Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to main content
There has been a recent explosion of interest in the epistemology of disagreement. Much of the recent literature is concerned with a particular range of puzzle cases (discussed in the first section of my paper). Almost all of the papers... more
There has been a recent explosion of interest in the epistemology of disagreement. Much of the recent literature is concerned with a particular range of puzzle cases (discussed in the first section of my paper). Almost all of the papers that contribute to that recent literature make mention of questions about religious disagreement in ways that suggest that there are interesting connections between those puzzle cases and real life cases of religious disagreement. One important aim of my paper is to cast doubt on that suggestion. More generally, the aim of my paper is to give a reasonably full account of the recent literature on the epistemology of disagreement, and then to give a serious discussion of some of the epistemological issues that are raised by real world religious disagreements.
The plan of the paper is as follows: Following time honoured tradition, I begin by drawing some distinctions, and adverting to some disputes upon which I will take up a stance without proper discussion. After briefly considering the... more
The plan of the paper is as follows:

Following time honoured tradition, I begin by drawing some distinctions, and adverting to some disputes upon which I will take up a stance without proper discussion. After briefly considering the distinction between norms that are distinctive of assertion and norms that are shared between assertion and other speech acts, I spend some time thinking about different ways in which norms can be classified (and, in particular, I consider ways in which the taxonomy of Wright (1992) can be improved upon). Then, after some brief consideration of the distinction between norms and secondary proprieties, I conclude with a rather short discussion of exactly what it takes for a speech act to be an assertion (as opposed to some other kind of speech act that is characteristically performed using declarative sentences).

In the second section of the paper, I turn my attention to the claim that it is a norm of assertion that one ought not to assert that which one does not know. I argue that the arguments that Williamson (along with others) has offered in defence of this claim fail, and that there are good reasons for rejecting the claim that it is a norm of assertion that one ought not to assert that which one does not know.

In the third section of the paper, I consider the claim that it is a norm of assertion that one ought not to assert that which is not true. I argue that the standard argument in favour of this claim fails, and that there is good reason to deny that it is a norm of assertion that one ought not to assert that which is not true. Moreover, I also argue that there is similar good reason to deny that it is a norm of assertion that one ought not to assert that for which one has insufficient warrant. However, I do not go on to reject all of Grice’s norms of quality, for I also contend that it is the central norm of assertion that one ought not to assert that which one does not believe.

In the fourth section of the paper, I consider some further reasons for supposing that the central norm of assertion is that one ought not to assert that which one does not believe. After noting that it is plausible to suppose that assertion is the sole speech act that has the expression of belief as its proper end, I argue that consideration of the Gricean ambition mentioned above—viz. the ambition to show that people should be expected to have an interest in participating in conversational exchanges of this kind only on the assumption that they are conducted in general accordance with this norm —supports the claim that the sole constitutive individuating norm of assertion is that one ought not to assert that which one does not believe.

In the fifth section of the paper, I turn my attention to the consequences that the view that I have defended about the constitutive individuating norms of assertion has for a recent dispute about minimalist and deflationary theories of truth. Wright (1992) claims that considerations about the norms of assertion establish that deflationism is an inherently unstable view. Horwich (1998) denies that Wright’s arguments establish the conclusion for which he argues. Price (1998) argues that, while Horwich is right to criticise Wright’s arguments, it is nonetheless true that there are norms of assertion that deflationary theories of truth fail to capture. I argue that Price’s argument fails to establish that there are norms of assertion that deflationary theories of truth fail to capture; and, moreover, I argue that Wright and Price get great mileage from mistaken views about the norms of assertion.

In the sixth, and final, section of the paper, I provide the beginnings of a discussion of the connections that exist between norms of assertion, norms of belief, and norms of inquiry. In particular, I consider the impact that the adoption of the views of Kelly (2005) on the epistemic significance of disagreement have for discussion of the norms of assertion. If we suppose that disagreement between doxastic peers need give none of them reasons to change their minds, then I do not think that we should suppose that it is even a proximate aim of assertion, in general, to secure agreement between parties to a conversation.
There are many contemporary philosophers of religion who defend putative proofs or arguments for the existence or non-existence of God. In particular, there are many contemporary philosophers of religion who set out explicit arguments... more
There are many contemporary philosophers of religion who defend putative proofs or arguments for the existence or non-existence of God. In particular, there are many contemporary philosophers of religion who set out explicit arguments that they claim to be proofs or successful arguments for the existence or non-existence of God. The aim of this paper is to examine the prospects for proofs or successful arguments for the existence or non-existence of God. I begin with an attempt to establish terms for the subsequent discussion.
In a very interesting recent article (“Who Are The Best Judges Of Theistic Arguments?” Sophia 35, 2, 1996, pp.1–12), Mark Nelson argues that the best judges of arguments for the existence of God are theists whose belief in God is properly... more
In a very interesting recent article (“Who Are The Best Judges Of Theistic Arguments?” Sophia 35, 2, 1996, pp.1–12), Mark Nelson argues that the best judges of arguments for the existence of God are theists whose belief in God is properly basic. I propose to take up two questions here. First, does Nelson’s argument succeed in establishing his conclusion? Second, if Nelson’s argument were successful, what significance would it be appropriate to attribute to his conclusion? I shall begin with a brief rehearsal of his argument, and a discussion of some questions about the proper formulation of his argument which immediately arise.
I think that there is much about contemporary philosophy of religion that should change. Most importantly, philosophy of religion should be philosophy of religion, not merely philosophy of theism, or philosophy of Christianity, or... more
I think that there is much about contemporary philosophy of religion that should change. Most importantly, philosophy of religion should be philosophy of religion, not merely philosophy of theism, or philosophy of Christianity, or philosophy of certain denominations of Christianity, or the like. Here, however, I shall complain about one fairly narrow aspect of contemporary philosophy of religion that really irks me: its obsession with derivations that have as their conclusion either the claim that God exists or the claim that God does not exist. I shall work myself up by degrees.
Nick Trakakis and Yujin Nagasawa (2004) criticise the argument in Almeida and Oppy (2003). According to Trakakis and Nagasawa, we are mistaken in our claim that the sceptical theist response to evidential arguments from evil is... more
Nick Trakakis and Yujin Nagasawa (2004) criticise the argument in Almeida and Oppy (2003). According to Trakakis and Nagasawa, we are mistaken in our claim that the sceptical theist response to evidential arguments from evil is unacceptable because it would undermine ordinary moral reasoning. In their view, there is no good reason to think that sceptical theism leads to an objectionable form of moral scepticism.

We beg to differ. In our view, the criticisms of Trakakis and Nagasawa do not touch the heart of our objection to sceptical theism. However, in order to defend this contention, we need to begin by recapping the discussion to this point.
A number of authors have developed evidential arguments from evil in the past thirty years. Perhaps the best known evidential arguments from evil are those presented in Rowe (1979) and Draper (1989). We shall spend most of this chapter... more
A number of authors have developed evidential arguments from evil in the past thirty years. Perhaps the best known evidential arguments from evil are those presented in Rowe (1979) and Draper (1989). We shall spend most of this chapter examining these two arguments.
Research Interests:
Many philosophers seem to suppose that the argument of Plantinga (1974)—or a suitably elaborated variant thereof—utterly demolishes the kinds of “logical” arguments from evil developed in Mackie (1955). I am not at all convinced that this... more
Many philosophers seem to suppose that the argument of Plantinga (1974)—or a suitably elaborated variant thereof—utterly demolishes the kinds of “logical” arguments from evil developed in Mackie (1955). I am not at all convinced that this is a correct assessment of the current state of play. First, I think that Plantinga’s free will defence involves a hitherto undetected inconsistency. Second, I think that even if Plantinga’s free will defence is consistent, it relies upon some indefensible metaphysical assumptions. Third, I think that even if the metaphysical assumptions upon which Plantinga’s free will defence relies are defensible, there are serious questions to be raised about the moral assumptions which are made in that defence. Finally, I think that, even if Plantinga’s free will defence is acceptable, there are arguments closely related to those developed in Mackie (1955) that are not vulnerable to any variant of Plantinga’s free will defence, and yet that are clearly deserving of further examination.
In recent times, a number of philosophers have championed ‘sceptical theist’ responses to evidential arguments from evil. [Alston 1991, 1996; Bergmann 2001; Fitzpatrick 1981; Howard-Snyder 1996a, 1996c; van Inwagen 1991; Plantinga 1979,... more
In recent times, a number of philosophers have championed ‘sceptical theist’ responses to evidential arguments from evil. [Alston 1991, 1996; Bergmann 2001; Fitzpatrick 1981; Howard-Snyder 1996a, 1996c; van Inwagen 1991; Plantinga 1979, 1988; Wykstra 1984, 1996] The core idea behind these responses to evidential arguments from evil is that considerations of human cognitive limitations are alone sufficient to undermine those arguments. This core idea is developed in different ways. Some ‘sceptical theists’—[Bergmann 2001; Howard-Snyder 1996a, 1996c]—claim that consideration of human cognitive limitations in the realm of value are alone sufficient to undermine evidential arguments from evil. Other ‘sceptical theists’—[Alston 1991, 1996; van Inwagen 1991; Plantinga 1979, 1988; Wykstra 1984, 1996]—claim that consideration of human cognitive limitations in various spheres including the realm of value are alone sufficient to undermine evidential arguments from evil. Our response to these ‘sceptical theists’ is in two parts. First, we argue—against [Bergmann 2001, et al.]—that it isn’t true that considerations of human cognitive limitations in the realm of value are alone sufficient to undermine evidential arguments from evil. Second, we argue against [Alston 1991, 1996, et al.]—that it isn’t true that considerations of human cognitive limitations in various spheres including the realm of values are alone sufficient to undermine evidential arguments from evil.
In “Defining Art Historically” (BJA, 1979, pp.232-250), Jerrold Levinson defends the following definition: (R) X is a work of art at time t iff X is an object of which it is true at t that some person or persons having the appropriate... more
In “Defining Art Historically” (BJA, 1979, pp.232-250), Jerrold Levinson defends the
following definition:

(R) X is a work of art at time t iff X is an object of which it is true at t that some
person or persons having the appropriate proprietary right over X, nonpassingly
intends (or intended) X for regard-as-a-work-of-art, i.e. regard in any
way (or ways) in which objects in the extension of “art work” prior to t are or
were correctly (or standardly) regarded. (p.240)

Moreover, he suggests that this definition can form the generative component of a
recursive definition of art, in harness with the initial condition:

(I) Objects of the ur-arts are art works at t0 (and thereafter).

It seems to me that there are numerous difficulties which confront this definition. In
particular, there are difficulties involving:

(1) the inclusion of a condition involving “appropriate proprietary rights”;
(2) the reliance upon the intentions of independent individuals;
(3) Levinson’s account of the notion of “regard-as-a-work-of-art”; and
(4) the implicit insistence that art is necessarily backward-looking.

Since Levinson’s paper has recently received some favourable press (e.g. see Noel
Carroll, “Art, Practice, And Narrative”, Monist, pp.140-156, at p.155n.9), I think that
some discussion of these problems is in order. I shall consider them in turn.
In "A Defence Of The Institutional Definition Of Art" (Southern Journal Of Philosophy (1988), Vol. XXVI, No. 3, pp.317-334), Stephen Davies argues for the view that George Dickie's institutional definition of art can be amended in a way... more
In "A Defence Of The Institutional Definition Of Art" (Southern Journal Of Philosophy (1988), Vol. XXVI, No. 3, pp.317-334), Stephen Davies argues for the view that George Dickie's institutional definition of art can be amended in a way which renders it immune to the sorts of objections which are most commonly supposed to undermine it, but without leaving the definition open to any new and crippling objections. I think that Davies' argument is interesting, but unsuccessful; consequently, the main purpose of this paper is to explain why Davies' attempt to rehabilitiate the institutional definition of art fails.
M. W. Rowe, in “Why ‘Art’ Doesn’t Have Two Senses” (British Journal Of Aesthetics, Vol.31, No.3, July 1991, pp.214-221) defends a view which is quite similar to views defended by Collingwood and Barrett. Rowe’s main claims are: (i) that... more
M. W. Rowe, in “Why ‘Art’ Doesn’t Have Two Senses” (British Journal Of Aesthetics, Vol.31, No.3, July 1991, pp.214-221) defends a view which is quite similar to views defended by Collingwood and Barrett. Rowe’s main claims are: (i) that there can be no classificatory sense of the word ‘art’; and (ii) that the fact that there can be no classificatory sense of the word ‘art’ reveals that recent attempts to define the expression ‘work of art’ without making any reference to the value of art are doomed to fail.

I think that Rowe is right to insist that a definition of the expression ‘art’ will need to accord an important role to the notion of the value of works of art. However, I do not think that one needs to deny that there is a non-derivative classificatory sense of the expression “work of art” in order to defend that claim. The purpose of my paper is to explain how the previous two remarks can sit comfortably together.
Agnosticism has had some bad press in recent years. Nonetheless, I hope to show that agnosticism can be so formulated that it is no less philosophically respectable than theism and atheism. This is not a mere philosophical exercise; for,... more
Agnosticism has had some bad press in recent years. Nonetheless, I hope to show that agnosticism can be so formulated that it is no less philosophically respectable than theism and atheism. This is not a mere philosophical exercise; for, as it happens, the formulated position is -- I think -- the one to which I subscribe. I include a qualification here since it may be that the position to which I subscribe is better characterised as fallibilist atheism -- but more of that anon.
Michael Martin’s Atheism: A Philosophical Justification has the explicitly stated aim of showing “that atheism is a rational position and that belief in God is not” (24, 460).1 Furthermore, Martin tells us that he aims “to provide good... more
Michael Martin’s Atheism: A Philosophical Justification has the explicitly stated aim of showing “that atheism is a rational position and that belief in God is not” (24, 460).1 Furthermore, Martin tells us that he aims “to provide good reasons for being an atheist” (24), and to “defend” and “justify” atheism. However, Martin goes on to warn the reader that “no extended theory of rationality or justification is given” (25). And, in defence of this “general philosophical limitation”, he says that:

It seems to me that any attempt to justify [the few general comments that I make about rationality and justification] by subsuming them under a larger theory would be premature, given the controversial state of general epistemological theories. It is far better, in my view, to develop certain middle-level principles of justification that are in accord with our ordinary and scientific rational practice and to argue for atheism in terms of these than to justify atheism in terms of some larger and more controversial theory. (25-6)

While one might grant to Martin that it is acceptable to rest content with “certain middle-level principles of justification that are in accord with ordinary and scientific rational practice”, it would then be important to ask whether the principles of rationality, justification and argumentation that Martin adopts are actually in accord with “ordinary and scientific rational practice”. And, even if one were not to grant to Martin that it is acceptable to rest content with “certain middle-level principles of justification that are in accord with ordinary and scientific rational practice”, one would likely still wish to examine the assumptions about rationality, justification and argumentation that are at least implicit in Martin’s book.

