back
05 / 06
bird bird

"What evidence do we have for God's existence?"

INTRODUCTION: Good afternoon everyone, and a special welcome for our distinguished guest, Dr. William Lane Craig. You're in the front row! You'll hear from him in a minute. My name is Paul Ehlenbach, and I am the senior management liaison to the Bible Study Group ERG which is the host for this afternoon's event. The key letter in the acronym ERG is E for employee. ERGs are such a rich and valuable source of diversity in our laboratory community precisely because they are driven by employee affinities, backgrounds, interests, and beliefs, not by the Department of Energy or laboratory management. As Kim noted in her Kim's Connection a few weeks ago, diverse perspectives, new ideas, emergent approaches, a broad variety of skills and knowledge, and unique ways of thinking all contribute to the kind of innovation that is central to our mission.

The fundamental questions of physics and metaphysics, how and why our physical world and humankind even came to exist, were asked eons ago by the ancients who were also often the scientists of their day, and they remain the subject of passionate discussion and argument. I personally welcome this break from our everyday work to hear Dr. Craig's presentation having relevance to those fundamental questions.

KEVIN: Hello, and welcome. My name is Kevin, and I am one of the co-chairs of the Bible Study Group, or BSG for short. I'd like to thank you all for coming today to listen to Dr. William Lane Craig. For those of you unfamiliar with Dr. Craig, he is an author, professor, and philosopher. He's devoted much of his career to understanding how to reconcile our knowledge of the universe with the knowledge of the Bible. He has also put his ideas to the test by publicly debating more than 40 scholars in academics from other academic institutions. But the purpose of today's lecture is to start a conversation regarding God's existence. Answering this question is a good starting point to explaining the Bible. As an analogy, consider how difficult it would be to explain what a molecule is while assuming that atoms do not exist. Now although there can be wisdom in remaining silent on God's existence, there is also wisdom in talking about it objectively. This kind of conversation helps us consider life's most important questions and provides a framework for making sense of our own experience, their existence. For sure 45 minutes is a short time to explain the evidence for God's existence so if you're left hungry for more then we will still consider that we have achieved our goal in starting the conversation. I'm pleased to welcome Dr. Craig to the podium. Thank you.

DR. CRAIG: Thank you very much! I am delighted to have the privilege of speaking at Lawrence Livermore Laboratory, and I'm grateful for your coming this afternoon.

In 1991, at a scientific conference on “The History and Philosophy of Thermodynamics,” the prominent British physicist P. T. Landsberg suddenly began to explore the theological implications of the scientific theory he was discussing. He observed,

To talk about the implications of science for theology at a scientific meeting seems to break a taboo. But those who think so are out of date. During the last 15 years, this taboo has been removed, and in talking about the interaction of science and theology, I am actually moving with a tide which is threatening to wash us away in a flood of publications.[1]

Over the last fifty years or so, a flourishing dialogue between science and theology has been going on in North America and Europe. Folks who think that science and theology are hostile or mutually irrelevant need to realize that the cat is already out of the bag; and I daresay there’s little prospect of stuffing it back in. Science and theology have discovered that they have important mutual interests and important contributions to make to each other, and those who don’t like this can choose not to participate in the dialogue, but that’s not going to shut down the dialogue or show it to be meaningless.

I can think of at least six ways in which science and theology are relevant to each other, starting with the most general and becoming more particular.

1. Theology furnishes a conceptual framework in which science can flourish.

2. Science can both falsify and verify certain theological claims.

3. Science encounters metaphysical problems which theology can help to solve.

4. Theology can augment the explanatory power of science.

5. Theology can help to adjudicate between scientific theories.

6. Science can establish a premise in an argument for a conclusion having theological significance.

Today I want to focus only on the last point. I want to share with you three arguments whose premises enjoy considerable scientific support and which imply conclusions having theological significance. In sharing these arguments, I am acutely aware of how superficial my presentation must be in light of the time constraints. Entire books have been written on each of these arguments, and so I can present only the tip of the iceberg. Beneath the surface lurks a massive amount of material waiting to be explored by those who are interested. If you are one of these, then I highly recommend to you this book The Blackwell Companion to Natural Theology, which includes lengthy discussions of various theistic arguments.

Kalām Cosmological Argument

Our first argument is a version of the cosmological argument known as the kalām cosmological argument. This argument has a long history in Jewish, Muslim, and Christian thought reaching all the way back to the early centuries after Christ. Let’s let one of its greatest medieval Muslim proponents speak for himself. Al-Ghazali was a twelfth century theologian from Persia. He frames the argument very simply: “Every being which begins has a cause for its beginning; now the world is a being which begins; therefore, it possesses a cause for its beginning.”[2]

Ghazali’s deductive formulation of the argument involves three simple steps:

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

2. The universe began to exist.

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

One can then do a conceptual analysis of what it is to be a cause of the universe, as a result of which a number of theologically striking properties of the First Cause can be recovered.

