and humans only have the experience of what it is for humans to create, not what it is for a creator to create or what it is to create the universe
  • If the human experience of design was valid, the design argument would be valid, but prove that the universe had a designer, not that the designer was the God of Classical Theism. It cannot be believed that the Theistic God created the universe because there is the problem of evil
  • The very existence of evil in the universe would therefore suggest a designer who is not the omnipotent, omniscient and/or omnipresent God
  • With regards to the analogy put forward by Paley: if we are going to use the analogy of a manufactured object (the watch) then this would suggest many creators (many Gods) as opposed to one creator (one God) as it is more usual for a machine to be designed and created by many hands
  • The analogy of the watch is weak and, the universe would be better compared to a vegetable – something that grows of its own accord as opposed to something made by hand that does not evolve erratically (leads back to problem of evil)

So, if the design argument is to be considered as valid, then, Hume states, God cannot have the attributes he is given, he cannot be all-loving, all-knowing and all-powerful, because of the problem of evil. Surely an all-loving, all-powerful, all-knowing God would be able to prevent evil and suffering in the world – and though this maybe argued against with the idea that evil is in the world because we have free will, surely natural disasters such as earthquakes and volcanoes would be prevented by a God who is all-loving, all-knowing and all-powerful. In any case, if there was a God, Hume argues that he could not be the God of Theism or else there would not be evil in the universe.

I will now comment on the criticisms of the Anthropic Principle. The basic criticism of the Anthropic Principle is concerning the basis of the argument – probability. I don’t think that the probability of something happening is evidence enough for a designer: although it is probable  that when I roll a dice, I will get a number less than 6, this is not a certainty – and I could well get a 6, although the chances are smaller than me getting a 5, 4, 3, 2 or 1. This may be argued against by the fact that the chances of the universe forming as it has done were infinitesimal; however, do we actually have the knowledge to predict that if the slightest thing had changed, the whole universe would be entirely different? Or that perhaps human life wouldn’t have evolved at all?

If there was a God who created the world for us to appreciate things such as aesthetics, what were his  motives? What did he want of humans? What is the meaning of life? The truth is, the Anthropic Principle causes us to ask more questions: what sort of God created humankind? And there is no logical explanation as to why there was a single, personal, transcendent God with qualities such as being omnipotent, omniscient etc. To make that much of a presumption without justification cannot be valid – where is any evidence for such a being?

In conclusion, I think both the Design Argument and the Anthropic Principle – although based upon probability rather than reality or evidence – can be understood and accepted by most philosophers and humans: the universe really is complex, and how is it so that our universe may have just fallen into place in such a precise way? It seems unlikely; and the universe now – in its current form – certainly points to the idea of a designer. However, whether this designer was/is God is another argument and personally I find the way in which Paley, for example, manages to jump from the idea of a designer to the idea of God rather illogical. How can there be a Theistic God when it is evident that we have suffering in the world? ...

‘This world, for all he knows, is very faulty and imperfect compared to a superior standard; and was only the first rude essay of some infant deity

who afterwards abandoned it.”

Bibliography

  • AS Design sheets –

      Mr Hickman

  • Philosophy of Religion for A Level  

     Anne Jordan, Neil Lockyer, Edwin Tate

  • An Introduction To Philosophical Analysis (third edition) -

      John Hospers

  • Design Argument RIP -  

     

  • A Critique of the Design Argument – Trevor Stone
  • Argument for the Existence of God -  
  • The Design Argument –

      http://saif_w.tripod.com/explore/i4wm/02k.htm

      http://www.aish.com/spirituality/philosophy/The_Design_Argument.asp

  The Greek word teleos can be translated as ‘end’ or ‘purpose’.

  Meaning ‘at the end’ – something that is known to be true through our own experience of it.

  The Christian theologian and philosopher (1225-1274) who put forward his ‘Five Ways’ to prove the existence of God.

  St Thomas Aquinas, Summa Theologica

  (1743-1805) A British philosopher-theologian. Paley was an apologist – meaning that he tried to demonstrate from human reason that Christianity (and the existence of God) was true.