Before we can turn to the task of trying to determine whether there are serious difficulties that arise here for Martin, we need to have a broad overview of the book before us. I shall present such an overview, and then return to an examination of the assumptions about rationality, justification and argumentation upon which Martin’s “defence” and “justification” of atheism appears to rest.
The thoughts that I have developed here depend upon assumptions that are highly controversial—and, in some cases, they depend upon assumptions that I have myself denied at other times and in other places. Consequently, I do not here... more
The thoughts that I have developed here depend upon assumptions that are highly controversial—and, in some cases, they depend upon assumptions that I have myself denied at other times and in other places. Consequently, I do not here pretend to be offering an argument on behalf of the views that I hold. Rather, I have offered the barest outlines of a view which I claim is capable of almost indefinite consistent refinement and development, and which I think is capable of standing with any of the competing worldviews that have been offered by those who believe in supernatural entities.
Research Interests:
In 1948, the BBC broadcast a debate between Bertrand Russell and Father Frederick Copleston on the existence of God (Russell and Copleston 1957). In that debate, Copleston claims: (1) that the existence of God can be proven by a... more
In 1948, the BBC broadcast a debate between Bertrand Russell and Father Frederick Copleston on the existence of God (Russell and Copleston 1957). In that debate, Copleston claims: (1) that the existence of God can be proven by a metaphysical argument from contingency; and (2) that only the postulation of the existence of God can make sense of our religious and moral experience. Russell replies by giving diverse reasons for thinking that these two claims are incorrect: there are various ways in which Copleston’s argument from contingency fails to be persuasive, and there are more plausible alternative explanations of our religious and moral experience. While there are many significant changes of detail, it is fair to say that the debate between Russell and Copleston typifies exchanges between theists and atheists in the second half of the twentieth century, and it is also fair to say that Russell’s contribution to this debate typifies the approaches of late twentieth-century atheists.

Speaking very roughly, we might divide the activities of atheists in the following way. First, some atheists have been concerned to argue that religious talk fails to be meaningful: there is no serious discussion to be had about, for example, the existence of God because one cannot even meaningfully deny the existence of God. Second, many atheists have been concerned to develop alternative worldviews to the kinds of worldviews that are presented in the world’s religions; and, in particular, many atheists have been concerned to develop naturalistic worldviews that leave no room for any kinds of supernatural entities. Third, some atheists have been interested in discussions of the ground rules for the arbitration of debates between theists and non-theists; and, in particular, some atheists have wanted to insist that there is an initial presumption in favor of atheism that leaves theistic opponents carrying the argumentative burden of proof. Fourth, many atheists have been concerned to raise objections against the plethora of theistic arguments that have been advanced, in particular on behalf of the claim that God exists. Fifth, some atheists have also been concerned to advance argument on behalf of atheism and, in particular, on behalf of the claim that God does not exist. Sixth, in the early part of the twenty-first century, some ‘new’ atheists have attempted to advance overarching critiques of religion—not merely theistic religion—in which even moderate religious belief is characterized as barbaric superstition. In what follows, we shall survey all of these different spheres of activity of atheists in the second half of the twentieth century.
Atheism is the rejection of theism: a‐theism. Atheists maintain some or all of the following claims: that theism is false; that theism is unbelievable; that theism is rationally unacceptable; that theism is morally unacceptable. Among... more
Atheism is the rejection of theism: a‐theism. Atheists maintain some or all of the following claims: that theism is false; that theism is unbelievable; that theism is rationally unacceptable; that theism is morally unacceptable.

Among arguments for atheism, there are arguments that are direct, indirect, and comparative.

Direct arguments for atheism aim to show that theism fails on its own terms: theism is meaningless, or incoherent, or internally inconsistent, or impossible, or inconsistent with known fact, or improbable given known fact, or less likely than not given known fact, or morally repugnant, and so forth.

Indirect arguments for atheism depend upon direct arguments for something else. Consider naturalism. Naturalism and theism are jointly inconsistent: they cannot both be true. Direct arguments for naturalism—arguments for the claim that naturalism is true, or rationally required, or morally required—are, eo ipso, arguments for atheism.

Comparative arguments for atheism are arguments for the theoretical superiority of something else to theism. Consider naturalism. An argument for the theoretical superiority of naturalism to theism is, eo ipso, an argument for atheism, even though such an argument need not aim to establish that naturalism is true, or rationally required, or morally required.
Philosophy in Australia and New Zealand has for some time now been experiencing something of a ‘golden age’. The richness of Australasia’s philosophical past, though less well known, should also not be forgotten. Australasian philosophy,... more
Philosophy in Australia and New Zealand has for some time now been experiencing something of a ‘golden age’. The richness of Australasia’s philosophical past, though less well known, should also not be forgotten. Australasian philosophy, although heavily indebted to overseas trends, includes much distinctive and highly original work.

The Companion contains a wide range of articles contributed by prominent philosophers and scholars, and includes biographical essays on selected philosophers, topics and controversies, as well as shorter entries on associations, research centres, departments, journals, pedagogy, and international links. Important philosophical contributions made by those working outside of the academy are also included, along with philosophy’s recent inroads into the wider community – in primary and secondary schools, community-based forums and ‘philosophy cafés’.

A Companion to Philosophy in Australia and New Zealand will provide scholars and the wider community, in Australia, New Zealand and beyond, with a greater appreciation of the philosophical heritage of this region, and will be a standard work of reference for many years to come.

Note that an electronic version of this book is freely available from the publishers: http://www.publishing.monash.edu/books/cpanz.html
In this second volume of The Antipodean Philosopher, Graham Oppy and N.N. Trakakis have brought together fourteen leading Australasian philosophers, inviting them to speak in a frank and accessible way about their philosophical lives: for... more
In this second volume of The Antipodean Philosopher, Graham Oppy and N.N. Trakakis have brought together fourteen leading Australasian philosophers, inviting them to speak in a frank and accessible way about their philosophical lives: for example, what drew them to a career in philosophy, what philosophy means to them, and their perceptions and criticisms of the ways in which philosophy is studied and taught in Australia and New Zealand.

The philosophers interviewed include Susan Dodds, Frank Jackson,  Moira Gatens, Philip Pettit,  Rae Langton, Graham Priest, Janna Thompson, Peter Singer and Michael Smith – philosophers who have distinguished themselves in the discipline, both nationally and internationally, over many years and in various branches of philosophy. What emerges from the discussion with these philosophers is a distinctive and engaging narrative of the history of philosophy in Australasia, its recent successes and flourishing, and the problems and prospects facing it in the twenty-first century.

These interviews will challenge and entertain anyone with an interest in contemporary philosophy and the challenges of living out the examined life today.
The 2008 Australasian Association of Philosophy (AAP) Conference in Melbourne was attended by 345 delegates: 245 from Australia, 27 from New Zealand, 8 from Singapore, and 55 from a range of other countries. Over the course of the... more
The 2008 Australasian Association of Philosophy (AAP) Conference in Melbourne was attended by 345 delegates: 245 from Australia, 27 from New Zealand, 8 from Singapore, and 55 from a range of other countries. Over the course of the conference, there were 265 papers presented in thirteen different subject streams and at a diverse range of symposia. The conference concluded with an overlapping three-day mini-conference on relations between ‘analytic’ and ‘continental’ philosophy.

By way of contrast, the 1999 AAP conference was also held in Melbourne. This conference was attended by 275 delegates who presented 213 papers. The papers were not streamed, though there were three special themes for the conference: Wittgenstein; Fictionalism; and ‘Beyond Analysis’. Among the delegates who presented papers, there were 164 from Australia, 11 from New Zealand, and 38 from a range of other countries. This conference served as an umbrella for conferences of the Australasian Association of Logic (AAL), Women in Philosophy (WiP), and the Australasian Society for Asian and Comparative Philosophy (ASACP), and 74 of the 213 papers presented were to these associated conferences.

Comparison between these two conferences suggests that philosophy in Australasia has prospered in the first decade of the twenty-first century. On almost every measure, these numbers indicate an increase over the course of the decade: more conference delegates, more conference papers, and more focussed debates on matters of contemporary concern. In what follows, we shall look more closely at the current state of Australasian philosophy, to see whether this optimistic view can be sustained.

We begin with a brief overview of the state of higher education in Australia, and then a similarly brief overview of the state of the humanities. This overview establishes context that is necessary for a proper evaluation of the performance of philosophy in the past decade.
On February 25, 2008, in an article entitled ‘Faith School Boom “Creates Division”’, Michael Bachelard of The Age reported: The principal of Chairo Christian School in Drouin, Rob Bray, said that both evolution and creationism were... more
On February 25, 2008, in an article entitled ‘Faith School Boom “Creates Division”’, Michael Bachelard of The Age reported:

The principal of Chairo Christian School in Drouin, Rob Bray, said that both evolution and creationism were taught in his school's science class. "We don't hide the fact that there is a theory of evolution, and that's how we'd present it, as a theory," Mr. Bray said. "We teach it, explain what it is, and at the same time we present clearly and fairly, and we believe convincingly, the fact that our position as a school is that God created the heaven and earth … There wouldn't be any point of being a faith-based school if we didn't think that God was the creator."

The article noted that there are now more than 200,000 children—almost 40% of non-government school students—attending a religious school outside the main Catholic, Anglican and Uniting systems. Further, it expressed the concerns of Professor Barry McGaw, head of the National Curriculum Board, and the Deputy Prime Minister, Julia Gillard, that the rapid growth of faith-based schools threatens the social cohesion of the nation. "These people often form a narrowly focused school that is aimed at cementing the faith it's based on … If we continue as we are, I think we'll just become more and more isolated sub-groups in our community”, said Professor McGaw.

Not so long ago, most people would have found it unthinkable that ‘creation science’ could come to occupy such an important place in public debate, that it could be taught so widely in science classes in our schools, and that it could be accepted by such a large part of the population. In what follows, I will provide a short history of this development, and then consider how we ought to respond to the current state of play.
The History of Philosophy in Australia and New Zealand is a comprehensive account of the historical development of philosophy in Australia and New Zealand, from the establishment of the first Philosophy Chair in Australasia in 1886 at the... more
The History of Philosophy in Australia and New Zealand is a comprehensive account of the historical development of philosophy in Australia and New Zealand, from the establishment of the first Philosophy Chair in Australasia in 1886 at the University of Melbourne to the current burgeoning of Australasian philosophy. The work is divided into two broad sections, the first providing an account of significant developments and events during various periods in the history of Australasian philosophy, and the second focusing on ideas and theories that have been influential in various disciplines within Australasian philosophy. The work consists of chapters contributed by various philosophers, on specific fields of inquiry or historical periods within Australasian philosophy.
The phrase “The Turing Test” is most properly used to refer to a proposal made by Turing (1950) as a way of dealing with the question whether machines can think. According to Turing, the question whether machines can think is itself “too... more
The phrase “The Turing Test” is most properly used to refer to a proposal made by Turing (1950) as a way of dealing with the question whether machines can think. According to Turing, the question whether machines can think is itself “too meaningless” to deserve discussion (442). However, if we consider the more precise—and somehow related—question whether a digital computer can do well in a certain kind of game that Turing describes (“The Imitation Game”), then—at least in Turing's eyes—we do have a question that admits of precise discussion. Moreover, as we shall see, Turing himself thought that it would not be too long before we did have digital computers that could “do well” in the Imitation Game.

The phrase “The Turing Test” is sometimes used more generally to refer to some kinds of behavioural tests for the presence of mind, or thought, or intelligence in putatively minded entities. So, for example, it is sometimes suggested that The Turing Test is prefigured in Descartes' Discourse on the Method. (Copeland (2000:527) finds an anticipation of the test in the 1668 writings of the Cartesian de Cordemoy. Gunderson (1964) provides an early instance of those who find that Turing's work is foreshadowed in the work of Descartes.) In the Discourse, Descartes says:

If there were machines which bore a resemblance to our bodies and imitated our actions as closely as possible for all practical purposes, we should still have two very certain means of recognizing that they were not real men. The first is that they could never use words, or put together signs, as we do in order to declare our thoughts to others. For we can certainly conceive of a machine so constructed that it utters words, and even utters words that correspond to bodily actions causing a change in its organs. … But it is not conceivable that such a machine should produce different arrangements of words so as to give an appropriately meaningful answer to whatever is said in its presence, as the dullest of men can do. Secondly, even though some machines might do some things as well as we do them, or perhaps even better, they would inevitably fail in others, which would reveal that they are acting not from understanding, but only from the disposition of their organs. For whereas reason is a universal instrument, which can be used in all kinds of situations, these organs need some particular action; hence it is for all practical purposes impossible for a machine to have enough different organs to make it act in all the contingencies of life in the way in which our reason makes us act. (Translation by Robert Stoothoff)

Although not everything about this passage is perfectly clear, it does seem that Descartes gives a negative answer to the question whether machines can think; and, moreover, it seems that his giving this negative answer is tied to his confidence that no mere machine could pass The Turing Test: no mere machine could talk and act in the way in which adult human beings do. Since Descartes explicitly says that there are “two very certain means” by which we can rule out that something is a machine—it is, according to Descartes, inconceivable that a mere machine could produce different arrangements of words so as to give an appropriately meaningful answer to whatever is said in its presence; and it is for all practical purposes impossible for a machine to have enough different organs to make it act in all the contingencies of life in the way in which our reason makes us act—it seems that he must agree with the further claim that nothing that can produce different arrangements of words so as to give an appropriately meaningful answer to whatever is said in its presence can be a machine. Given the further assumption—which one suspects that Descartes would have been prepared to grant—that only things that think can produce different arrangements of words so as to give an appropriately meaningful answer to whatever is said in their presence, it seems to follow that Descartes would have agreed that the Turing Test would be a good test of his confident assumption that there cannot be thinking machines. Given the knowledge that something is indeed a machine, evidence that that thing can produce different arrangements of words so as to give an appropriately meaningful answer to whatever is said in its presence is evidence that there can be thinking machines.