Ghazālī offered philosophical arguments in support of his premises, arguments which still merit consideration today. But in one of the most astonishing developments of modern astronomy and astrophysics, which Ghazālī could never have anticipated, we now have pretty strong scientific evidence in support of the crucial second premise that the universe began to exist.

Consider, for example, the evidence from the expansion of the universe. The standard (Friedman-LeMaître Robertson-Walker) Big Bang cosmogonic model implies that the universe is not infinite in the past but had an absolute beginning a finite time ago. As you trace the expansion of the universe back in time, the galaxies will begin to coalesce. Eventually the distance between any two points in space becomes zero. So at that point you have reached the boundary of space and time. Space and time cannot be extended any further back than that. It is literally the beginning of the universe.

In the standard model there’s simply nothing prior to the initial boundary point. Here it is very important that we not be misled by words. When astrophysicists say, “There is nothing prior to the initial boundary,” they do not mean that there is something prior to it, and that is a state of nothingness. That would be to treat nothing as though it were something! The word “nothing” is just a term of universal negation, meaning “not anything.” So when scientists say there is nothing prior to the Big Bang or nothing prior to the boundary point, what they mean is there was not anything prior to the boundary point.

Incredibly, the standard Big Bang model thus predicts an absolute beginning of the universe. If this model is correct, then we have amazing scientific confirmation of the second premise of the kalām cosmological argument – the universe began to exist.

The question then is: is the standard model correct? Or, I think, more importantly, is it correct in predicting a beginning of the universe? Although advances in astrophysical cosmology have forced various revisions in the standard model, nothing has called into question its fundamental prediction of the finitude of the past and the beginning of the universe. Indeed, 20th century cosmogony has seen a parade of failed theories trying to avert the absolute beginning predicted by the standard model.[3] Meanwhile, a series of remarkable singularity theorems has increasingly tightened the loop around empirically tenable cosmogonic models by showing that under more and more generalized conditions, a beginning is inevitable. In 2003 Arvind Borde, Alan Guth, and Alexander Vilenkin were able to show that any universe which is, on average, in a state of cosmic expansion throughout its history cannot be infinite in the past but must have a beginning.[4] In 2012 Vilenkin showed that cosmogonic models which do not fall under this condition fail on other grounds to avert the beginning of the universe. Vilenkin concluded, “There are no models at this time that provide a satisfactory model for a universe without a beginning.”[5] “All the evidence we have says that the universe had a beginning.”[6]

The Borde-Guth-Vilenkin theorem proves that under a single very general condition classical space-time must shrink down to a boundary at some point in the past. Now either there was something on the other side of that boundary or not. If not, then that boundary simply was the beginning of the universe. If there was something on the other side of that boundary, it will be the quantum gravity regime described by the yet-to-be-discovered quantum theory of gravity. In that case, Vilenkin says, it will be the beginning of the universe. So the Borde-Guth-Vilenkin theorem shows either that the universe began to exist at this past boundary or that, if there was a quantum gravity regime, then that regime is the beginning of the universe.

Vilenkin's confidence in this fact, even in the absence of a quantum theory of gravity, is based upon the fact that such a quantum regime is metastable. That is to say, it cannot endure for very long. Certainly, it would be impossible for such a metastable condition to endure for infinite time doing nothing and then suddenly begin to expand about 13.8 billion years ago. So even though we do not yet have a description of this earliest phase of the universe, we can be confident that if such a quantum regime does exist, then it was the beginning of the universe.

In the online scientific magazine Inference in 2015, in an article entitled “Did the Universe Have a Beginning?”[7] Vilenkin interacts explicitly with the kalām cosmological argument. Vilenkin is himself an agnostic. So he is going to reject one of the two premises of the kalām cosmological argument. But he doesn't reject the second premise, that the universe began to exist. Quite the contrary, he affirms it. In the article he says:

We have no viable models of an eternal universe. The BGV [Borde-Guth-Vilenkin] theorem gives us reason to believe that such models simply cannot be constructed.

How, then, does Vilenkin, as an agnostic, respond to the kalām argument? He chooses to reject the first premise – that everything that begins to exist has a cause. He maintains that the universe just popped into existence uncaused out of nothing. What justification does he have for such a remarkable hypothesis? Well, he says, in a closed universe (that is, one that is finite in volume), the positive energy and the negative energy in such a universe balance each other out so that the net energy is zero. Therefore if the universe pops into being uncaused out of nothing, the conservation laws of matter and energy are not violated. Therefore the universe can simply come into being [uncaused] from nothing.