  Arguing by analogy is to argue that since things are alike in some ways, they will probably be alike in others also.

  William Paley, Natural Theology

  (supposing that our arrows mean creator of)

  (1711-1776) regarded as Britain’s greatest philosopher

  Meaning to be linked to the science and study of mankind.

  First developed by F. R. Tennant (1866-1957) and followed by philosophers (i.e. Swinburne) and cosmologists (i.e. Davies). ‘Anthropos’ is a Greek word, translating as ‘man’ in the generic sense.

  Sir Fred Hoyle, cosmologist

  Bullet points taken from Philosophy of Religion: Anne Jordan, Neil Lockyer, Edwin Tate, p 77

  The Mind of God, Paul Davies, Penguin, 1993, p 204

  Having an appreciation of beauty

 the extent to which something is probable; the likelihood of something happening or being the case

  deconstructed the design argument in Dialogues Concerning Natural Religion (1779)

  not meaning to give God a gender (not meaning he  as male)

  (in response to Paley’s Design Argument). David Hume, Dialogues Concerning Natural Religion, 1779

Examine the design argument for the existence of God.

Document Details

  • Word Count 3498
  • Page Count 7
  • Subject Religious Studies (Philosophy & Ethics)

Related Essays

Examine the design argument for the existence of God.

Examine the design argument for the existence of god.

Examine the design argument for the existence of God.

  • Trending Now
  • Foundational Courses
  • Data Science
  • Practice Problem
  • Machine Learning
  • System Design
  • DevOps Tutorial
  • Email Writing - Format and Samples
  • English Essay Writing Tips, Examples, Format
  • Letter to Principal, Format And Samples
  • Analytical Writing Section in GRE General
  • 12 Best ChatGPT Prompts for Academic Writing Assistance in 2024
  • Writing data from a Python List to CSV row-wise
  • CBSE Sample Papers for Class 9 Social Science Set 2 with Solutions
  • A Guide to Writing an Essay for Job Interviews
  • CBSE Sample Papers for Class 11 History Set 2 with Solutions 2023-24
  • How to prepare for IELTS?
  • 5 Content Writing Tools to Improve your Content Writing Skills
  • IBPS Clerk Prelims Reasoning Question Paper 2020
  • CBSE Sample Papers for Class 11 Economics (2023-24) Set 1 with Solutions
  • IBPS Clerk Prelims Reasoning Question Paper 2021
  • IBPS Clerk Prelims English Question Paper 2019
  • IBPS Clerk Prelims English Question Paper 2021
  • Fill In The Blanks - Rules, Tips and Tricks with Examples
  • CBSE Sample Papers Class 11 History (2023-24) Set-1 with Solution
  • UPSC Prelims 2018 General Studies Paper I With Detailed Solutions

IELTS Writing Task 2: Format, Sample, Tips

The IELTS Writing Task 2: The second portion of the writing test, known as IELTS Writing Task 2, asks you to produce an essay in response to a point of view, argument, or problem. Your essay should be written in a formal tone, be at least 250 words long, and take no more than 40 minutes to finish.

Table of Content

IELTS Writing Task 2- Format

1. task question, 2. word limit, 4. response structure, 5. evaluation criteria, difference between ielts writing task 2- academic vs general, understanding the evaluation criteria, common ielts writing task 2 topics, band descriptors ielts writing task 2, ielts essay types for writing task 2, ielts writing task 2 preparation tips, ielts writing task 2 sample, ielts writing task 2- faqs, what are indigenous cultures and languages, why is it important to protect indigenous cultures and languages, what are some challenges in protecting indigenous cultures and languages, what role can governments play in protecting indigenous cultures and languages, are there any potential drawbacks to prioritizing the protection of indigenous cultures and languages.