The phrase “The Turing Test” is also sometimes used to refer to certain kinds of purely behavioural allegedly logically sufficient conditions for the presence of mind, or thought, or intelligence, in putatively minded entities. So, for example, Ned Block's “Blockhead” thought experiment is often said to be a (putative) knockdown objection to The Turing Test. (Block (1981) contains a direct discussion of The Turing Test in this context.) Here, what a proponent of this view has in mind is the idea that it is logically possible for an entity to pass the kinds of tests that Descartes and (at least allegedly) Turing have in mind—to use words (and, perhaps, to act) in just the kind of way that human beings do—and yet to be entirely lacking in intelligence, not possessed of a mind, etc.

The subsequent discussion takes up the preceding ideas in the order in which they have been introduced. First, there is a discussion of Turing's paper (1950), and of the arguments contained therein. Second, there is a discussion of current assessments of various proposals that have been called “The Turing Test” (whether or not there is much merit in the application of this label to the proposals in question). Third, there is a brief discussion of some recent writings on The Turing Test, including some discussion of the question whether The Turing Test sets an appropriate goal for research into artificial intelligence. Finally, there is a very short discussion of Searle's Chinese Room argument, and, in particular, of the bearing of this argument on The Turing Test.
There has recently been a surge in publications espousing arguments from consciousness for the existence of God. In particular, J. P. Moreland has produced a series of articles in which he promotes the virtues of the following... more
There has recently been a surge in publications espousing arguments from consciousness for the existence of God.

In particular, J. P. Moreland has produced a series of articles in which he promotes the virtues of the following argument:

1. Mental events are genuine nonphysical mental entities that exist.
2. Specific mental and physical event types are regularly correlated.
3. There is an explanation for these correlations.
4. Personal explanation is different from natural scientific explanation.
5. The explanation for these correlations is either a personal or natural scientific explanation.
6. The explanation is not a natural scientific one.
7. Therefore, the explanation is a personal one.
8. If the explanation is personal, then it is theistic.
9. Therefore, the explanation is theistic.

In this chapter, I propose to focus on Moreland’s defence of arguments from consciousness. In particular, I shall argue against his claim that considerations about consciousness favour theism over naturalism.

Moreland’s argument that considerations about consciousness favour theism over naturalism depends crucially upon his account of naturalism, his account of theoretical virtues, and his method of assessing the relative merits of theism and naturalism. So I begin with some discussion of his treatment of each of these topics.
In The Christian God, Richard Swinburne provides a definition of ‘physical property’ and ‘mental property’ which has various counter–intuitive consequences. The purpose of this note is to draw attention to these consequences, and to... more
In The Christian God, Richard Swinburne provides a definition of ‘physical property’ and ‘mental property’ which has various counter–intuitive consequences. The purpose of this note is to draw attention to these consequences, and to suggest some ways in which Swinburne’s account could be improved. (Essentially the same definitions are to be found in The Evolution of the Soul; I shall focus on the discussion in The Christian God because it is much more recent. All page numbers refer to this later work.)
This book divides naturally into three parts. The first part consists of two chapters, the first of which sets out what Moreland takes to be ‘the epistemic backdrop’ against which ‘the argument from consciousness’ is properly assessed,... more
This book divides naturally into three parts. The first part consists of two chapters, the first of which sets out what Moreland takes to be ‘the epistemic backdrop’ against which ‘the argument from consciousness’ is properly assessed, and the second of which presents several ‘versions’ of ‘the argument from consciousness.’ The second part consists of five chapters, each of which is devoted to a close analysis of the work of a particular theorist: John Searle, Tim O’Connor, Colin McGinn, David Skrbina, and Philip Clayton. The third part consists of two chapters, the first of which develops and defends ‘the Autonomy thesis’—roughly, the claim that, where central questions of philosophy have answers, those answers do not substantively depend on science—and the second of which argues that it is fear of God that drives ‘current and confident acceptance of strong physicalism and naturalism and rejection of dualism’ (176).

I have discussed much of the material in the first two chapters of this book elsewhere—see my chapter on arguments from consciousness in C. Meister, J. P. Moreland, and K. Sweis (eds.) (forthcoming). In this review, I propose to focus more attention on the final two chapters. However, I shall begin with a discussion of the presentation of ‘the argument from consciousness’ in Chapter Two.
Philipse (2012) provides an extended critique of the argument of Swinburne (1979). This critique is the centrepiece of a broader critique of natural theology: according to Philipse, there are several devastating criticisms that can be... more
Philipse (2012) provides an extended critique of the argument of Swinburne (1979). This critique is the centrepiece of a broader critique of natural theology: according to Philipse, there are several devastating criticisms that can be made of theism, but Swinburne's 'argumentative strategy … [is] … the most promising one from an apologetic point of view' (xiii). Philipse uses 'the strategy of subsidiary arguments' to press against theism. First, he claims, theism is not a meaningful hypothesis: in particular, the defence of the coherence of theism in Swinburne (1977) fails. Second, he says, theism has no explanatory power: in particular, Swinburne's attributions of creative intentions to God are nothing more than anthropomorphic projections. Third, according to Philipse, theism is defeated by empirical data: in particular, the cumulative argument of Swinburne (1979) fails to establish that, given all of the relevant evidence, the existence of God is more likely than not. There is a prima facie awkwardness involved in Philipse's use of the strategy of subsidiary arguments. The 'cumulative' case that he builds for the rejection of theism is not one in which there are three mutually supporting lines of criticism, but rather one in which there are three pair-wise strongly incompatible lines of criticism. If, for example, the first line of criticism succeeds, and theism is shown not to be a meaningful hypothesis, then it seems that all of the discussion of the second and third lines of criticism is shown to be meaningless: if theism is meaningless, then there cannot be detailed meaningful discussion of how theism fares when it is confronted with empirical data. But the very act of engaging in detailed discussion of how theism fares when it is confronted with empirical data takes for granted the meaningfulness of that kind of discussion! Perhaps one might think that there are rationally defensible views according to which theism is not a meaningful hypothesis. However, if there are rationally defensible views of this kind, then, it seems to me, those who endorse them will decline any invitation to debate the question how theism fares when confronted with empirical data: because, by the lights of those who suppose that theism is not a meaningful hypothesis, any such debate will itself have no meaning. Given these remarks, and given that I wish to take up the question of how theism fares when confronted with empirical data, it is perhaps not surprising that I shall simply accept without further argument that theism is a meaningful hypothesis, and that theism has at least some explanatory power.
Research Interests:
Perhaps unsurprisingly, Robert Koons finds fault with my criticisms of his “new” cosmological argument; perhaps no less surprisingly, I find his replies unsatisfactory. There seems to be little prospect that our views about the merits of... more
Perhaps unsurprisingly, Robert Koons finds fault with my criticisms of his “new” cosmological argument; perhaps no less surprisingly, I find his replies unsatisfactory. There seems to be little prospect that our views about the merits of the argument will converge; and that might be taken to be a reason for me to hold my counsel. However, there are various ways in which Koons rather egregiously misrepresents what I said previously; some good may come from getting clearer about where we genuinely disagree. So here goes.
Koons (2008) argues for the very surprising conclusion that ‘any exception to the principle of general causation [i.e., the principle that everything has a cause] that is narrow enough to avoid a collapse into global scepticism about... more
Koons (2008) argues for the very surprising conclusion that ‘any exception to the principle of general causation [i.e., the principle that everything has a cause] that is narrow enough to avoid a collapse into global scepticism about empirical knowledge is also narrow enough to permit the construction of a successful proof of God’s existence’ (106). While Koons supposes that there are two ways in which a ‘principle of general causation’ could be connected to the possibility of empirical knowledge—namely (i) as an objective fact needed as the ground for the reliability of our cognitive processes, and (ii) as a subjectively required presumption needed for immunity to internal defeaters—he does little more than sketch the beginnings of the development of an argument of the first kind, reserving almost all of his attention for the development of an argument of the second kind. We shall follow his lead.
Richard Gale and Alexander Pruss—“A New Cosmological Argument”, Religious Studies 35, 1999, pp.461–76—present a cosmological argument which they claim is an improvement over familiar cosmological arguments because it relies upon a weaker... more
Richard Gale and Alexander Pruss—“A New Cosmological Argument”, Religious Studies 35, 1999, pp.461–76—present a cosmological argument which they claim is an improvement over familiar cosmological arguments because it relies upon a weaker version of the principle of sufficient reason than is used in those more familiar arguments. I shall argue that this claim is mistaken: their new argument is no better than the more familiar arguments which they take as their benchmark. In order to explain why this is so, I shall need briefly to explain the theoretical framework in which their proof is located, and to recapitulate the main details of the proof. After I have done this, I shall go on to give my explanation.
We can think of Christian philosophers as beginning with the following picture: GOD ⇒ THE WORLD where both ‘THE WORLD’ and ‘⇒’ are items in need of further explanation. In typical cosmological arguments, what Christian philosophers then... more
We can think of Christian philosophers as beginning with the following picture:

GOD ⇒ THE WORLD

where both ‘THE WORLD’ and ‘⇒’ are items in need of further explanation. In typical cosmological arguments, what Christian philosophers then aim to show is that there is a * such that

* ⇒ THE WORLD

where ‘THE WORLD’ and ‘⇒’ are given that particular further explanation.

What I propose to do in this paper, is to set out my reasons for thinking that there is no * such that that * ⇒ THE WORLD, on any interpretation of ‘⇒’ and ‘THE WORLD’ that would set this claim at odds with naturalism.
The subject of physical eschatology is in its infancy. While popular books about the Big Bang and the cosmological origins of life have inundated bookshops during the past couple of decades, it is only in the past few years that a trickle... more
The subject of physical eschatology is in its infancy. While popular books about the Big Bang and the cosmological origins of life have inundated bookshops during the past couple of decades, it is only in the past few years that a trickle of books about the distant future and the probable cosmological extinction of life has appeared. Moreover, this disproportion is also apparent in the technical, scientific literature: there have only been a handful of serious studies of far distant cosmological futures. Perhaps some of this relative neglect can be attributed to the more speculative nature of eschatology—and perhaps more can be attributed to the psychological discomfort to which serious study of the future of the universe seems apt to give rise—but, whatever the reasons, it seems that we are currently seeing some redressing of this imbalance. In particular, the work of Adams and Laughlin (1997) serves as a useful benchmark for progress in this area.

I shall begin with the assumption that the causally connected universe is all the universe that there is, i.e. I shall begin with the assumption that there are no causally disconnected regions. Under this assumption, I shall explore the likely future of the universe under the further assumption that the currently observable universe is representative of the universe as a whole. Since the current best opinion is that the currently observable universe is open, this means that in this most likely future, the universe itself is open, and ‘exists forever’. The future of open universes much like our own is the topic of Section 1 of this paper.

If we give up the assumption that the currently observable universe is representative of the universe as a whole—but while still holding on to the assumption that there are no causally disconnected regions of the universe—then it is possible to suppose that the universe is closed. If we suppose that the universe is closed, then we can explore its likely future as it journeys towards the Big Crunch. The future of closed universes much like our own is the topic of Section 2 of this paper.

If we give up the assumption that there are no causally disconnected regions—i.e. if we allow that there are, or might be, regions of the universe with which our region of the universe shall never be in causal contact—then all kinds of speculative possibilities emerge. We shall canvass a few of these possibilities, without making any attempt to assign a value to their likelihood. These speculations are the topic of Section 3 of this paper.

One main focus of interest in the future of the universe concerns the future of human beings, life, intelligent entities, and the like. What does the future hold for us, and those like us? In Section 4 of this paper, we shall look at some suggestions about the likely fate of our descendants in each of the kinds of scenarios discussed in the first three sections of the paper.