I have to say that I find this argument difficult to take seriously. Vilenkin assumes that if something doesn't violate the conservation laws, then it is metaphysically possible. But there is no reason to adopt such an assumption. Philosopher of science Michael Ruse points out, “The man who says that it is morally acceptable to rape little children is just as mistaken as the man who says 2+2 = 5.” Certain ethical principles are characterized by the same logical necessity as arithmetic truths. Yet it’s being good to rape little children would not violate any conservation laws. Similarly, just because coming into being uncaused out of nothing wouldn't violate the conservation laws doesn't mean that it is metaphysically possible for something to come into being uncaused from nothing.

Worse, when you think about it, the situation that Vilenkin imagines, that the universe can come into being uncaused if its positive energy is exactly balanced by its negative energy, just seems completely wrong-headed. It is like saying that if your financial assets and your financial debits exactly balance each other out, then your net worth is zero and therefore there is no cause of your financial condition. Clearly, that would be a mistake. Christopher Isham, who is Great Britain's leading quantum cosmologist, points out that even if the positive and negative energy balance each other out, so that the net energy is zero, there still needs to be, in his words, “ontic seeding” to create the positive and negative energy in the first place! So even if you had the exact balance of positive and negative energy, that wouldn't eliminate the need for a cause of the origin of the universe.

Of course, scientific results are always provisional. Science doesn't deal in certainties. It deals in probabilities. We can fully expect that new theories will be proposed, trying to avoid the universe’s beginning. These proposals are to be welcomed and tested. Nevertheless, I think it’s pretty clear which way the evidence points. Today the proponent of the kalām cosmological argument stands comfortably within the scientific mainstream in holding that the universe began to exist.

The Unreasonable Effectiveness of Mathematics

Our second argument concerns the problem of mathematics’ applicability to physical phenomena. Applicability has to do with what mathematician and physicist Eugene Wigner famously called “the unreasonable effectiveness of mathematics in the natural sciences.”[8] Mathematics is the language of nature. That is to say, the laws of nature may be expressed as mathematical equations which describe the phenomena to an astonishing degree of accuracy.

In his seminal paper Wigner states his principal point as follows:

Mathematical concepts turn up in “entirely unexpected connections” in physics and often permit “an unexpectedly close and accurate description” of the phenomena in these connections.[9]

In unfolding his point, Wigner inquires, first, as to the role of mathematics in physics and, second, as to why mathematics’ success in that role appears “so baffling.” With respect to mathematics’ role in physics, Wigner notes that while mathematics is useful in physics for evaluating the consequences of the laws of nature, a role which he associates with applied mathematics, it also plays a more “important” and “sovereign” role in physics, namely, to enable the formulation of the laws of nature themselves in the language of mathematics in order to be an apt object for the use of applied mathematics.

Wigner notes that “the mathematical formulation of the physicist’s often crude experience leads in an uncanny number of cases to an amazingly accurate description of a large class of phenomena.”[10] He provides three examples in support: Newton’s second law of motion, elementary quantum mechanics, and quantum electrodynamics. Wigner takes his examples—“which could be multiplied almost indefinitely”—to illustrate the “appropriateness” and “almost fantastic accuracy” of the mathematical formulation of the laws of nature.

With respect to why mathematics’ role in physics appears “so baffling,” Wigner notes that his three examples represent a hierarchy of increasing independence from empirical experience in favor of reliance on mathematical concepts which are chosen for aesthetic reasons rather than empirical applicability:

whereas it is unquestionably true that the concepts of elementary mathematics and particularly elementary geometry were formulated to describe entities which are directly suggested by the actual world, the same does not seem to be true of the more advanced concepts, in particular the concepts which play such an important role in physics. . . . Most more advanced mathematical concepts, such as complex numbers, algebras, linear operators, Borel sets - and this list could be continued almost indefinitely - were so devised that they are apt subjects on which the mathematician can demonstrate his ingenuity and sense of formal beauty.[11]

Wigner’s point is well-taken. As philosopher of mathematics Penelope Maddy emphasizes, mathematicians employ what she calls “maximizing principles of a sort quite unlike anything that turns up in the practice of natural science: crudely, the scientist posits only those entities without which she cannot account for our observations, while the set theorist posits as many entities as she can, short of inconsistency.”[12] Maddy identifies quite a few of these “rules of thumb” followed by mathematicians, such as maximize, richness, diversity, one step back from disaster, etc.[13] Similarly Wigner observes, “The great mathematician fully, almost ruthlessly, exploits the domain of permissible reasoning and skirts the impermissible.”[14]

Wigner’s “principal point” is that mathematicians are free to define new concepts with a view, not of applicability or scientific utility, but of “permitting ingenious logical operations which appeal to our aesthetic sense both as operations and also in their results of great generality and simplicity.”[15] Wigner is not talking about aesthetics in the artistic sense, but in the sense of mathematical beauty, what Maddy calls mathematical depth. Mathematics is an a priori discipline which is independent of the physical world. Moreover, when we reflect that mathematical objects, even if they exist, are causally effete, it is surprising that such objects should be significantly effective in physics. Wigner muses, “It is difficult to avoid the impression that a miracle confronts us here.”[16]

Accordingly, the following seems to be a faithful formulation of Wigner’s argument:

1. Mathematical concepts arise from the aesthetic impulse in humans and have no causal connection to the physical world.

2. It would be surprising to find that what arises from the aesthetic impulse in humans and has no causal connection to the physical world should be significantly effective in physics.