  • You will be presented with a topic or statement related to a contemporary issue or problem.
  • The task question may ask you to discuss a particular problem, present a solution, evaluate a situation, or provide your opinion on a given topic.
  • You are expected to write at least 250 words for the IELTS Writing Task 2.
  • It is advisable to write within the range of 250300 words, as responses shorter than 250 words are penalized, and longer responses do not necessarily receive higher scores.
  • 3. Time Allotment:
  • You have 40 minutes to complete the IELTS Writing Task 2.
  • Your response should be structured as an essay with an introduction, body paragraphs, and a conclusion.
  • The introduction should provide an overview of the topic and outline the main points you will discuss.
  • The body paragraphs should develop your ideas and arguments, with one main idea per paragraph supported by relevant examples or evidence.
  • The conclusion should summarize your main points and provide a final perspective on the topic.
  • Your response will be evaluated based on four criteria: Task Response, Coherence and Cohesion, Lexical Resource (vocabulary), and Grammatical Range and Accuracy.
  • You should aim to address all parts of the task question, present a clear and coherent argument, use a wide range of vocabulary accurately, and demonstrate a good command of grammar and sentence structures.
Must Read: IELTS Academic vs General Tests – What’s the Difference?

The IELTS Writing Task 2 covers a wide range of topics related to contemporary issues and problems. Here are some common topics that frequently appear in the IELTS Writing Task 2:

1. Education:

  • The role of technology in education
  • The importance of extracurricular activities
  • The advantages and disadvantages of single gender schools

2. Environment:

  • Climate change and its impacts
  • Sustainable development and environmental conservation
  • The use of renewable energy sources
  • The impact of lifestyle choices on health
  • The role of government in promoting public health
  • The advantages and disadvantages of alternative medicine

4. Society and Culture:

  • The effects of globalization on local cultures
  • The impact of social media on human interactions
  • The role of religion in modern society

5. Technology:

  • The advantages and disadvantages of artificial intelligence
  • The impact of technology on employment and job markets
  • The role of technology in communication and information sharing

6. Urbanization and Transportation:

  • The challenges of urban growth and city planning
  • The benefits and drawbacks of public transportation
  • The impact of transportation on the environment

7. Crime and Justice:

  • The causes and prevention of crime
  • The effectiveness of different types of punishment
  • The role of the criminal justice system in society

8. Economics and Business:

  • The impact of globalization on international trade
  • The role of advertising in influencing consumer behavior
  • The advantages and disadvantages of outsourcing

9. Government and Politics:

  • The importance of freedom of speech and press
  • The role of government in regulating the economy
  • The impact of immigration on societies

10. Arts and Culture:

  • The importance of preserving cultural heritage
  • The role of art in society
  • The impact of censorship on artistic expression

In the IELTS Writing Task 2, candidates are required to write an essay in response to a prompt or question. There are several common types of essays that may appear in Task 2:

1. Argumentative/Opinion Essays: These essays require candidates to express their opinion on a given topic and support it with reasons and examples. They often involve discussing both sides of an issue and presenting a clear argument in favor of one viewpoint.

2. Discussion/Two-sided Essays: Similar to argumentative essays, discussion essays require candidates to discuss both sides of an issue before expressing their opinion or preference. They need to provide balanced arguments and consider opposing viewpoints.

3. Advantages and Disadvantages Essays : In these essays, candidates need to discuss the pros and cons of a particular issue, situation, or trend. They should provide examples to illustrate each point and offer a balanced analysis.

4. Problem-Solution Essays: These essays involve identifying a problem or issue, discussing its causes and effects, and proposing possible solutions or measures to address it. Candidates need to present logical arguments and support their solutions with evidence.

5. Cause and Effect Essays: Cause and effect essays focus on analyzing the reasons behind a specific phenomenon or event and its subsequent effects. Candidates should clearly outline the causal relationships and provide relevant examples.

6. Comparison/Contrast Essays: These essays require candidates to compare and contrast two or more ideas, concepts, or approaches. They should highlight similarities and differences and draw conclusions based on their analysis.

7. Process Essays: Process essays explain a sequence of steps or actions involved in a particular process, such as how to do something or how something works. Candidates need to provide clear explanations and use appropriate transition words to guide the reader through each step.