In the final section of the paper, we shall ask some questions about the significance of the answers to—and the significance of the activity of seeking answers to—the questions which are the topic of investigation in Section 4. Since the middle of the nineteenth century, the ‘Heat Death of the Universe’ has exercised a considerable grip on the popular imagination—and on the imagination of scientists as well—but is there really any good reason for this obsession? Since this is a vast—and interesting—topic, the remarks to be made in this final section are preliminary at best.
In this paper, I shall assume that some kind of Big Bang theory is correct, i.e. that the currently observable universe evolved from a superdense state over a period of ten to twenty billion years. While this is clearly the view which is best supported by the current evidence, there are still supporters of various kinds of Steady State theories, which hold that the resources of the universe are continually replenished by the spontaneous creation of matter and energy to fill the space created by expansion. On these kinds of views, there is in principle no reason why there couldn’t be galaxies existing indefinitely into the future: some galaxies would die, collapsing into black holes which then evaporate away via Hawking radiation; but new galaxies would continue to form, and the process of forming stars and planets would continue forever. No doubt, there are difficulties in getting the spontaneous creation of matter and energy to happen in the right way to ensure that there will be new galaxies: but I shall leave these matters to enthusiasts of steady state theories.
There has been a recent upsurge of interest in ‘physical eschatology’ and, in particular, in the idea that, in the far–distant future, intelligent life will spread from earth to the rest of the universe. Although some of the early seeds... more
There has been a recent upsurge of interest in ‘physical eschatology’ and, in particular, in the idea that, in the far–distant future, intelligent life will spread from earth to the rest of the universe. Although some of the early seeds were sown by Freeman Dyson, the main proponent of this idea has been Frank Tipler. Tipler (1995)—a book devoted solely to the elaboration of Tipler’s ‘Omega Point Theory’—was released with an enormous fanfare, and it spent many weeks near the top of the bestseller charts. In this paper, I cast a critical eye over some aspects of the theory which Tipler develops.
A person is offered a choice between two envelopes, A and B. She is told that one envelope contains twice as much money as the other but has no information as to which one that is. She chooses A, say. Before opening A she asks herself... more
A person is offered a choice between two envelopes, A and B. She is told that one envelope contains twice as much money as the other but has no information as to which one that is. She chooses A, say. Before opening A she asks herself whether she ought to have taken B instead. There is a line of reasoning which suggests that she should. Suppose that the amount of money in A is $x. Then B either contains $2x or $0.5x. Each possibility is equally likely, hence the expected value of taking B is 0.5.$2x + 0.5.$0.5x = $1.25x, a gain of $0.25x. This conclusion cannot be right. The mere choosing of A cannot give her a reason to say that she ought to have picked up B instead. For the situation is symmetrical as between A and B, at least until one of them is opened. Moreover, had she chosen B initially the same reasoning would suggest that she ought to have chosen A instead. No matter which choice she makes the reasoning leads to the conclusion that she made the wrong choice!

It is easy to say how she ought to have done the expected value calculation so as to avoid the absurd (clearly mistaken) result. She should have reasoned that the following two situations are equally likely: A contains $x, B contains $2x; and A contains $2x, B contains $x. The expected value of taking A is then $1.5x, and the expected value of taking B is also $1.5x, and hence there is no reason for her to judge that she should have chosen B instead of A. The problem is to say what is wrong with the first way of doing the calculation.
Sober claims: (i) that Hume produces no good criticisms of design arguments; (ii) that Hume offers no serious alternative explanations of (biological) design; and (iii) that design was the best explanation available prior to 1859. Dawkins... more
Sober claims: (i) that Hume produces no good criticisms of design arguments; (ii) that Hume offers no serious alternative explanations of (biological) design; and (iii) that design was the best explanation available prior to 1859. Dawkins claims: (i) that one could not have been an intellectually fulfilled atheist prior to 1859; (ii) that Hume offered no alternative explanations of (biological) design; and (iii) that Paley knew that there must be some non-Humean explanation of apparent biological design. Mackie claims: (i) that the work of Darwin greatly diminishes the plausibility of design arguments; and (ii) that there are no reciprocal adjustments in cosmology which are as plausible candidates for evidence of design as cases from biology. None of these claims can withstand scrutiny, as I shall now attempt to demonstrate. More exactly, I shall argue: (a) that the claim that Hume offered no (serious) alternative explanations of apparent (biological) design is, perhaps, strictly correct, but that, taken alone, it misrepresents the strength of the resources which Hume provided for developing (serious) alternative explanations of apparent (biological) design; and (b) that all of the other claims are simply false.
The main aim of this paper is to examine an almost universal assumption concerning the structure of Paley’s argument for design. Almost all commentators suppose that Paley’s argument is an inductive argument—either an argument by analogy... more
The main aim of this paper is to examine an almost universal assumption concerning the structure of Paley’s argument for design. Almost all commentators suppose that Paley’s argument is an inductive argument—either an argument by analogy or an argument by inference to the best explanation. I contend, on the contrary, that Paley’s argument is actually a straightforwardly deductive argument. Moreover, I argue that, when Paley’s argument is properly understood, it can readily be seen that it is no good. Finally—although I do not stress this very much—I note that the points that I make about Paley’s argument can carry over to modern design arguments that are based upon the argument which Paley actually gives.
In Oppy (2002), I argued for the view that, contrary to received opinion, Paley’s argument for design is a deductive argument that is subject to decisive objections. Schupbach (2005) argues that I fail to show that Paley’s argument for... more
In Oppy (2002), I argued for the view that, contrary to received opinion, Paley’s argument for design is a deductive argument that is subject to decisive objections. Schupbach (2005) argues that I fail to show that Paley’s argument for design is a deductive argument, whence it surely follows that the objections that I raised are irrelevant. While I think that Schupbach overstates the case against the view that Paley’s argument for design is a deductive argument, I am persuaded that, at best, it is unclear whether or not we should hold that Paley’s argument is deductive. However, I insist that it doesn’t matter whether Paley’s argument is deductive or inductive: what matters is that the kinds of objections that I raised in Oppy (2002) serve to defeat Paley’s argument even if it is properly taken to be an inductive argument.
Some theists argue that nothing comes from nothing on grounds of experience. But experience no less confirms that there is no causing without changing: no thing causes a change in a second thing without itself undergoing some change.... more
Some theists argue that nothing comes from nothing on grounds of experience. But experience no less confirms that there is no causing without changing: no thing causes a change in a second thing without itself undergoing some change. Considerations about the kinds of causal principles licensed by experience do nothing to favour theism over naturalism.
I shall begin by criticising the two most promising extant accounts of omnipotence . After providing various reasons for finding these accounts unsatisfactory, I shall then go on to make some suggestions about how the notion of... more
I shall begin by criticising the two most promising extant accounts of omnipotence . After providing various reasons for finding these accounts unsatisfactory, I shall then go on to make some suggestions about how the notion of omnipotence should be understood.
There are many different views that have been held about the content of the idea or concept of God, and many different suggestions that have been made about how to define or analyse the name ‘God’. In this paper, I defend the suggestion... more
There are many different views that have been held about the content of the idea or concept of God, and many different suggestions that have been made about how to define or analyse the name ‘God’. In this paper, I defend the suggestion that to be God is just to be the one and only god, where to be a god is to be a supernatural being or force that has and exercises power over the natural world but that is not, in turn, under the power of any higher ranking or more powerful category of beings or forces. While many will take this to be a rather radical suggestion, it seems to me that there are many good reasons for adopting this proposal, and that there are no telling reasons that speak against it. Among the other controversial claims that are defended in this paper—and that I take to be plausible consequences of this main claim—I might mention in particular, the claim that there can be no more than one God, the claim that ‘God’ is not a title-term, and the claim that the use of the name ‘God’ by non-believers is not parasitic on the use of this name by believers. Thinking hard about the use of the name ‘God’ turns up all kinds of interesting consequences.
The aim of this paper is to examine the difficulties that belief in a paradisiacal afterlife creates for orthodox theists. In particular, we consider the difficulties that arise when one asks whether there is freedom in Heaven, i.e.... more
The aim of this paper is to examine the difficulties that belief in a
paradisiacal afterlife creates for orthodox theists. In particular, we consider the
difficulties that arise when one asks whether there is freedom in Heaven, i.e. whether
the denizens of Heaven have libertarian freedom of action. Our main contention is
that this “Problem of Heaven” makes serious difficulties for proponents of free will
theodicies and for proponents of free will responses to arguments from evil.
Of all the core doctrines of traditional Western theology, perhaps none has received more summary dismissals than the doctrine of divine simplicity. Consider, for example, the throwaway remark made by Paul Fitzgerald (1985:262) in his... more
Of all the core doctrines of traditional Western theology, perhaps none has received more summary dismissals than the doctrine of divine simplicity. Consider, for example, the throwaway remark made by Paul Fitzgerald (1985:262) in his discussion of the views of Eleanor Stump and Norman Kretzmann on issues concerning divine eternity:

This feature of eternality may be impossible to reconcile with the divine simplicity. But that doctrine never had much to recommend it anyhow. (My emphasis.)

Fitzgerald’s attitude towards the doctrine of divine simplicity is quite widespread, and is shared by many who are not otherwise particularly ill-disposed towards traditional Western theology. Even those who are far better disposed towards traditional Western theology are usually prepared to concede that the doctrine is hard to understand, let alone to accept. Thus, for example, at the beginning of her entry on simplicity in the Blackwell Companion to Philosophy of Religion, Eleanor Stump (1997:250) writes:

Among the traditionally recognised divine attributes regularly discussed by medieval theologians and accepted by them as part of orthodox religious belief, the strangest and hardest to understand is simplicity. (My emphasis.)

Given the nature of his book, it should come as no surprise that Richard Gale (1991) gives pretty short shrift to the doctrine of divine simplicity. Taking into account his delight in philosophical combat, it seems appropriate to use this occasion to try to defend the apparently indefensible (and not only from Gale’s attack on it, though that shall serve as our paradigm).

I shall begin with Gale’s explanation of the doctrine, and his reasons for dismissing it. I shall then go on to suggest an alternative understanding of the view which seems to me to avoid the difficulties which Gale and other contemporary philosophers have raised. I shall not be claiming that my “alternative understanding” is what traditional theologians had in mind; however, I do want to claim that my “rational reconstruction” has many of the properties which those theologians wanted the doctrine to have.

Of course, in undertaking to defend the doctrine, I am not undertaking to defend its truth; I’m not a theist so, a fortiori, I think that the doctrine is false. What I do want to defend is the rational acceptability of the doctrine: given that you buy Western theism in general, there is a coherent—albeit controversial—doctrine concerning divine simplicity which is also available to you.
One family of challenges to theistic belief derives from considerations concerning the claim that there is an omnipotent, omniscient, eternal, perfectly free, perfectly evil sole creator of the universe ex nihilo. These challenges begin... more
One family of challenges to theistic belief derives from considerations concerning the claim that there is an omnipotent, omniscient, eternal, perfectly free, perfectly evil sole creator of the universe ex nihilo. These challenges begin with the claim that a case can be made for the existence of this being—call it God*—which “parallels” the case which can be made for the existence of God. (Perhaps one might think that it would be more accurate to say that the claim is that the case for God* is just as good or bad as the case for God, and that part of the case for God* is contrived simply by mimicking or paralleling the case for God. For, prima facie at least, it seems that there are extra wrinkles which are needed in the case of God* to construct arguments from scripture, or revelation, or religious experience, or religious authority, etc. However, proponents of the challenges to theism which are under consideration ought to reply that the kinds of ‘evidence’ adverted to here are equally well explained on the hypothesis that God exists—where the explanation goes via God’s good intentions to help us—and on the hypothesis that God* exists—where the explanation goes via God*’s evil intentions to harm us. I shall suppose that we should allow this generous construal of the notion of a ‘parallel’ case, and that no harm will follow from this concession.)

One kind of response to this family of challenges on behalf of theistic belief would be to deny that the mimicking arguments are genuinely parallel—e.g. to claim that the ontological or cosmological or teleological or ... argument for God is stronger than the corresponding argument for God*, or that the problem of evil is a weaker argument against God than the problem of good is against God*. It seems to me that the kind of response looks prima facie rather unpromising; in any case, I propose to proceed under the pro tem assumption that this kind of response won’t work. (Those who disagree with my judgement here should for now take me to conducting a ‘conditional’ investigation: what can be said in response to these kinds of challenges to theistic belief if one concedes that the mimicking arguments for God* do genuinely parallel the traditional arguments for God?) Instead, I shall focus attention on a line of response which aims to establish that there are reasons for thinking that the concept of God* is incoherent in a way in which the concept of God is not. In particular, I shall consider the suggestion that the notion of an omniscient and perfectly evil being can be shown to be incoherent in ways which tend not at all to establish that the notion of an omniscient and perfectly good being is incoherent. If this suggestion is correct, then—other things being equal (as the proponents of the objection hold that they are!)—it seems that the hypothesis that God exists is clearly to be preferred to the hypothesis that God* exists.

I propose to argue that, even if these counterarguments do establish that the hypothesis that God exists is clearly to be preferred to the hypothesis that God* exists, this is not enough to show that theists are home free—for there are many other alternative Gods for whom ‘parallel’ cases could be constructed, and for which this particular counterargument is ineffective. I shall then go on to consider the consequences of this claim for the status of the debate between theists and their opponents. (I shall also argue that there are serious questions to be raised about the counterarguments against God*. However, I shall not place too much emphasis on these questions in this paper.) In order to get to these considerations, some preliminary scene–setting is required.
As a matter of historical fact, most philosophers and theologians who have defended traditional theistic views have been moral realists. Some “divine command” theorists have held that the good is constituted by the content of divine... more
As a matter of historical fact, most philosophers and theologians who have defended traditional theistic views have been moral realists. Some “divine command” theorists have held that the good is constituted by the content of divine approval -- i.e. that things are good because, and insofar as, they have divine approval. However, even amongst those theists who hold that the good is independently constituted -- i.e. those who hold that God’s pattern of approval is explained by the fact that he approves of all and only that which is good -- the dominant meta-ethic has been strongly realistic.

As a further matter of historical fact, one of the main motives for the development of non-realist meta-ethics has been the desire to give an adequate atheistic account of the nature of the good. Thus many subjectivist, projectivist, and error-theoretic accounts of the good were developed in the context of atheistic enquiries. Of course, atheistic accounts of the good can be realist -- e.g. G. E. Moore’s non-natural objectivism. However, the question which I wish to take up is whether theistic accounts of the good can be non-realist.