3. Therefore, it would be surprising to find that mathematical concepts should be significantly effective in physics.

4. The laws of nature can be formulated as mathematical descriptions (concepts) which are often significantly effective in physics.

5. Therefore, it is surprising that the laws of nature can be formulated as mathematical descriptions that are often significantly effective in physics.

Given that something surprising merits prima facie an explanation, we wonder as to the explanation of the fact that the laws of nature can be formulated as mathematical descriptions that are often significantly effective in physics.

Wigner, despite his characterization of the applicability of mathematics to the physical world as a miracle, never actually considered in his essay whether theism might not furnish a good explanation of mathematics’ applicability. He considered at most naturalistic explanations of it and, finding none to be satisfactory, therefore concluded “that the enormous usefulness of mathematics in the natural sciences is something bordering on the mysterious and that there is no rational explanation for it.”[17]

But suppose we take the theistic hypothesis seriously. Theists will have a considerably easier time, I think, explaining the applicability of mathematics than will naturalists. Theists hold that there is a personal, transcendent being (a.k.a. God) who is the Creator and Designer of the universe. Naturalists hold that all that exists concretely is space-time and its physical contents. Now whether one is a realist or an anti-realist about mathematical objects, it appears that the theist enjoys a considerable advantage over the naturalist in explaining the uncanny success of mathematics.

Consider, first, realism’s take on the applicability of mathematics to the world. For the non-theistic realist, the fact that physical reality behaves in accord with the dictates of acausal mathematical entities existing beyond space and time is, in the words of philosopher of mathematics Mary Leng, “a happy coincidence.”[18] For consider: If, per impossibile, all the abstract objects in the mathematical realm were to disappear overnight, there would be no effect on the physical world. This is simply to underscore the fact that abstract objects are causally inert. The idea that realism somehow accounts for the applicability of mathematics “is actually very counterintuitive,” muses philosopher of mathematics Mark Balaguer. He says, “The idea here is that in order to believe that the physical world has the nature that empirical science assigns to it, I have to believe that there are causally inert mathematical objects, existing outside of spacetime,” an idea which is inherently implausible.[19]

By contrast, the theistic realist can argue that God has fashioned the world on the structure of the mathematical objects He has chosen. This is essentially the view that Plato defended in his dialogue Timaeus. Plato draws a fundamental distinction between the realm of static being and the realm of temporal becoming. The realm of becoming is comprised primarily of physical objects, while the static realm of being is comprised of logical and mathematical objects. God looks to the realm of mathematical objects and models the world on it. The world has its mathematical structure as a result. Thus, the realist who is a theist has a considerable advantage over the [naturalistic realist] in explaining why mathematics is so effective is describing the physical world.

Now consider anti-realism, first of a non-theistic sort. Leng says that on anti-realism relations which are said to obtain among pretended mathematical objects just mirror the relations obtaining among things in the world, so that there is no happy coincidence. Philosopher of physics Tim Maudlin muses, “The deep question of why a given mathematical object should be an effective tool for representing physical structure admits of at least one clear answer: because the physical world literally has the mathematical structure; the physical world is, in a certain sense, a mathematical object.”[20]

Well and good, but what remains wanting on naturalistic anti-realism is an explanation why the physical world should exhibit so elegant and stunning a mathematical structure in the first place. Not only so, but by choosing examples like the infinite-dimensional Hilbert space and complex numbers, Wigner implicitly precluded the explanation that physical reality is isomorphous to such mathematical structures, since these cannot be physically realized. Mark Steiner provides numerous examples of the applicability of mathematical concepts that cannot be physically instantiated.[21] Some of his examples are the same ones to which Wigner already appealed, such as the descriptive applicability of the Hilbert space formalism to quantum mechanics, which Steiner calls “physically unintelligible.” So even if the physical universe had to have some mathematical structure, that fails to address the question raised by Wigner.

By contrast, the theistic anti-realist has a ready explanation of the applicability of mathematics to the physical world: God has created it according to a certain mental model which He had in mind. This was the view of the first century Jewish philosopher Philo of Alexandria, who maintained in his treatise On the Creation of the World that God created the physical world on the mental model in His mind. For a Jewish monotheist like Philo, the realm of Ideas does not exist, as Plato thought, independently of God but as the contents of His mind. Philo referred to the mind of God as God’s Logos (Word). The sensible world is made on the model of the conceptual or intelligible world that pre-exists in the Logos. Just as the ideal architectural plan of a city exists only in the mind of the architect, so the world of ideas exists solely in the mind of God.