8. Agree/Disagree Essays: In these essays, candidates are given a statement or opinion, and they need to express whether they agree or disagree with it. They should support their stance with reasons and examples.

IELTS Writing Task 2 preparation tips to help you improve your performance:

1. Understand the Task Question

  • Read the task question carefully and identify the key components, such as the topic, the instructions (e.g., discuss, evaluate, give your opinion), and any specific aspects to be addressed.
  • Underline or highlight the essential elements to ensure you address all parts of the question.

2. Plan Your Essay

  • Spend a few minutes planning your essay before you start writing.
  • Brainstorm ideas and organize them into an introduction, body paragraphs, and a conclusion.
  • Develop a clear thesis statement and main points to guide your essay.

3. Manage Your Time

  • Allocate your time wisely, allowing enough time for planning, writing, and reviewing.
  • Aim to spend around 510 minutes planning, 2530 minutes writing, and 5 minutes reviewing and making corrections.

4. Use Appropriate Structure and Paragraphing

  • Follow a standard essay structure with an introduction, body paragraphs, and a conclusion.
  • Each body paragraph should focus on one main idea and include supporting details, examples, or evidence.
  • Use clear topic sentences and logical transitions between paragraphs.

5. Develop Your Ideas

  • Provide relevant and welldeveloped ideas to support your main points.
  • Use examples, personal experiences, facts, or hypothetical situations to illustrate your arguments.
  • Show critical thinking by analyzing different perspectives and addressing counterarguments.

6. Use Appropriate Language and Vocabulary

  • Use a range of appropriate vocabulary related to the topic.
  • Vary your sentence structures and avoid repetition.
  • Demonstrate your ability to use idiomatic expressions and collocations accurately.

7. Pay Attention to Grammar and Accuracy

  • Review and proofread your essay for grammatical errors, spelling mistakes, and punctuation issues.
  • Ensure subjectverb agreement, correct tense usage, and appropriate word forms.
  • Avoid overly complex sentences that may increase the risk of errors.

8. Practice with Sample Questions

  • Familiarize yourself with different types of IELTS Writing Task 2 questions by practicing with sample prompts.
  • Set a timer and practice writing complete essays under timed conditions.
  • Seek feedback from experienced IELTS teachers or online resources to identify areas for improvement.

9. Learn from Model Answers

  • Study highscoring model answers to understand the expected level of writing and the organization of ideas.
  • Analyze the structure, language use, and development of arguments in these model answers.
  • Incorporate effective strategies and techniques into your own writing practice.

10. Stay UptoDate with Current Affairs

  • Stay informed about current events, global issues, and debates related to various topics.
  • Read reputable news sources, magazines, or online articles to broaden your knowledge and enhance your ability to discuss contemporary topics.
Here is a practice IELTS Writing Task 2 topic for you: Topic: Some people believe that governments should make more efforts to protect indigenous cultures and languages from disappearing. To what extent do you agree or disagree with this statement? You should spend about 40 minutes on this task. Write at least 250 words discussing both viewpoints and giving your opinion.
  • Make a plan before you start writing. Outline your introduction, body paragraphs and conclusion.
  • The introduction should paraphrase the topic and outline what will be discussed.
  • Discuss both sides of the argument in the body paragraphs. One paragraph arguing for protecting indigenous cultures/languages, one paragraph arguing against or giving the opposite view.
  • Use examples, data or personal experiences to support your arguments.
  • The conclusion should summarize your main points and give a clear opinion.
  • Use a range of vocabulary and sentence structures. Avoid repetition.
  • Check for grammar, spelling and punctuation errors.

In conclusion, while protecting indigenous cultures and languages is undoubtedly important for preserving human diversity and heritage, it should be balanced with practical considerations and the interests of the wider community. A nuanced approach that promotes understanding and appreciation while accommodating evolving societal needs is ideal.