In fact, I shall argue that the traditional philosophical conception of God requires a commitment to moral realism. Despite the range of non-realist meta-ethics which have been developed, there is none which is compatible with theism. Consequently, there is a hitherto ignored argument against theism which emerges for assessment, viz:

1. Ethical realism is a necessary consequence of traditional theism.
2. Ethical realism is false.
3. (Therefore) Traditional theism is false.

In this paper I shall focus on 1. -- i.e. I shall not attempt to provide an argument for 2. However, I believe that there are independent arguments -- i.e. arguments independent of the debate about theism -- which favour the view that ethical realism is false. Consequently, I believe that the argument of this paper makes a contribution to the case against theism.
Among challenges to Molinism, the challenge posed by divine prophecy of human free action has received insufficient attention. We argue that this challenge is a significant addition to the array of challenges that confront Molinism.
The three attached documents are pre-proof copies of my contributions to P. Gould (ed.) Beyond the Control of God? Six Views on the Problem of God and Abstract Objects New York: Bloomsbury (44-5; 72-4; 104-6; 134-6; 162-4; 169-81;... more
The three attached documents are pre-proof copies of my contributions to P. Gould (ed.) Beyond the Control of God? Six Views on the Problem of God and Abstract Objects New York: Bloomsbury (44-5; 72-4; 104-6; 134-6; 162-4; 169-81; 192-5).

I defend the view that, in the context of the broader dispute between theists and naturalists, there is no significant difference between the defensible accounts that theists and naturalist can give of abstract objects.
This paper provides a critical review of Brian Leftow's *God and Necessity* (OUP, 2012). Much of this chapter is reprinted in my *Describing Gods* (CUP, 2014). I have also have a much shorter review of Leftow's book in the *Times Literary... more
This paper provides a critical review of Brian Leftow's *God and Necessity* (OUP, 2012). Much of this chapter is reprinted in my *Describing Gods* (CUP, 2014). I have also have a much shorter review of Leftow's book in the *Times Literary Supplement*, July 14, 2013.
Suppose that we take Anselmian Theism to consist in the following two claims: (a) there is a being than which none greater can be conceived; and (b) it is knowable on purely—solely, entirely—a priori grounds that there is a being than... more
Suppose that we take Anselmian Theism to consist in the following two claims: (a) there is a being than which none greater can be conceived; and (b) it is knowable on purely—solely, entirely—a priori grounds that there is a being than which none greater can be conceived. A key question in the assessment of Anselmian Theism concerns the interpretation of the expression 'being than which none greater can be conceived'. In particular, a question that is suggested by some of the recent literature in this topic is whether we should interpret this expression in terms of perfect— ideal—excellence, or whether we should interpret it in terms of maximal—maximal possible—excellence. In this paper, I set out to examine the notions of particular excellence, overall excellence, perfect excellence, and maximal excellence. I argue that, when we get clear about these notions, we see that Anselmian Theism gains traction by conflating notions that ought to be carefully distinguished; and we also see that there are grounds for thinking that a careful separation of notions that ought to be distinguished casts serious doubt on claim (b), i.e. on the second of the two claims that is constitutive of Anselmian Theism. There is also an appendix to my paper, in which I examine the recent defence of Anselmian Theism in Nagasawa (2008). Here, I argue that Nagasawa's defence of Anselmian Theism is undermined by the conflation identified in the main body of my paper.
Research Interests:
ABSTRACT We criticise Shepard's notions of “invariance” and “universality,” and the incorporation of Shepard's work on inference into the general framework of his paper. We then criticise Tenenbaum and Griffith's account of Shepard... more
ABSTRACT We criticise Shepard's notions of “invariance” and “universality,” and the incorporation of Shepard's work on inference into the general framework of his paper. We then criticise Tenenbaum and Griffith's account of Shepard (1987b), including the attributed likelihood function, and the assumption of “weak sampling.” Finally, we endorse Barlow's suggestion that minimum message length (MML) theory has useful things to say about the Bayesian inference problems discussed by Shepard and Tenenbaum and Griffiths. [Barlow; Shepard; Tenenbaum & Griffiths]
The advent of formal definitions of the simplicity of a theory has important implications for model selection. But what is the best way to define simplicity? Forster and Sober ((1994)) advocate the use of Akaike's Information Criterion... more
The advent of formal definitions of the simplicity of a theory has important implications for model selection. But what is the best way to define simplicity? Forster and Sober ((1994)) advocate the use of Akaike's Information Criterion (AIC), a non-Bayesian formalisation of the notion of simplicity. This forms an important part of their wider attack on Bayesianism in the philosophy of science. We defend a Bayesian alternative: the simplicity of a theory is to be characterised in terms of Wallace's Minimum Message Length (MML). We show that AIC is inadequate for many statistical problems where MML performs well. Whereas MML is always defined, AIC can be undefined. Whereas MML is not known ever to be statistically inconsistent, AIC can be. Even when defined and consistent, AIC performs worse than MML on small sample sizes. MML is statistically invariant under 1-to-1 re- parametrisation, thus avoiding a common criticism of Bayesian approaches.
Research Interests:
Oderberg (2002b) is a series of criticisms of Oppy (2002). Oppy (2002) is, in turn, a series of comments on Oderberg (2002a). Oderberg (2002b) claims that he “can scarcely find one thing that [Oppy (2002)] says about [Oderberg (2002a)] …... more
Oderberg (2002b) is a series of criticisms of Oppy (2002). Oppy (2002) is, in turn, a series of comments on Oderberg (2002a). Oderberg (2002b) claims that he “can scarcely find one thing that [Oppy (2002)] says about [Oderberg (2002a)] … that is not wrong or confused” (351). Restricting myself to the things that I said in Oppy (2002) about which Oderberg (2002b) passes comment, I propose to argue that, in quite large measure, it is actually Oderberg’s criticisms of my comments that are either wrong or confused. Moreover—and more importantly—I shall discuss some points of independent philosophical interest as I go along. (Perhaps it is worth noting that Oppy (2002) and Oderberg (2002b) were both prepared in great haste. This may well account for some of the errors that occur on each side.)
In some recent articles, Professor William Craig (1986) (1990) (1992) has argued that critiques of kalam cosmological arguments by Paul Davies, Stephen Hawking, and Adolf Grunbaum are superficial, ill-conceived, and based on... more
In some recent articles, Professor William Craig (1986) (1990) (1992) has argued that critiques of kalam cosmological arguments by Paul Davies, Stephen Hawking, and Adolf Grunbaum are superficial, ill-conceived, and based on misunderstanding. These judgements seem to me to be unfair. While I concede that some of the discussion of Davies and Hawking is not philosophically sophisticated, it seems to me that Davies, Hawking and Grunbaum do raise serious difficulties for the view that kalam cosmological arguments are rationally compelling pieces of natural theology. Of course, this is not to say that Davies, Hawking and Grunbaum offer compelling reasons for Craig to give up his belief that kalam cosmological arguments are sound -- but it is important to see that this is an entirely separate issue. At several points in his critiques, Craig makes things easy for himself by supposing that Davies, Hawking and Grunbaum must demonstrate that he -- Craig -- ought to give up his belief in the soundness of the argument; when, in fact, all that Davies, Hawking and Grunbaum need to show is that there is no good, non-question-begging, reason for them to be persuaded that the arguments which Craig offers are sound. What is at issue is a choice between two quite different kinds of models of the origins of the universe; if it turns out that there are no suitably independent reasons for preferring Craig’s favoured theistic model, then there is sufficient justification for those who wish to pursue alternatives.
Craig (1979) presents and defends several different kalam cosmological arguments. The core of each of these arguments is the following ur–argument: 1. The universe began to exist. 2. Whatever begins to exist has a cause of its existence.... more
Craig (1979) presents and defends several different kalam cosmological arguments. The core of each of these arguments is the following ur–argument:

1. The universe began to exist.
2. Whatever begins to exist has a cause of its existence.
3. (Hence) The universe has a cause of its existence.
4. (Hence) God exists.

What distinguishes between the different kalam arguments which Craig defends is the sub–argument which is given on behalf of the first premise in this ur–argument.
One of the sub–arguments appeals to a posteriori scientific considerations: there is empirical evidence that the universe has only existed for a finite amount of time, and hence that it began to exist a finite number of years ago. The other two sub–arguments appeal to a priori philosophical considerations: there are broadly logical arguments which establish the conclusion that the universe could only have existed for a finite amount of time, and hence that it must have begun to exist no more than some finite number of years ago. According to one of these sub–arguments, it is impossible for there to be completed infinities: since the past would be a completed infinity if it were infinite, it follows that the past must be finite. According to the other of these sub–arguments, it is impossible for there to be completed infinities formed by successive addition: since the past would be a completed infinity formed by successive addition if it were infinite, it follows that the past must be finite.

In this paper, I wish to look at the second of the two a priori arguments. Set out in full, this argument might be represented as follows:

1. It is not possible for a series formed by successive addition to be both infinite and completed.
2. The temporal series of (past) events is formed by successive addition.
3. The temporal series of past events is completed (by the present).
4. (Hence) It is not possible for the temporal series of past events to be infinite.
5. (Hence) The temporal series of past events is finite.
6. (Hence) The universe began to exist.
7. Whatever begins to exist has a cause of its existence.
8. (Hence) The universe has a cause of its existence.
9. (Hence) God exists.

There are many criticisms which a non–theist might choose to make of this argument. However, in this paper, I shall be focussing my attention on the second premise, i.e. on the claim that the temporal series of events is formed by successive addition. I shall argue that non–theists can have good reason to refuse to accept this premise—and hence that non–theists can have good reason to reject arguments which make use of this premise. I shall not be claiming that this is the strongest—or most important—objection which non–theists can make to the argument; however, towards the end of the paper, I shall—in effect—explore the suggestion that reasons for refusing to accept this premise can be generalised to reasons for refusing to accept the conjunction of the two premises in the ur–argument with which we began.
Let ‘reality’ refer to the largest whole every eligible part of which is connected to our eligible part of reality under a privileged external relation (think: ‘a is causally related to b’). We suppose that this privileged external... more
Let ‘reality’ refer to the largest whole every eligible part of which is connected to our eligible part of reality under a privileged external relation (think: ‘a is causally related to b’). We suppose that this privileged external relation has an analogue that is directed, so that some eligible parts of reality are anterior to other eligible parts of reality under the analogue of the privileged relation (think: ‘a is causally anterior to b’). We do not suppose that this analogue relation is total: we allow that there are non-overlapping eligible parts of reality such that neither is anterior to the other under the analogue relation. However, as noted, we do suppose that any two non-overlapping eligible parts of reality are connected by a chain of non-overlapping eligible parts of reality under the relation itself. Moreover, we suppose that each eligible part of reality is such that none of its parts is anterior to any other of its parts under the analogue relation, and also such that there are no other eligible parts of reality to which it fails to be connected under the relation itself.

To illustrate this rather abstract account of reality—and to exhibit a justification for its complexity—we consider a concrete example. Pretend that reality is exhausted by an instance of a standard general relativistic universe that originates in an initial singularity. Eligible parts of reality will be maximal sub-parts all of whose parts have only space-like connections to each other. These eligible parts will stand in causal—light-like and time-like—relations to one another. If, for example, the initial singularity is a singular surface, it may be that there are non-overlapping eligible parts of reality that are causally related to one another only in virtue of the fact that they trace back to non-overlapping regions of that initial singular surface. (Among the reasons why we can only pretend that reality is exhausted by an instance of a standard general relativistic universe that originates in an initial singularity, perhaps the most important is that we should not immediately rule out the possibility that reality has parts ‘on the other side of the initial singularity’ from which our observable universe has arisen. We suppose only that, if reality does have parts ‘on the other side of the initial singularity’ from which our observable universe has arisen, then there are least analogues of causation and space-like relation that determine the extent of reality.)
There are many hypotheses that one can frame about the overall ‘shape’ taken by reality under our privileged external relation. We begin by distinguishing the following two very general competing hypotheses:

Infinite Regress: Under the external relation, each eligible part of reality which is posterior to some non-overlapping eligible part of reality belongs to a chain of non-overlapping eligible parts of reality which satisfies the following condition: for each member of the chain, there is an anterior (and not posterior) member of the chain that does not overlap with any other members of the chain.

Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any eligible part of reality.

In the subsequent discussion, we simply—though perhaps improperly—ignore all of the many competing hypotheses that one might frame about the general shape of reality under the privileged external relation.

The Initial Boundary hypothesis divides into two competing sub-hypotheses, depending upon the modal status of the initial boundary:

Contingent Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any non-overlapping eligible part of reality, and that smallest eligible part of reality exists contingently.

Necessary Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any non-overlapping eligible part of reality, and that smallest eligible part of reality exists of necessity.

The Initial Boundary hypothesis also divides into two competing sub-hypotheses depending upon the ontological status of the initial boundary:

Immanent Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any non-overlapping eligible part of reality, and that smallest eligible part of reality is continuous in nature with the rest of reality

Transcendent Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any non-overlapping eligible part of reality, and that smallest eligible part of reality is radically different in nature from the rest of reality.

And the Initial Boundary hypothesis divides into two competing sub-hypotheses depending upon the psychological status of the initial boundary:

Impersonal Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any non-overlapping eligible part of reality, and that smallest eligible part of reality has no personal—mental, psychological—properties.

Personal Initial Boundary: Under the external relation, there is a smallest eligible part of reality that is anterior to every other non-overlapping eligible part of reality and not posterior to any non-overlapping eligible part of reality, and that smallest eligible part of reality has personal—mental, psychological—properties.

Given only the distinctions that we have drawn to this point, we now have eight competing versions of the Initial Boundary hypothesis, and we also have the competing Infinite Regress hypothesis. Are there reasons to prefer one of these nine hypotheses above all of the others?