Thus, the theist—whether he be a realist or an anti-realist about mathematical objects—has the explanatory resources to account for the otherwise unreasonable effectiveness of mathematics in physical science—resources which the naturalist lacks.

We may thus extend Wigner’s argument:

6. Therefore, the fact that the laws of nature can be formulated as mathematical descriptions that are often significantly effective in physics merits explanation.

7. Theism provides a better explanation of the fact that the laws of nature can be formulated as mathematical descriptions that are often significantly effective in physics than does atheism.

8. Therefore, the fact that the laws of nature can be formulated as mathematical descriptions that are often significantly effective in physics provides evidence for theism.

It would be helpful to have a simpler formulation of this argument. It seems to me that we would not be misleading to epitomize our argument as follows:

1. If God does not exist, the applicability of mathematics is just a “happy coincidence.”

2. The applicability of mathematics is not just a “happy coincidence.”

3. Therefore God exists.

To quote Paul Dirac, “God is a mathematician.”[22]

Teleological Argument

Our third argument is the teleological argument. Widely thought to have been demolished by Hume and Darwin, the teleological argument for God’s existence has come roaring back into prominence during the latter half of the twentieth century. Scientists have been stunned by the discovery that the existence of intelligent, interactive life depends upon a complex and delicate balance of fundamental constants and quantities, like the gravitational constant and the amount of entropy in the early universe, which are fine-tuned to a degree that is literally incomprehensible.

What is meant by “fine-tuning”? By “fine-tuning for life” one means that small deviations from the actual values of the constants and quantities in question would render the universe life-prohibiting or, alternatively, that the range of life-permitting values is exquisitely narrow in comparison with the range of assumable values. Since fine-tuning compares only worlds governed by the same laws of nature as ours, but with different values of the constants and quantities, scientists can predict fairly reliably what would happen if those values were to be changed. The results would be catastrophic. In the absence of fine-tuning there would not even be chemistry, there would not even be matter, much less stars and planets where life might evolve.

Will the fine-tuning “go away” with the advance of physics? The multiplicity and variety of cases of fine-tuning make this unlikely. Indeed, the pattern to date has been that fine-tuning, like the stubborn bump in the carpet, is suppressed at one point only to re-appear elsewhere.

Now there are three hypotheses debated in the literature for explaining the presence of this remarkable fine-tuning: physical necessity, chance, or design. The question then is: Which of these alternatives is the most plausible?

Accordingly, we can offer this very simple formulation of the argument:

1. The fine-tuning of the universe is due to either physical necessity, chance, or design.

2. It is not due to physical necessity or chance.

3. Therefore, it is due to design.

Given the fact of fine-tuning, (1) simply lists the alternative hypotheses which are live options in the current literature for explaining the fine-tuning evidence. If someone comes up with another alternative that is a live option, then we shall add it to the list.[23]

Turn, then, to premiss (2). Here we must ask whether the hypotheses of physical necessity or of chance are as plausible as the hypothesis of design.

Consider the first alternative, physical necessity. This alternative seems extraordinarily implausible because the constants and quantities are independent of the laws of nature. The laws of nature are consistent with a wide range of values for these constants and quantities. For example, the most promising candidate for a Theory of Everything to date, super-string theory or M-Theory, allows a “cosmic landscape” of around 10500 different universes governed by the present laws of nature, so that it does nothing to render the observed values of the constants and quantities physically necessary.

So what about the second alternative, that the fine-tuning is due to chance? The problem with this alternative is that the odds against the universe’s being life-permitting are so incomprehensibly great that they cannot be reasonably faced. In order to rescue the alternative of chance, its proponents have therefore been forced to adopt the hypothesis that there exists a sort of World Ensemble or multiverse of randomly ordered universes of which our universe is but a part. Now comes the key move: since observers can exist only in finely tuned worlds, of course we observe our world to be fine-tuned!

So this explanation of fine tuning relies on (i) the existence of a specific type of World Ensemble and (ii) an observer self-selection effect. Now this explanation, wholly apart from objections to (i), faces a very formidable objection to (ii), namely, the Boltzmann Brain problem. In order to be observable the entire universe need not be fine-tuned for our existence. Indeed, the smallest observable universe consists of a single brain which fluctuates into existence with illusory perceptions of an external world, what is called, after the Austrian physicist Ludwig Boltzmann, a Boltzmann Brain. It is vastly more probable that a random fluctuation of mass-energy would yield a universe dominated by Boltzmann Brain observers than one dominated by ordinary observers like ourselves.