Also Read: IELTS Full Form: Check Its Significance IELTS Average Score: Across Worldwide and India IELTS Minimum Score for Top Universities in 2024 IELTS Exam Pattern 2024: Section-wise IELTS Exam Paper Pattern, Question Types
Indigenous cultures and languages refer to the traditional practices, belief systems, and modes of expression of ethnic groups native to a particular region or country.
Protecting indigenous cultures and languages helps preserve unique identities, traditional knowledge, and cultural diversity, which are valuable aspects of human heritage and can contribute to our understanding of history, societies, and the environment.
Challenges include globalization, urbanization, lack of resources, and a shift towards more dominant cultures and languages, which can lead to the erosion of indigenous practices and languages over time.
Governments can implement policies to support the use and teaching of indigenous languages, provide funding for cultural preservation efforts, and promote awareness and appreciation of indigenous cultures through education and media.
Potential drawbacks include the allocation of limited resources towards this effort at the expense of other priorities, the potential for cultural stagnation or resistance to cultural evolution, and the risk of creating divisions or conflicts within diverse societies.

Please Login to comment...

Similar reads.

  • Study Abroad

Improve your Coding Skills with Practice

 alt=

What kind of Experience do you want to share?

Man charged with helping to fraudulently acquire Laguna Beach home

A man has been charged with a fraudulent purchase of a home in the Victoria Beach neighborhood in Laguna Beach.

  • Show more sharing options
  • Copy Link URL Copied!

A 34-year-old man pleaded not guilty Tuesday to conspiring with several others to fraudulently acquire a Laguna Beach home worth about $2 million.

Vicente Anzu was charged in June 2023 to 23 felony counts including conspiracy, identity theft, procuring or offering false or forged documents to be filed, registered or recorded, forgery and mortgage fraud.

Laguna Beach Police Department Det. Kyle Milot said in court papers that Anzu and four other defendants “unlawfully conspired against victims ... to commit a grand theft in which they fraudulently acquired the property at 2845 Wards Terrace. Anzu forged the signature of real persons and recorded false and forged instruments in an effort to carry out the acquisition of the Wards Terrace property and later attempted to monetize the property through loans or a sale of the property.”

The defendant’s wife, Monica, wrote a letter to court officials last year saying her husband couldn’t respond to an arraignment in the case because “he was due at the Maricopa Courthouse in Arizona for sentencing on the same charges.”

She said he was sentenced in Arizona to five years in prison with his release date expected for October 2027.

Co-defendant Deric Ryals, 59, was next due in court for a pretrial hearing July 1 in the Harbor Justice Center in Newport Beach.

Co-defendants Jermaine Bush, 51, Selwyn Benjamin, 54, and Edmound Daire, 66, were listed as fugitives who have not yet appeared in court on the case.

All the latest on Orange County from Orange County.

Get our free TimesOC newsletter.

You may occasionally receive promotional content from the Daily Pilot.

City News Service is the nation’s largest regional wire service and is headquartered in Los Angeles.

More on this Subject

siren two

Bicyclist killed in Fountain Valley hit-and-run

Patients, staff, public officials, physicians, and administrators gather in front of the new City of Hope, Seacliff location for ribbon-cutting and grand opening ceremony in Huntington Beach on Tuesday morning.

City of Hope expands with Seacliff facility in Huntington Beach

May 15, 2024

A potential construction project involving the Balboa Branch Library, may once again threaten the future of the Eucalpytus tree that holds blue heron nests behind the library in Newport Beach.

Balboa Library and Fire Station No. 1 project moves forward to final design

Ceramicist Agnes Noble joins husband Don with her functional clay art pieces.

Nature and art merge for spring show in Corona del Mar

Rusmania

  • Yekaterinburg
  • Novosibirsk
  • Vladivostok

the design argument essay

  • Tours to Russia
  • Practicalities
  • Russia in Lists
Rusmania • Deep into Russia

Out of the Centre

Savvino-storozhevsky monastery and museum.