Theists suppose that the answer to this question is affirmative: we should prefer Necessary Transcendent Personal Initial Boundary above the other eight hypotheses. On the other hand, it is not at all clear that naturalists suppose that the answer to the question is affirmative: it may well be that naturalists suppose that we have no reasons for preferring any one of Infinite Regress, Necessary Immanent Impersonal Initial Boundary, and Contingent Immanent Impersonal Initial Boundary to the other two hypotheses in this group. However, naturalists will hold that at least one of Infinite Regress, Necessary Immanent Impersonal Initial Boundary, and Contingent Immanent Impersonal Initial Boundary is preferable to all of the six remaining alternatives; and if naturalists are undecided between two or more of the hypotheses that are congenial to naturalism, then they will hold that all of the hypotheses between which they are undecided are preferable to the remaining six alternatives.

In the ensuing discussion, we shall restrict our attention to the dispute between theists and naturalists about the standing of the four hypotheses singled out in the preceding paragraph. As before, we simply—though perhaps improperly—ignore all other disputes concerning the relative standing of the nine hypotheses that we have identified, and concerning the many hypotheses that we have not even attempted to frame. (It is, for example, an interesting question what pantheists should say about the relative standing of our nine hypotheses. But that is not a question properly taken up in the present context.)

One way of reading William Lane Craig’s voluminous writings on the kalām cosmological argument is as the construction of a case for the superiority of Necessary Initial Boundary to both Infinite Regress and Contingent Initial Boundary. While Craig’s writings focussed solely on the kalām cosmological argument offer no means of deciding between Necessary Immanent Impersonal Initial Boundary and Necessary Transcendent Personal Initial Boundary, the success of the case developed in those writings would remove Infinite Regress and Contingent Immanent Impersonal Initial Boundary from the contest, or, at any rate, would provide some non-negligible but defeasible reason to prefer Necessary Transcendent Personal Initial Boundary to Infinite Regress and Contingent Immanent Impersonal Initial Boundary. Consequently, naturalists have good reason to think carefully about the case that Craig develops: if his case is compelling, then naturalists are obliged to adopt Necessary Immanent Impersonal Initial Boundary. We shall start with the case that Craig develops against Infinite Regress.
In Oppy (2001), I argued that non-theists can reasonably reject the claim—which plays a key role in some versions of the kalam cosmological argument—that the temporal series of (past) events is formed by successive addition. David... more
In Oppy (2001), I argued that non-theists can reasonably reject the claim—which plays a key role in some versions of the kalam cosmological argument—that the temporal series of (past) events is formed by successive addition. David Oderberg (2001) makes some interesting criticisms of the arguments that I gave in Oppy (2001). As I said in the concluding footnote to that previous paper of mine, I do not think that there is much that I need to change or retract in what I said there. However, Oderberg’s criticisms do give me the opportunity to expand on some points which only appeared in very compressed form in that earlier paper. Moreover, there are some points where Oderberg misunderstands or misrepresents what I said previously; at the very least, it will be worthwhile to try to get clearer about exactly where it is that we have major disagreements. In what follows, I will respond—in turn—to the four criticisms that Oderberg makes of Oppy (2001).
I hold that the considerations adduced in kalam cosmological arguments do not embody reasons for reflective atheists and agnostics to embrace the conclusion of those arguments, viz. that the universe had a cause of its existence. I do not... more
I hold that the considerations adduced in kalam cosmological arguments do not embody reasons for reflective atheists and agnostics to embrace the conclusion of those arguments, viz. that the universe had a cause of its existence. I do not claim to be able to show that reflective theists could not reasonably believe that those arguments are sound; indeed, I am prepared to concede that it is epistemically possible that the arguments proceed validly from true premises. However, I am prepared to make the same concession about the following argument: Either 2+2=5 or God exists; 2+2≠5; therefore God exists. But nobody could think that this argument deserves to be called a proof of its conclusion (even if it is sound). Of course, this latter argument is obviously circular: (almost) no-one who was not antecedently persuaded of the truth of the conclusion would (have reason to) believe the first premise. But this fact does not entail that admittedly non-circular arguments, such as the kalam cosmological arguments, cannot fail to be equally dialectically ineffective. And, indeed, that is the view which I wish to defend: there is not the slightest reason to think that kalam cosmological arguments should be dialectically effective against reasonable and reflective opponents.

I take it that proponents of kalam cosmological arguments wish to maintain that the arguments are dialectically effective -- i.e. that reasonable and reflective opponents ought to be persuaded by them. Moreover, I take it that in order to refute these arguments, it is sufficient to show that one can reasonably refuse to be persuaded by them. Of course, one might also seek to show that, in fact, one ought reasonably disbelieve one or more of the premises or the conclusion of the argument -- but that would be a much more difficult undertaking, and one in which I have no interest. (Perhaps John Mackie wanted to argue for the stronger view; in that case, all I wish to claim is that his arguments can be adapted to substantiate the weaker thesis.)
At the outset, I take it that there is a prime facie presumption that there can be reasonable and reflective atheists, agnostics, and theists. It is epistemically possible that one or more of these positions should turn out to be logically inconsistent, or ad hoc with respect to uncontroversial evidence, or clearly deficient in explanatory power, etc. However, it seems clear that, at the beginning of the dialectic between theists and their opponents, all sides should concede the (prima facie) reasonableness of their opponents views. For, if this is not conceded, then there is really no way of proceeding with the dialectic. Moreover, I assume that the aim of the debate is to bring one’s opponent to see that, by her own lights, she ought to accept the conclusion for which one is arguing. In other words: in order to win, one needs to show that, by her own lights, a reasonable and reflective opponent will improve her view by adopting the conclusion(s) for which one is arguing.

Now, I do not know whether Professor Craig wishes to defend the claim that kalam cosmological arguments are dialectically effective. However, while some of his comments suggest that he only wishes to defend the view that theists can reasonably believe that the arguments are sound, I suspect that he does wish to defend this claim. At any rate, I shall proceed to respond to his objections under the assumption that it may be the case that he wishes to defend the claim that kalam cosmological arguments are dialectically effective. (Note that it would be pointless to say that all Craig wishes to do is to defend the view that kalam cosmological arguments are sound -- i.e. that he has no interest in questions of dialectical efficacy. For, of course, all parties to the debate about these arguments will agree that what they are really interested in is the truth of the conclusion of the arguments. And moreover, each will surely be entitled to believe -- and indeed will be obliged to believe -- that her own beliefs on this matter are true. The only interesting question is whether one can be shown that one has reason to change one’s beliefs. Note, too, that it would be equally pointless to suppose that what one wants to do is to produce arguments which serve to justify the beliefs which one is fact has. For how could the exhibition of logical relationships between propositions -- e.g. a demonstration that it would be inconsistent for one to reject the conclusion of a kalam cosmological argument, given that one accepts its premises -- serve to
show -- even to oneself -- that one is justified in believing that conclusion?)
William Lane Craig’s defence of kalam cosmological arguments has generated a large amount of discussion in the years since the publication of his book on that topic. John Mackie, Quentin Smith, Adolf Grunbaum and I are just some of the... more
William Lane Craig’s defence of kalam cosmological arguments has generated a large amount of discussion in the years since the publication of his book on that topic. John Mackie, Quentin Smith, Adolf Grunbaum and I are just some of the philosophers who have engaged in prolonged debates in the journals with Craig on issues which arise in connection with his defence of those arguments. Given the amount of discussion which has already taken place, one might be given to wonder whether there are any important points which remain to be made. Perhaps unsurprisingly, I think that it is clear that there are important issues which have not been taken up in the previous discussion; I propose to take up one such set of issues in this paper.

The main question I want to address here is this: Given that someone presents you with an argument for a given conclusion, under what conditions should the person who presents the argument be prepared to agree that the argument is unsuccessful? Ultimately, my aim is to argue that, in light of some of the criticisms which have been made in the literature, Craig ought to be prepared to concede that what he calls “the kalam cosmological argument” is plainly an unsuccessful argument. However, much of the paper is taken up with consideration of preliminary matters which need to be addressed in order to provide a satisfactory answer to the main question.

In a full discussion of this main question, one would need to provide: (a) an account of rationality and rational belief revision; (b) an account of arguments; (c) an account of rational argumentation among rational agents; and (d) a discussion of the problems which arise as a result of the fact that we are not perfectly rational agents. I shall say a little about each of these topics, though I recognise that what I have to say will not amount to a thorough discussion of any of them. I shall then go on to apply the results from this part of the paper to the on-going debate about kalam cosmological arguments. Finally, I shall discuss two cases—one from Craig’s criticisms of Grunbaum (2000) and one from Craig’s criticisms of Oppy (1995b)—which it seems to me that those criticisms turn on misunderstandings of the proper framework within which rational argument takes place.
In "Professor Mackie And The Kalam Cosmological Argument" (Religious Studies, 1984, Vol.20, pp.367-375), Professor William Lane Craig undertakes to demonstrate that J. L. Mackie's analysis of the kalam cosmological argument in The Miracle... more
In "Professor Mackie And The Kalam Cosmological Argument" (Religious Studies, 1984, Vol.20, pp.367-375), Professor William Lane Craig undertakes to demonstrate that J. L. Mackie's analysis of the kalam cosmological argument in The Miracle Of Theism (Oxford University Press, Oxford, 1982) is "superficial", and that Mackie "has failed to provide any compelling or even intuitively appealing objection against the argument". (p.367) I disagree with Craig's judgement; for it seems to me that the considerations which Mackie advances do serve to refute the kalam cosmological argument. Consequently, the purpose of this paper is to reply to Craig's criticisms on Mackie's behalf.

This paper has three parts. In the first part, I outline the kalam argument, and introduce the objections which Mackie makes to it. In the second part, I present the replies which Craig makes to Mackie's objections. Finally, in the third part, I explain why I think that Craig's replies are unsuccessful.
In “Reply To Smith: On The Finitude Of The Past”, Professor William Craig writes: "I reiterate that Smith has yet to deal with my strongest arguments in favour of the impossibility of the existence of an actual infinite, those based on... more
In “Reply To Smith: On The Finitude Of The Past”, Professor William Craig writes:

"I reiterate that Smith has yet to deal with my strongest arguments in favour of the impossibility of the existence of an actual infinite, those based on inverse operations performed with transfinite numbers."

I think that this claim is mistaken; for: (i) there is no problem about allowing the inverse operations in question -- subtraction, division, extracting roots, etc. -- into transfinite ordinal arithmetic; and (ii) there is no problem about the exclusion of these operations from transfinite cardinal arithmetic. I shall take up these points in turn.
William Lane Craig's 'Reflections on Uncaused Beginnings' 1-hereafter 'Craig (2010)'-is a sustained critique of my 'Uncaused Beginnings' 2-hereafter 'Oppy (2010)'. I argue that the central arguments of Oppy (2010) survive the critique of... more
William Lane Craig's 'Reflections on Uncaused Beginnings' 1-hereafter 'Craig (2010)'-is a sustained critique of my 'Uncaused Beginnings' 2-hereafter 'Oppy (2010)'. I argue that the central arguments of Oppy (2010) survive the critique of Craig (2010) unscathed. When we make a fair and accurate comparison of naturalist and theist claims about global causal reality, we see that considerations about causation and the shape of causal reality do not decide between naturalism and theism. Moreover, the Edwards/Prior/Craig objection does not rule out the view that there is an initial global causal state involving none but contingently existing entities.
Davis (2008) is a critique of Oppy (1993). I think that Davis’ critique fails to touch the argument of my early paper, because it is based on serious misunderstanding of that argument. Perhaps this misunderstanding is my fault: perhaps I... more
Davis (2008) is a critique of Oppy (1993). I think that Davis’ critique fails to touch the argument of my early paper, because it is based on serious misunderstanding of that argument. Perhaps this misunderstanding is my fault: perhaps I failed to write with sufficient clarity. In any case, my aim here is to give a clear account of the argument of Oppy (1993), and then to explain why there is nothing in Davis (2008) that threatens this argument. I shall also discuss criticisms that Davis makes of some of my other writings; there are some things that he says about my account of begging the question in Oppy (1995) that are worthy of further investigation.
Microphysicalists hold that microphysics provides a subvenient basis upon which all else supervenes: fix the microphysical properties of a world like ours, and one fixes all the properties of a world like ours. I would like to endorse... more
Microphysicalists hold that microphysics provides a subvenient basis upon which all else supervenes: fix the microphysical properties of a world like ours, and one fixes all the properties of a world like ours. I would like to endorse this thesis of microphysical supervenience.

David Lewis has defended the philosophical tenability of a different thesis, the principle of Humean Supervenience, according to which the spatiotemporal arrangement of local qualities provides an irredundant subvenient basis upon which all else supervenes. Under the natural assumption that the fundamental point properties on which all else supervenes are microphysical, the doctrine of Humean supervenience entails that the spatiotemporal arrangement of local microphysical qualities provides an irredundant subvenient basis upon which all else supervenes. I do not want to accept this natural consequence of the principle of Humean Supervenience. Moreover, I do not think that microphysicalists should look with too much fondness on Lewis’ defence of the philosophical tenability of the principle of Humean Supervenience. The best reasons for looking askance at the principle of Humean Supervenience are also good reasons for thinking that the principle may not be ‘philosophically tenable’—or so I shall argue.