In other words, the observer self-selection effect is explanatorily vacuous. Simon Friederich, writing in the Stanford Encyclopedia of Philosophy, reports, “The broad consensus in the literature on multiverse cosmology is that, in order for a multiverse scenario to qualify as empirically confirmed, it must entail that those conditions that we find in our own universe are typical among those found by observers across the multiverse.”[24] Appeal to an observer self-selection effect accomplishes nothing because there’s no reason whatsoever to think that most observable worlds or the most probable observable worlds are worlds in which that kind of observer exists. Indeed, the opposite appears to be true: most observable worlds will be Boltzmann Brain worlds. Thus, the hypothesis of an infinite World Ensemble leads to a radical scepticism: we cannot trust our perceptions that there is an external world.

With the failure of the Many Worlds Hypothesis the last ring of defense for the alternative of chance falls. Thus, we have good reason for thinking that neither physical necessity nor chance provides a plausible explanation of the fine-tuning of the universe. It follows that the fine-tuning is therefore due to design, unless the design hypothesis can be shown to be even more implausible than its competitors. But that is a talk for another day.

Conclusion

In conclusion, today I have given three examples of science’s providing evidence for a premise of an argument leading to a conclusion of theological significance. And that is just one reason why the dialogue between science and theology is flourishing today.

MODERATOR: We’d like to thank Dr. Craig for that riveting presentation. Leading up to this event, Livermore employees were able to submit questions on this very topic. Dr. Craig, I'm not sure if you've heard, but we've been known as the smartest square mile on Earth, and the questions we've received have certainly met that mark.

DR. CRAIG: I hope I haven't diminished your degree then by coming! [laughter]

MODERATOR: In the interest of time, we selected a few of these questions to ask Dr. Craig for his answers. So, Dr. Craig, first question. Miracles such as the resurrection of Jesus Christ are difficult to reconcile with a modern scientific understanding of the world. Why should we believe these miracles occurred, rather than concluding that they are most likely legends, given their scientific implausibility?

DR. CRAIG: I think it's important to understand that there's absolutely no incompatibility between what modern science says about the operations of the natural world and the occurrence of a miracle caused by God. What science and scientific laws tell us is that what happens if no supernatural factors are interfering. The laws of nature have in them implicit ceteris paribus conditions – that there are no supernatural factors interfering. But if there is a supernatural agent at work then science cannot say that that is impossible, and the law of nature would not be violated by a miracle because the miracle assumes that there are no supernatural agents at work. If you give up that philosophical prejudice against the possibility of miracles, and if you have arguments like this that there is a transcendent Creator and Designer of the universe then the plausibility of events like Jesus’ resurrection go way up because you've already got in place the causal conditions that could produce such a resurrection. As Peter Slezak, an Australian philosopher that I debated, once put it: for the Creator and Designer of the universe, the odd resurrection would be child's play. So when you assess whether or not that event occurred, what you need to do is to compare the historical hypotheses for what happened following Jesus’ crucifixion including the ones you mentioned: legend or conspiracy and so forth. You weigh these hypotheses by the typical criteria used to assess historical hypotheses: explanatory scope, explanatory power, degree of ad-hocness, plausibility, and so forth. I've argued that what I call “the resurrection hypothesis” (namely, that God raised Jesus from the dead) passes those tests better than any naturalistic counter-explanation that I know of. In particular, the hypothesis of legend is actually against the consensus of New Testament scholarship. That was a view that was popular back in the 30s and 40s but today the wide majority of New Testament historians agree on the fundamental facts underlying the inference to the resurrection of Jesus; namely, the discovery of Jesus’ empty tomb on the first morning of the week by a group of his women followers, secondly the post-mortem appearances of Jesus to various individuals and groups, and thirdly the very origin of the disciples’ belief that God raised Jesus from the dead. Those do not represent the opinions of conservative scholars. Those represent the broad consensus of New Testament scholarship on the historical Jesus whether secular or Christian, liberal or conservative. That would illustrate the fact that we've got to come up with some sort of historical hypothesis that will credibly account for that data, and I don't know of any historical hypothesis that would do so better than the one that the disciples themselves gave – that God raised Jesus from the dead. I apologize for that long-winded answer but it was a complex question.

MODERATOR: Our second question. Why does God allow such great human suffering? Either God does not care about human suffering, in which case God is not all-caring, or God is powerless to stop human suffering, in which case God is not all-powerful. Why would I give worship to a God who was either incapable or indifferent to the suffering of his creations?