Savvino-Storozhevsky Monastery and Museum

Zvenigorod's most famous sight is the Savvino-Storozhevsky Monastery, which was founded in 1398 by the monk Savva from the Troitse-Sergieva Lavra, at the invitation and with the support of Prince Yury Dmitrievich of Zvenigorod. Savva was later canonised as St Sabbas (Savva) of Storozhev. The monastery late flourished under the reign of Tsar Alexis, who chose the monastery as his family church and often went on pilgrimage there and made lots of donations to it. Most of the monastery’s buildings date from this time. The monastery is heavily fortified with thick walls and six towers, the most impressive of which is the Krasny Tower which also serves as the eastern entrance. The monastery was closed in 1918 and only reopened in 1995. In 1998 Patriarch Alexius II took part in a service to return the relics of St Sabbas to the monastery. Today the monastery has the status of a stauropegic monastery, which is second in status to a lavra. In addition to being a working monastery, it also holds the Zvenigorod Historical, Architectural and Art Museum.

Belfry and Neighbouring Churches

the design argument essay

Located near the main entrance is the monastery's belfry which is perhaps the calling card of the monastery due to its uniqueness. It was built in the 1650s and the St Sergius of Radonezh’s Church was opened on the middle tier in the mid-17th century, although it was originally dedicated to the Trinity. The belfry's 35-tonne Great Bladgovestny Bell fell in 1941 and was only restored and returned in 2003. Attached to the belfry is a large refectory and the Transfiguration Church, both of which were built on the orders of Tsar Alexis in the 1650s.  

the design argument essay

To the left of the belfry is another, smaller, refectory which is attached to the Trinity Gate-Church, which was also constructed in the 1650s on the orders of Tsar Alexis who made it his own family church. The church is elaborately decorated with colourful trims and underneath the archway is a beautiful 19th century fresco.

Nativity of Virgin Mary Cathedral

the design argument essay

The Nativity of Virgin Mary Cathedral is the oldest building in the monastery and among the oldest buildings in the Moscow Region. It was built between 1404 and 1405 during the lifetime of St Sabbas and using the funds of Prince Yury of Zvenigorod. The white-stone cathedral is a standard four-pillar design with a single golden dome. After the death of St Sabbas he was interred in the cathedral and a new altar dedicated to him was added.

the design argument essay

Under the reign of Tsar Alexis the cathedral was decorated with frescoes by Stepan Ryazanets, some of which remain today. Tsar Alexis also presented the cathedral with a five-tier iconostasis, the top row of icons have been preserved.

Tsaritsa's Chambers

the design argument essay

The Nativity of Virgin Mary Cathedral is located between the Tsaritsa's Chambers of the left and the Palace of Tsar Alexis on the right. The Tsaritsa's Chambers were built in the mid-17th century for the wife of Tsar Alexey - Tsaritsa Maria Ilinichna Miloskavskaya. The design of the building is influenced by the ancient Russian architectural style. Is prettier than the Tsar's chambers opposite, being red in colour with elaborately decorated window frames and entrance.

the design argument essay

At present the Tsaritsa's Chambers houses the Zvenigorod Historical, Architectural and Art Museum. Among its displays is an accurate recreation of the interior of a noble lady's chambers including furniture, decorations and a decorated tiled oven, and an exhibition on the history of Zvenigorod and the monastery.

Palace of Tsar Alexis

the design argument essay

The Palace of Tsar Alexis was built in the 1650s and is now one of the best surviving examples of non-religious architecture of that era. It was built especially for Tsar Alexis who often visited the monastery on religious pilgrimages. Its most striking feature is its pretty row of nine chimney spouts which resemble towers.

the design argument essay

Plan your next trip to Russia

Ready-to-book tours.

Your holiday in Russia starts here. Choose and book your tour to Russia.

REQUEST A CUSTOMISED TRIP

Looking for something unique? Create the trip of your dreams with the help of our experts.