The plan of the paper is as follows. I begin, in section I, by rehearsing Lewis’ formulation of the principle of Humean Supervenience, and provide a brief discussion of some objections to it which he considers and sets aside. In sections II and III, I discuss some further objections to the principle of Humean Supervenience, including some which I wish to endorse. As far as I know, these objections are novel—i.e. they have not hitherto been discussed in the literature. In section IV, I take up the question of the bearing of the objections which I endorse on the claim that the principle of Humean Supervenience is ‘philosophically tenable’. Finally, in section V, I close with some considerations about the proper characterisation of the doctrine of microphysicalism.
Neale [2001] is an expanded version of a paper that appeared in Mind in 1995. In Oppy [1997], I claimed to find fault with various aspects of that original paper. Neale and Dever [1997] then provided a vigorous defence of Neale’s original... more
Neale [2001] is an expanded version of a paper that appeared in Mind in 1995. In Oppy [1997], I claimed to find fault with various aspects of that original paper. Neale and Dever [1997] then provided a vigorous defence of Neale’s original paper against my complaints. In the preface to the book, Neale [2001:x] says that he “stands by the main points” of both the original paper and the reply to Oppy [1997] My aim in the present paper is to assess the extent to which Neale is right to claim that he has now succeeded in meeting the objections that I brought against Neale [1995].
There are many possible models for the causal shape of reality. Amongst the simple models to be considered—even if only to be subsequently rejected as models of real possibilities—we should certainly mention: REGRESS, CIRCLE, NECESSARY... more
There are many possible models for the causal shape of reality. Amongst the simple models to be considered—even if only to be subsequently rejected as models of real possibilities—we should certainly mention: REGRESS, CIRCLE, NECESSARY INITIAL STATE, and CONTINGENT INITIAL STATE.1 Each of these simple models admits of both theistic and naturalistic interpretations. It is widely recognised that in several of these cases—REGRESS, CIRCLE, CONTINGENT INITIAL STATE—naturalism would be preferable to theism: if reality had the causal shape in question, then there would be good reason to accept naturalism and to reject theism.

I think that naturalism is preferable to theism even in the case of NECESSARY INITIAL STATE. Consequently, I think that considerations about the causal shape of reality provide grounds for naturalism: for I judge that, on any of the most plausible causal shapes that might be taken by reality, naturalism is more plausible than theism. Of course, these judgments of mine are highly controversial; however, what seems less controversial is that the case for naturalism is strengthened if the models on which naturalism is clearly preferable to theism remain in play.

In this paper, I shall be examining an argument for the claim that CONTINGENT INITIAL STATE is not a model that should be kept in play. This argument runs as follows.

1. If it is possible for reality to have a contingent initial state under the causal relation—i.e. it is possible for reality to have a contingent initial state that has no cause—then it is possible for other (non-overlapping) parts of reality to have no cause. (Premise)

2. It is not possible for other (non-overlapping) parts of reality to have no cause. (Premise)

3. (Hence) It is not possible for reality to have a contingent initial state that has no cause. (From 1, 2)

This argument is my reconstruction of an argument introduced in Craig (1985: 371n3)—repeated in Craig (1986:167-8), Craig (1991), and Craig (1993:7)—which Craig claims has antecedents in the work of Edwards (1754) and Prior (1962). While it may be that this is not the best possible reconstruction of Craig’s argument, that will not matter for present purposes, since my main purpose here is to investigate the first premise of this argument, and, in particular, to explore potential naturalistic objections to it. I take it that Craig clearly commits himself to this first premise, and that Edwards and Prior do likewise; further investigation of exactly who argues what and how can be left to some other occasion.
In “The Philosophical Significance of Godel’s Slingshot”1, Stephen Neale claims to answer all technical questions about Godel’s Slingshot, and also to highlight the philosophical significance of that argument. Moreover, he (at least... more
In “The Philosophical Significance of Godel’s Slingshot”1, Stephen Neale claims to answer all technical questions about Godel’s Slingshot, and also to highlight the philosophical significance of that argument. Moreover, he (at least implicitly) defends the view that Godel’s Slingshot is philosophically important -- i.e. he (at least implicitly) suggests that Godel’s Slingshot has important philosophical consequences for theories of facts and for referential treatments of definite descriptions. Finally, he explicitly claims: (i) that Godel’s Slingshot highlights unpleasant consequences of referential treatments of descriptions (p.817); (ii) that Godel’s Slingshot poses genuine philosophical questions about the logical and philosophical consequences of rejecting Russell’s Theory of Descriptions (p.765); (iii) that Godel’s Slingshot forces friends of facts to take a definite position on the semantics of definite descriptions (pp.764, 817); (iv) that, on plausible assumptions, Godel’s Slingshot forces anyone who is giving a logic for causal statements to endorse Russell’s Theory of Descriptions (p.814); and (v) that Godel’s Slingshot has important consequences for any theory of facts because of the consequences which it has for the sentential operator FIC -- “The fact that () = the fact that []”. (p.815)

I have doubts about all of this. Most importantly, I am sceptical of the philosophical significance of Godel’s Slingshot (and of Slingshot arguments in general). In particular, I do not believe that Godel’s Slingshot has any interesting and important philosophical consequences for theories of facts or for referential treatments of definite descriptions. More generally, I do not believe that any Slingshot arguments have interesting and important philosophical consequences for theories of facts or for referential treatments of definite descriptions. Friends of facts and referential treatments of definite descriptions can, and should, proceed with the construction of their theories, blithely ignoring the many Slingshots which now litter the landscape. (Of course, there may be other considerations which should give such theorists pause -- but those are other considerations.)

What would it take for a Slingshot argument to have interesting and important philosophical consequences for theories of facts or referential treatments of definite descriptions? I assume that (i) such an argument would need to impose a further constraint upon those theories, beyond those constraints which may be deduced simply from consideration of general theoretical desiderata such as consistency, completeness, simplicity, accommodation of data, and so on; (ii) such an argument should impose a constraint for which it is prima facie plausible that there are interesting theories of the kind in question which fail to meet the constraint -- i.e. it should not be obvious that all of the theories of the kind in question will meet the constraint, regardless of how they are developed; and (iii) such an argument should impose a constraint which operates independently of theoretical assumptions which would, by themselves, suffice to undermine or demolish the theories in question. That is, I assume that it would make sense now to claim that Godel’s Slingshot has interesting and important philosophical consequences for theories of facts and definite descriptions provided that Godel’s Slingshot imposes some particular constraint on those kinds of views for which it is true both (a) that it is at least prima facie plausible that there are interesting versions of the views in question which fail to meet the constraint; and (b) that the constraint which Godel’s Slingshot does impose isn’t due either to independent theoretical assumptions or to general theoretical considerations which would have been taken into account anyway, regardless of the development of the Slingshot.

The claim that I wish to defend in this note is that Godel’s Slingshot fails to meet these constraints. While not wishing to deny that the discussion of Godel’s Slingshot has technical and historical significance, I want to insist that Godel’s Slingshot does not impose any philosophically interesting constraints upon theories of facts and referential treatments of definite descriptions. On the one hand, there are many views which one might have about facts but -- as far as I can tell -- it is almost impossible to construct a prima facie plausible theory of facts which falls foul of Godel’s Slingshot. On the other hand, while there are, perhaps, not many plausible referential treatments of definite descriptions, it seems to me that assumptions involved in the construction of Godel’s Slingshot must be denied by any such referential treatment of definite descriptions. In particular, it seems to me that any referential treatment of definite descriptions will require rejection of semantic innocence (or direct reference, or both) -- i.e. rejection of the claim that the semantic content of vocabulary does not vary as the non-quotational sentential frames in which that vocabulary is embedded varies (or rejection of the claim that the semantic content of (say) singular terms in extensional contexts is just the objects (if any) to which those singular terms refer, or both) -- and yet, the assumption of semantic innocence (or direct reference, or both) is implicitly built into the construction of Slingshot arguments.

In order to bolster the arguments which I give, I shall also try my hand at sketching a kind of theory which is fact-friendly, and in which definite descriptions can be treated as singular terms. The aim will be to provide enough detail to suggest (i) that it isn’t obvious on independent grounds that no theory of this kind can succeed; and yet (ii) that it is quite clear that this theory will not fall to Godel’s Slingshot. (Of course, even if the kind of theory which I sketch only meets the second of these criteria, it will still serve to illustrate the main point which I wish to make: Godel’s Slingshot gets no credit if this kind of theory is only independently defeated.) I shall then close with some comments about statements about the identity of facts.
Many contemporary philosophers claim to be ‘physicalists’; many of these philosophers take themselves to be heirs to Greek atomism and seventeenth century materialism. Many other contemporary philosophers hold that ‘physicalism’ either... more
Many contemporary philosophers claim to be ‘physicalists’; many of these philosophers take themselves to be heirs to Greek atomism and seventeenth century materialism. Many other contemporary philosophers hold that ‘physicalism’ either admits of no intelligible formulation, or else is hopelessly vulgar and undeserving of serious philosophical attention. Before we can arbitrate this apparent dispute, we need to get clearer about what ‘physicalism’ might mean. In the circumstances, it would not be surprising to learn that those who claim to be ‘physicalists’ defend a far more modest doctrine than those ‘physicalist’ views which others allege to be hopelessly vulgar and undeserving of serious philosophical attention.

The plan of the discussion is as follows. In the first section of the paper, I consider some initial difficulties which arise in the formulation of a statement of what it is that ‘physicalists’ believe. These difficulties concern the range of entities which are quantified over—objects or properties?—and ways of handling mathematics, logic, and the like. In the second section of the paper, questions about the status of ‘physicalism’ are considered: should it be taken to be necessary, and/or analytic, and/or a priori; and should it be taken to be telling physicists how to conduct their investigations? This section includes some discussion of microphysicalism, and some discussion of the doctrine of Humean supervenience. The third section of the paper is devoted to consideration of issues concerning reduction and elimination: what should ‘physicalists’ say about everything which lies outside of physics, or their favoured part of physics, or the physical sciences more broadly construed? Here, I argue that the most promising form of ‘physicalism’ provides for non–analytic reduction of the non–physical to the physical. In the fourth section of the paper, a range of supervenience theses is canvassed. One aim is to show that there are no decent prospects for ‘non–reductive physicalism’. Another aim is to exhibit a new supervenience claim which, I argue, succeeds in capturing what it is that ‘physicalists’ should want to say about the relation between the physical and the non–physical. The fifth section of the paper takes up some questions about the importance of physicalism as thus characterised. I shall suggest that physicalism is a relatively anodyne doctrine, without much importance for anything other than fundamental metaphysics. In the sixth section of the paper, I turn to a brief examination of reasons for supposing that non–analytic reductive physicalism is true. Finally, I conclude with some brief remarks about the spirit in which this investigation has been conducted.
This paper presents an attempt to integrate theories of causal processes - of the kind developed by Wesley Salmon and Phil Dowe - into a theory of causal models using Bayesian networks. We suggest that arcs in causal models must... more
This paper presents an attempt to integrate theories of causal processes - of the kind developed by Wesley Salmon and Phil Dowe - into a theory of causal models using Bayesian networks. We suggest that arcs in causal models must correspond to possible causal processes. Moreover, we suggest that when processes are ren- dered physically impossible by what occurs on distinct paths, the original model must be restricted by removing the relevant arc. These two techniques su ce to explain cases of late preëmption and other cases that have proved problematic for causal models.
In this paper, I examine—but do not endorse—a position which I shall call ‘thorough–going resemblance nominalism’. This position has two distinct aspects. On the one hand, it is a thorough–going ‘nominalism’, in one familiar sense of that... more
In this paper, I examine—but do not endorse—a position which I shall call ‘thorough–going resemblance nominalism’. This position has two distinct aspects. On the one hand, it is a thorough–going ‘nominalism’, in one familiar sense of that term: it is a view which holds that there is nothing but concrete individuals. On the other hand, it is a resemblance ‘nominalism’, in the other familiar sense of that term: it is a view which holds that pretty much all predication can be analysed or explained in terms of a single primitive resemblance predicate.

Although I do not propose to endorse thorough–going resemblance nominalism, I hold that it is worthy of examination. Some people may take the view that each of the separate aspects of the view has already been thoroughly discredited: there are devastating objections to be lodged against thorough–going nominalism, and against resemblance nominalism. However, recent work by Lewis (1983) (1986) (1991), and others, suggests—to me at any rate—that these familiar objections may not be quite so devastating as they appear. More about this anon.

The plan of the paper is as follows. I begin with some introductory remarks about different kinds of nominalist projects and the motivations which one might have for adopting them. I then turn to a discussion of the primitive resources of thorough–going resemblance nominalism, and of the benefits which might accrue to such a view. Finally, I consider some of the difficulties which stand in the way of the development of any kind of thorough–going resemblance nominalism.
In [3], Quentin Smith claims that ‘the Hartle-Hawking cosmology’ is inconsistent with classical theism in a way which redounds to the discredit of classical theism; and, moreover, that the truth of ‘the Hartle-Hawking cosmology’ would... more
In [3], Quentin Smith claims that ‘the Hartle-Hawking cosmology’ is inconsistent with classical theism in a way which redounds to the discredit of classical theism; and, moreover, that the truth of ‘the Hartle-Hawking cosmology’ would undermine reasoned belief in any other varieties of theism which hold that the universe is created.

I don’t think that Smith manages to substantiate these prima facie implausible claims. In particular, I do not think that he manages to provide an intelligible account of what he takes to be the crucial consequence of ‘the Hartle-Hawking cosmology’ — viz. that there is a probability strictly less than 100% and strictly greater than 0% of a universe like ours coming into existence ex nihilo. The main purpose of this paper is to explain why it seems to me that this claim is simply incoherent.

There are other points at which Smith’s arguments could be attacked. For example, Markosian (1995) objects to Smith’s assumption, that the claim that God wills that a universe with a given initial state exists entails the claim that the objective probability of such a universe coming into existence is 100% (and similarly to Smith’s assumption, that the claim that God wills that Hawking’s wave function law obtain entails that the objective probability of our universe coming into existence is (say) 95%). Markosian’s objection relies on the idea that one could evaluate the consequents of these conditionals at times earlier than the times at which the willings in the antecedents occur. However — as I shall go on to argue — it seems to me to be quite doubtful that one should allow that there are times of the kind which Markosian’s objection requires. In any case, rather than object to the entailments which Smith’s argument requires, I shall focus my attention on the coherence of the objective probabilities which Smith invokes.