DR. CRAIG: You don't ask small questions, do you?! This is one of the most profound questions facing the theist. Let me try to give a succinct overview here. I'm painfully aware this is going to be superficial. I think it's helpful to distinguish between what I call the intellectual problem of suffering and the emotional problem of suffering. The intellectual problem concerns how to reconcile the existence of an omnipotent, morally perfect being with the horrible suffering and evil in the world. The emotional problem concerns how to dissolve people’s dislike of a God who would permit them or others to suffer so terribly. I am convinced in talking with folks that for the vast majority of people this is not really an intellectual problem. They've never really thought about it very deeply. Rather, it's an emotional problem. But as a philosophical problem that I am called upon to deal with as a professional philosopher, it places a burden of proof upon the atheist which is so heavy that no atheist has ever been able to successfully shoulder that burden. The atheist must show that it is either impossible or highly improbable that God could have morally sufficient reasons for permitting the evil and suffering in the world, and there's simply no way that he could show that. That involves probability judgments that are way beyond our capacity to make. In my published work (for example, my book Philosophical Foundations for a Christian Worldview), I give several illustrations of this point. One, in fact, from contemporary science – chaos theory indicates the fact that from observing a particular historical event (say, an instance of evil), it can have a reverberating effect through history that make it utterly impossible to predict its outcome and to say, “God probably doesn't have a morally sufficient reason for allowing that.” I think that intellectually the problem of evil and suffering is a colossal failure. Having said that, that doesn't do anything to remove people's emotional dislike of a God who would allow people to suffer terribly. But here, as a Christian theologian, I think the answer to the emotional problem is to be found in the person of Christ, and in particular in the cross of Christ. On the cross of Christ, we see how God himself in the person of his Son voluntarily gives his life to redeem us from our own evil by bearing the punishment that we deserve thereby setting us free from the problem of our own evil and guilt before God. When you comprehend his incomprehensible love for you and the extent to which he was willing to bear innocent suffering of an incomprehensible nature for you, for your sake, that can give us the emotional strength to bear the cross that he calls upon us to bear in this finite life until we go home to be with him forever in a life of unspeakable joy. So I think that we can find comfort and succor in the cross of Christ to enable us to bear the suffering that is our lot in this finite existence.

MODERATOR: We have time for one more question. It's a long one. Scientists by nature have a healthy amount of skepticism towards even their own theories and readily discard theories for which contradictory evidence is provided. Furthermore, in science just because there is evidence for something does not make it immune to doubt, skepticism, and scrutiny. Assuming that there is actually evidence for God's existence, it is still only evidence. Does that mean that there still needs to be a doubt and skepticism within religion? And if God's existence can be totally proven beyond a shadow of a doubt then what even is faith or belief? Wouldn't there be no need for faith?

DR. CRAIG: There's two questions going there. In answer to the first question: certainly there is room for doubt and uncertainty in religion and particularly in Christianity. I love the story in the Gospels where the father comes to Jesus asking for healing for his son and Jesus says, “If you have faith he can be healed.” And the man says, “Lord, I believe! Help thou my unbelief.” And Jesus accepts that kind of partial, tentative, uncertain faith as perfectly legitimate. So we shouldn't think that having religious faith is a matter of certainty or not having doubts or questions. Not at all. Now, as for the incompatibility of faith with certainty, for those who are blessed with certainty, I don't think that obviates the need for faith at all. Faith in the New Testament sense of the term means trusting in what you know to be true. It's trusting in what you have good reason [to believe] to be true. So you may have very good reasons, for example, for believing in your doctor. But that's not enough for you to actually then trust him. The act of faith is putting your trust in the hands of that person for whom you have good reason to believe. So certainty doesn't eliminate the need for faith because faith involves trusting in someone or something that you have good reason to believe is true.

MODERATOR: Thank you, Dr. Craig, for answering those questions. Unfortunately, we're running towards the end of our time here. If you submitted a question that wasn't answered or if you have had more questions we encourage you to check out more of Dr. Craig's work. But at this time I'd like to invite Stacy to come and close us out.

DR. CRAIG: Thank you very much. Thank you.

STACY KANE: I'm Stacy Kane. I'm also a co-chair with the BSG, and we'd really like to thank Dr. Craig for delivering this informative lecture. We'd also like to thank all those who made this possible, of which there's too many to name here. But we especially like to thank all of you who came to attend, and we appreciate your time and interest. If you'd like to learn more, feel free to reach out to us in the BSG. We're a resource for everyone regardless of your faith. Before you leave we have special gifts outside. We have a book table. We'd like you to stop by on your way out and feel free to pick up a book. We just ask that you would allow others to also have access to these books, and please read what you take. We just hope that the rest of your day would be blessed, and we just want to thank you again for coming. Thank you, Dr. Craig, and just have a safe and blessed day. Thanks.

 

[1] P. T. Landsberg, “From Entropy to God,” in Thermodynamics: History and Philosophy, ed. K. Martinas, L. Ropolyi, and P. T. Landsberg (Singapore: World Scientific, 1991), p. 380.

[2] Al-Ghazālī, Kitab al-Iqtisad fi’l-Iqtiqad (Ankara: University of Ankara Press, 1962), pp. 15-16.