IMAGES

  1. The Design Argument

    the design argument essay

  2. Essay discussing Hume and the design argument

    the design argument essay

  3. A design argument essay question

    the design argument essay

  4. Design Argument

    the design argument essay

  5. (PDF) The Design Argument

    the design argument essay

  6. Design Argument essay.docx

    the design argument essay

VIDEO

  1. Academic reading and writing in English Part 10: Building logical arguments

  2. The Three Stages of the Design Argument! #Shorts

  3. Framing an argument #argument #debate #empathy

  4. AtheistDebates

  5. Design Is the Way Forward

  6. The History of the Design argument

COMMENTS

  1. Design Arguments for the Existence of God

    There are a number of classic and contemporary versions of the argument from design. This article will cover seven different ones. Among the classical versions are: (1) the "Fifth Way" of St. Thomas Aquinas; (2) the argument from simple analogy; (3) Paley's watchmaker argument; and (4) the argument from guided evolution.

  2. The Design Argument for the Existence of God

    The design argument is the simplest, most straightforward argument for the existence of God. Unlike the cosmological argument, the design argument can be stated in a few, easy-to-understand steps. In a nutshell, the design argument claims that the fact that everything in nature seems to be put together in just the right manner suggests that an ...

  3. The Argument from Design: A Guided Tour of William Paley's

    According to the classic "argument from design," observations of complex functionality in nature can be taken to imply the action of a supernatural designer, just as the purposeful construction of human artifacts reveals the hand of the artificer. The argument from design has been in use for millennia, but it is most commonly associated with the nineteenth century English theologian ...

  4. An Overview of the Design Argument and its Opponents

    Abstract. The design argument, also known as the teleological argument, is a philosophical concept that asserts that the complexity and order of the natural world imply the existence of an intelligent designer. This essay provides an introduction to the design argument, specifically examining the versions put forth by Thomas Aquinas and William ...

  5. Design Arguments for the Existence of God

    This essay introduces design arguments for the existence of God. 1. The Arguments. The standard 'Design' or 'Teleological' arguments for theism hold that there is evidence of design in nature and that this is evidence for the existence of God.[1] There are three general versions of this argument:

  6. The existence of God The design argument

    The design argument. This is an argument for the existence of God. It points to evidence that suggests our world works well - ie that it was designed in a specific way. The argument follows that ...

  7. Teleological argument

    The teleological argument (from τέλος, telos, 'end, aim, goal'; also known as physico-theological argument, argument from design, or intelligent design argument) is an argument for the existence of God or, more generally, that complex functionality in the natural world which looks designed is evidence of an intelligent creator. [1] [2] [3] [4]

  8. Teleological Arguments for God's Existence

    Teleological arguments (or arguments from design) by contrast begin with a much more specialized catalogue of properties and end with a conclusion concerning the existence of a designer with the intellectual properties (knowledge, purpose, understanding, foresight, wisdom, intention) necessary to design the things exhibiting the special ...

  9. The Design Argument (Chapter 6)

    The design argument is one of three main arguments for the existence of God; the others are the ontological argument and the cosmological argument. Unlike the ontological argument, the design argument and the cosmological argument are a posteriori. And whereas the cosmological argument can focus on any present event to get the ball rolling ...

  10. The Argument from Design

    The First Premise: Apparent Design. The claim that natural objects exhibit the appearance of design can be understood in a number of ways. Some versions of the argument stress the structural complexity or intricacy of many natural things: crystals, cells, spider webs and the like are all remarkably complex, but they are also remarkably regular or orderly. This notion of orderly complexity is ...

  11. God and Design: the Teleological Argument and Modern Science

    Sober raises critical points against both the biological and the cosmological ("fine tuning") versions of the argument from design. This essay is followed by John Leslie's reflections on the meaning of design. Where Sober's essay was rooted in rigorous analysis of the structure of arguments trading in probabilities, Leslie's essay is ...

  12. PDF An Introduction to Design Arguments

    5.1 Nieuwentyt's design argument, which is a kind of argument from order 71 6.1 Berkeley's 'Visual Language' argument 88 7.1 Cleanthes' analogical argument diagramed 105 7.2 An argument from order diagramed 105 8.1 Paley's design argument interpreted as an inference to the best explanation 129

  13. The Design Argument

    The Design Argument. Published online by Cambridge University Press: 07 December 2018. Elliott Sober. Show author details. Elliott Sober. Affiliation: University of Wisconsin, Madison. Summary. This Element analyzes the various forms that design arguments for the existence of God can take, but the main focus is on two such arguments.

  14. The Design Argument

    This chapter contains sections titled: What is the Design Argument. Clarifications. Other Formulations of the Design Argument, and Their Defects. Three Possible Objections to the Likelihood Argument. The Relationship of the Organismic Design Argument to Darwinism. Anthropic Reasoning and Cosmic Design Arguments. A Prediction.

  15. The Design Argument

    This Element analyzes the various forms that design arguments for the existence of God can take, but the main focus is on two such arguments. The first concerns the complex adaptive features that organisms have. Creationists who advance this argument contend that evolution by natural selection cannot be the right explanation. The second design argument - the argument from fine-tuning - begins ...

  16. 18 The Design Argument and Natural Theology

    Philosophers call this 'the Teleological Argument', from the Greek word ' telos ', meaning 'goal' or 'end'. It is more commonly called 'the Design Argument'. The Design Argument touches on an enormous variety of issues in science, logic, metaphysics, epistemology, and value theory. Furthermore, there is not one version of it ...

  17. William Paley and The "Argument from Design"

    The "Argument from Design" is comprehended best when split into two phases. In Phase I of his argument, Paley asserts—via syllogism—that an object, such as a watch, must entail an intelligent designer. To do this he employs an inference to the best explanation, or a "best-fit" reason assigned to the seemingly inexplicable phenomenon.

  18. Examine the design argument for the existence of God

    In this essay, I will be focusing on the different variations of the design argument, as expressed by Paley, as well as the version expressed by Hume, through Cleanthes, and its modern version in the form of the Anthropic Principle. I will describe the ideas of design 'qua regularity' and design 'qua order'; explaining what each means.

  19. How to Write an Argumentative Essay

    Make a claim. Provide the grounds (evidence) for the claim. Explain the warrant (how the grounds support the claim) Discuss possible rebuttals to the claim, identifying the limits of the argument and showing that you have considered alternative perspectives. The Toulmin model is a common approach in academic essays.

  20. IELTS Writing Task 2: Format, Sample, Tips

    In the IELTS Writing Task 2, candidates are required to write an essay in response to a prompt or question. There are several common types of essays that may appear in Task 2: 1. Argumentative/Opinion Essays: These essays require candidates to express their opinion on a given topic and support it with reasons and examples. They often involve ...

  21. Elektrostal

    In 1938, it was granted town status. [citation needed]Administrative and municipal status. Within the framework of administrative divisions, it is incorporated as Elektrostal City Under Oblast Jurisdiction—an administrative unit with the status equal to that of the districts. As a municipal division, Elektrostal City Under Oblast Jurisdiction is incorporated as Elektrostal Urban Okrug.

  22. Krasnogorsk, Moscow Oblast

    The street of Krasnogorsk. Krasnogorsk ( Russian: Красногорск) is a city in Moscow Oblast in Russia. It is the administrative center of Krasnogorsky District of Moscow Oblast. As of 2010, Krasnogorsk has 196,896 people. In 2024, Islamic State killed over 130 people in a massacre at Crocus City Hall .

  23. Man charged with helping to fraudulently acquire Laguna Beach home

    A 34-year-old man pleaded not guilty Tuesday to conspiring with several others to fraudulently acquire a Laguna Beach home worth about $2 million. Vicente Anzu was charged in June 2023 to 23 ...

  24. Elektrostal

    Elektrostal. Elektrostal ( Russian: Электроста́ль) is a city in Moscow Oblast, Russia. It is 58 kilometers (36 mi) east of Moscow. As of 2010, 155,196 people lived there.

  25. Savvino-Storozhevsky Monastery and Museum

    The design of the building is influenced by the ancient Russian architectural style. Is prettier than the Tsar's chambers opposite, being red in colour with elaborately decorated window frames and entrance. At present the Tsaritsa's Chambers houses the Zvenigorod Historical, Architectural and Art Museum. Among its displays is an accurate ...