[3] William Lane Craig and James Sinclair, “The Kalam Cosmological Argument,” in The Blackwell Companion to Natural Theology, ed. Wm. L. Craig and J. P. Moreland (Oxford: Wiley-Blackwell, 2009), pp. 101-201; idem, “On Non-Singular Spacetimes and the Beginning of the Universe,” in Scientific Approaches to the Philosophy of Religion, ed. Yujin Nagasawa, Palgrave Frontiers in Philosophy of Religion (London: Macmillan, 2012), pp. 95-142.

[4] A. Borde, A. Guth, A. Vilenkin, “Inflationary Spacetimes Are Incomplete in Past Directions,” Physical Review Letters 90 (2003): 151301, http://arxiv.org/abs/gr-qc/0110012.

[5] Mithani and Vilenkin, “Did the universe have a beginning?” p. 1; cf. p. 5. For application to the Carroll-Chen model, see http://www.youtube.com/watch?v=NXCQelhKJ7A, where Vilenkin concludes, “there are no models at this time that provide a satisfactory model for a universe without a beginning.” See further Alexander Vilenkin, “Arrows of time and the beginning of the universe,” arXiv:1305.3836v2 [hep-th] 29 May 2013.

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

[8] Eugene Wigner, “The Unreasonable Effectiveness of Mathematics in the Natural Sciences,” in Communications in Pure and Applied Mathematics 13/1 (New York: John Wiley & Sons, 1960).

[9] Wigner, “Unreasonable Effectiveness of Mathematics,” p. 2. Wigner re-phrases this first point as “mathematics plays an unreasonably important role in physics.”

[10] Ibid., p. 8.

[11] Ibid., pp. 2-3.

[12] Ibid., p. 131.

[13] Penelope Maddy, “Believing the Axioms I,” Journal of Symbolic Logic 53/2 (1988): pp. 481-511.

[14] Wigner, “Unreasonable Effectiveness of Mathematics,” p. 3.

[15] Ibid.

[16] Ibid., p. 7.

[17] Ibid., p. 2.

[18] Mary Leng, Mathematics and Reality (Oxford: Oxford University Press, 2010), p. 239.

[19] Mark Balaguer, Platonism and Anti-Platonism in Mathematics, p. 136.

[20] “On the Foundations of Physics,” July 5, 2013, http://www.3ammagazine.com/3am/philosophy-of-physics/.

[21] The turning point from examples of descriptive applicability which Steiner deems not mysterious because they can be explained in terms of physical properties of nature to examples of descriptive applicability which do seem mysterious because they have no physical basis, occurs at pp. 35-36 of Steiner’s book. For discussion of the following examples, see Steiner, Applicability of Mathematics, pp. 36-40, 95-97, 102.

[22] P. A. M. Dirac, “The Evolution of the Physicist’s Picture of Nature,” Scientific American 208/5 [May, 1963]: p. 53.

[23] These three alternatives encompass Waller’s five options for explaining fine-tuning, for physical necessity would encompass metaphysical necessity, since metaphysical necessity entails physical necessity, and brute factuality is encompassed by chance, since if something is a brute fact then on Waller’s view it obtains by chance. The real problem, overlooked by Waller, with the brute fact option is that, as a chance hypothesis, the odds that the one universe inexplicably obtaining should be fine-tuned for life is infinitesimal and therefore not to be accepted when there are better alternatives (Waller, Cosmological Fine-Tuning Arguments, pp. 217-23). For a similar taxonomy see Andrew Loke, The Teleological and Kalam Cosmological Arguments Revisited, Palgrave Frontiers in Philosophy of Religion (Cham, Switz.: Springer, 2022), pp. 156-57.

[24] Stanford Encyclopedia of Philosophy, s.v. “Fine-Tuning,” by Friederich. Remarkably, Waller, despite his thoroughness, has no discussion of the problem of typicality. He is content to argue that in World Ensembles having each an actually infinite number of randomly ordered members, it is virtually certain that there would be life-permitting worlds. Thus, for any randomly chosen multiverse, the probability of getting a fine-tuned universe is nearly certain. He takes no cognizance of Boltzmann Brains or the measure problem (Waller, Cosmological Fine-Tuning Arguments, pp. 190–91, 223-34). Indeed, his Random Multiverse Model seems itself to founder on the measure problem, since he speaks of proportions of the actually infinite collection (pp. 230, 256). See further Holder, God, the Multiverse, and Everything, chap. 9. Holder argues that the likelihood of the naturalistic Many Worlds Hypothesis is reduced because (i) the probability that any universe in an infinite World Ensemble is fine-tuned for life is zero, so that the proportion of life-permitting universes remains undefined; (ii) of the uncountable infinity of possible universes above a certain size, only a countable subset can be realized in a single spacetime, so that the existence of even one life-permitting universe is not guaranteed; (iii) the hypothesis is inconsistent with the amount of order in this universe, which is vastly greater than could be expected on any chance hypothesis.