One might also worry about the way in which Smith formulates the problem. He defines ‘classical theism’ to be ‘the theory that there necessarily exists a disembodied person who is necessarily omniscient, omnipotent and omnibenevolent and necessarily the cause of whatever universe there is’ (p.238). His main claim is that ‘the Hartle-Hawking cosmology’ is inconsistent with ‘classical theism’ thus construed. But this claim may be quite uninteresting: for it may be that ‘classical theism’ is internally inconsistent; and, even if it is not internally inconsistent, it may be that ‘classical theism’ is trivially inconsistent with ‘the Hartle-Hawking cosmology’. (For example, a reading of Smith’s definition which has it that there is only one initial state of the universe which God could create -- all others being inconsistent with his necessary omnibenevolence -- is trivially inconsistent with any theory which assigns non-zero probabilities to non-actual initial states, and may be internally inconsistent as well.) Smith’s claim might turn out to be comparatively uninteresting for other reasons. Suppose that most theists would commit themselves to no more than the claim that there is a very good, very wise and very powerful person who created the universe — and suppose further that there are no compelling reasons for them to shift to stronger positions such as the one which Smith outlines. In that case, the important question will be whether this position is inconsistent with ‘the Hartle-Hawking cosmology’. (Hawking himself seems to claim that his cosmology makes the hypothesis of any kind of creator otiose; if Smith really means to defend Hawking, then surely it is this claim which he ought to be defending.) However, I don’t need to fuss about these details: if I am right to claim that Smith’s account of ‘the Hartle-Hawking cosmology’ is incoherent, then the further details of Smith’s argument can be safely ignored.
It is sometimes said that causal theories of reference -- when conjoined with plausible metaphysical principles -- entail that it is logically or metaphysically impossible for reference to be made to entirely future individuals. In... more
It is sometimes said that causal theories of reference -- when conjoined with plausible metaphysical principles -- entail that it is logically or metaphysically impossible for reference to be made to entirely future individuals. In particular, it is sometimes said that, because backwards causation is logically or metaphysically impossible, causal theories of reference entail that it is logically or metaphysically impossible for reference to be made to entirely future individuals. Moreover, it is then said that this claim about the logical or metaphysical impossibility of reference to entirely future individuals entails that four-dimensionalist or “tenseless” theories of time are mistaken.

The premise about backwards causation is controversial. Many philosophers -- including many who are sympathetic to four-dimensionalist theories of time -- do not believe that backwards causation is logically or metaphysically impossible. Some philosophers hold that backwards causation is merely nomically impossible; others hold that it is merely contingently absent from the world; and yet others hold that, in fact, there is lots of microscopic backwards causation in the world. However, very few philosophers think that there is any macroscopic backwards causation in the world. And, if there is no macroscopic backwards causation in the world, then -- if the argument from causal theories of reference is correct -- there is no reference to entirely future individuals. But that is enough to create problems for four-dimensionalist theories of time, since these theories do suppose, e.g., that “tenseless” quantifiers range over entirely future individuals. Since I think that it is plausible to suppose that there is no macroscopic backwards causation in the world, I maintain that this weakened argument still presents a prima facie challenge to four-dimensionalism about time.

In order to assess this challenge, we need to know more about the causal theories of reference which are invoked. If all that is intended is Kripke’s suggestion that the meaning of proper names is transmitted along causal communicative chains, then it is open to the objection that, for all Kripke says, it may be possible to effect non-causal naming baptisms, by using suitable descriptions to fix the referents of names. Even if one accepts a broadly Kripkean account of names, one can still allow that it is possible to use descriptions to fix the referents of names which refer to entirely future individuals. Clearly, then, the causal theories of reference which are invoked by the argument must impose a stronger constraint.

I conjecture that the intended argument goes like this: There is no macroscopic backwards causation. So there is no direct causal contact with future individuals. So no-one stands in directly causally mediated epistemic relations to entirely future individuals. So no-one has de re singular thoughts about entirely future individuals. But reference to entirely future individuals can only occur if there are de re singular thoughts about those individuals. So there is no reference to entirely future individuals. So four-dimensionalism about time is mistaken.
There are various kinds of questions that might be asked by those in search of ‘ultimate explanations’. Why is there anything at all? Why is there something rather than nothing? Why is there causal stuff? Why is there causal stuff rather... more
There are various kinds of questions that might be asked by those in search of ‘ultimate explanations’. Why is there anything at all? Why is there something rather than nothing? Why is there causal stuff? Why is there causal stuff rather than complete absence of causal stuff? Why is there causal stuff that behaves as it does? Why is there causal stuff that behaves as it does rather than causal stuff that behaves in other ways?

In this chapter, my focus will be on ‘ultimate causal explanations’ and ‘ultimate explanations of the natural world’—or, more exactly, on the relative merits of theistic and naturalistic ‘ultimate causal explanations’ and ‘ultimate explanations of the natural world’. If we suppose that there are non-causal things—abstracta and the like—then we will not suppose that this discussion exhausts what there is to say about the relative merits of theistic and naturalistic ‘ultimate explanations’. However, I leave discussion of the relative merits of theistic and naturalistic accounts of the existence of non-causal things—abstracta and the like—for another day.

It is not part of my project to argue for the absolute virtue of the naturalistic ‘ultimate causal explanations’ that will be canvassed in this article. The explanations in question depend upon controversial assumptions about causality, modality, the meaningfulness of talk about ‘ultimate explanation’, and perhaps other things as well. What I do want to argue is that, against the background of these controversial assumptions, there is good reason to prefer naturalistic ‘ultimate explanations’ to theistic ‘ultimate explanations’. Moreover, I shall argue that, if I am right in thinking that naturalistic ‘ultimate explanations’ are better than theistic ‘ultimate explanations’, then those considerations alone are sufficient to defeat all cosmological arguments for the existence of God.
According to Tim O’Connor, ‘once one sees that unreduced modality is unavoidable for ordinary explanatory purposes, a modalised response to the question of contingent existence is both natural and prima facie viable, and there is much to... more
According to Tim O’Connor, ‘once one sees that unreduced modality is unavoidable for ordinary explanatory purposes, a modalised response to the question of contingent existence is both natural and prima facie viable, and there is much to commend classical monotheism as the framework best suited to providing an outline of a comprehensive and non-arbitrary ultimate explanation’ (2). I disagree. Even if we grant—at least for the sake of argument—that unreduced modality is unavoidable for ordinary explanatory purposes, it seems to me that naturalism is the framework best suited to providing an outline of a comprehensive and non-arbitrary ultimate explanation.
How one answers the question whether time could be two-dimensional depends upon what one takes to be the essential properties of time. I assume that it is essential to time that it has a 'sense' or 'direction'; and, on the basis of this... more
How one answers the question whether time could be two-dimensional depends upon what one takes to be the essential properties of time. I assume that it is essential to time that it has a 'sense' or 'direction'; and, on the basis of this assumption, I argue that no-one has yet succeeded in giving a clear account of how it could be that time is two-dimensional. In particular, I argue that no-one has yet succeeded in describing possible circumstances in which we would have serious reason to entertain the hypothesis that time is two-dimensional.
Peter Forrest — in “How Innocent Is Mereology?” Analysis 56.3, July 1996, pp.127-131 — disputes David Lewis’ claim that mereology is ontologically innocent, on the grounds that mereology entails something (Countable Fusion) which is... more
Peter Forrest — in “How Innocent Is Mereology?” Analysis 56.3, July 1996, pp.127-131 — disputes David Lewis’ claim that mereology is ontologically innocent, on the grounds that mereology entails something (Countable Fusion) which is inconsistent with something else (a Whiteheadian account of space) which we do not know a priori to be false. However, it seems to me that the inconsistency which Forrest extracts from the conjunction of Countable Fusion with the Whiteheadian account of arises only because the Whiteheadian account of space is formulated in terms which poach on the preserves of mereology. There are formulations of theories which share almost all of the interesting features of the Whiteheadian account of space, but which do not lead to contradiction when wedded to Countable Fusion. Moreover, the other kinds of allegedly objectionable consequences which Forrest extracts from the conjunction of Countable Fusion with the Whiteheadian account of space can be extracted from his Whiteheadian account of space alone. In these circumstances, the oddities should not be taken to redound to the discredit of Countable Fusion.
Non-cognitivism in ethics holds that ethical sentences are not in the business of being either true or false – for short, they are not truth apt. No-truth theories of indicative conditionals (on one labelling of the relevant class of... more
Non-cognitivism in ethics holds that ethical sentences are not in the business of being either true or false – for short, they are not truth apt. No-truth theories of indicative conditionals (on one labelling of the relevant class of conditionals) hold that indicative conditionals have assertability or acceptability conditions, but not truth conditions; they are not truth apt. The arguments for these views are typically local to ethics and conditionals, respectively. They are not usually set within a specific theory of truth, and the question of how they connect to the various theories of truth is typically left unaddressed. This is surprising. An obvious question to ask about non-cognitivism and no-truth theories of conditionals is how they fare in the light of various views about truth.

This paper is concerned with an increasingly popular position on the obvious question. According to this position, a certain view about truth, often called ‘minimalism about truth’, leads pretty well immediately, and in any case without recourse to the considerations distinctive of the debates in the ethics’ and conditionals’ literatures, to the conclusion that ethical sentences and indicative conditionals are truth apt. A variant on this position distinguishes kinds of truth, and holds that thin truth, or disquotational truth, or at any rate some core, non-robust notion of truth, cannot sensibly be denied to ethical sentences and conditionals, and that, accordingly, the live issue should be thought of as whether, on some thick or robust notion of truth – perhaps tied to correspondence with reality in some metaphysically heavyweight sense, or to mind independence, or to evidence transcendence, or ... – ethical sentences and conditionals are truth apt. In this paper, we argue that this is a mistake. The distinctive considerations cannot be side-stepped.

As the argumentation to come is, of necessity, somewhat complex, it may help to have a map. In section one we characterise minimalism about truth and argue that the usual arguments from minimalism about truth to the falsity of non-cognitivism and no-truth theories of conditionals fail, and that only an appropriate minimalism about truth aptness could yield the sought for refutation of non-cognitivism and no-truth theories. In sections two and three we discuss two appropriately minimal accounts of truth aptness – syntacticism and disciplined syntacticism – and argue that they fail. In section three we defend a general condition on any acceptable account of truth aptness, a condition that means that no (correct) account of truth aptness can be appropriately minimal, and indeed, as we argue in section four, this general condition vindicates the relevance of the kinds of considerations distinctive of the debates in the ethics’ and conditionals’ literatures. The attempt to short-circuit them by appeal to minimalism about truth or minimalism about truth aptness is a mistake. The final section notes a moral for minimalism about truth arising from the discussion earlier of minimalism about truth aptness.

Most of what we say applies equally to non-cognitivism in ethics and to no-truth theories of conditionals. Accordingly, the discussion will usually be developed for the case of ethics alone, and often in terms of the one example, the question of the truth aptness of ‘Torture is wrong’.
Minimalism about truth has received considerable attention of late. We think that much of the discussion suffers from a pair of deficiencies. First, there has been a failure to discriminate different varieties and dimensions of minimalism... more
Minimalism about truth has received considerable attention of late. We think that much of the discussion suffers from a pair of deficiencies. First, there has been a failure to discriminate different varieties and dimensions of minimalism about truth. Second, some serious and fundamental problems for the most popular varieties of minimalism about truth have not yet received sufficient attention. This paper aims to remedy those deficiencies.

The paper is divided into three sections. In the first section, we distinguish six main varieties of minimalism about truth. In the second section, we identify four dimensions along which views about truth can be more or less minimal, thus clarifying the range of relevant notions of “minimality”. In the third section, we critically discuss four minimalist theses.
Robert Adams tells us that “Leibniz is (in my opinion) the most interesting writer on the ontological argument for the existence of God between St. Anselm and the twentieth century”. Maybe so; but I think that a good case can be made for... more
Robert Adams tells us that “Leibniz is (in my opinion) the most interesting writer on the ontological argument for the existence of God between St. Anselm and the twentieth century”. Maybe so; but I think that a good case can be made for the opinion that there is at least as much material of interest to be found in the discussion of ontological arguments in the Objections and Replies generated by Descartes’ Meditations. At any rate, I propose to attempt to show that, between them, Caterus, Mersenne -- or, at least, the “divers Theologians and Philosophers” whose views Mersenne collected and collated -- Hobbes and Gassendi provide criticisms of Descartes’ arguments which are the equal of any criticisms of ontological arguments raised between St. Anselm and the twentieth century. Moreover, I also want to suggest that the glosses which Descartes provides on the arguments of Meditation V are of considerable independent interest. I shall begin with a brief recapitulation of the arguments of Meditation V.
This paper extends the dialogue developed in Matthews, G. and L. R. Baker forthcoming. Reply to Oppy’s Fool. Analysis 71:1. This is the fourth instalment in this exchange. In my extension of the dialogue, the Fool ends up on the side of... more
This paper extends the dialogue developed in Matthews, G. and L. R. Baker forthcoming. Reply to Oppy’s Fool. Analysis 71:1. This is the fourth instalment in this exchange. In my extension of the dialogue, the Fool ends up on the side of the angels.
In Oppy (1996), I claimed that it is possible to parody Godel’s ontological argument—or, at any rate, C. Anthony Anderson’s variant of Godel’s ontological argument—in the same way in which Gaunilo parodied Anselm’s ontological argument.... more
In Oppy (1996), I claimed that it is possible to parody Godel’s ontological argument—or, at any rate, C. Anthony Anderson’s variant of Godel’s ontological argument—in the same way in which Gaunilo parodied Anselm’s ontological argument. Michael Gettings (1999) claims that the parodies which I provided do not work. I agree, more or less. I would add that there is room for dispute about whether Gettings’ Axiom 7 is really part of Godel’s argument—or of C. Anthony Anderson’s variant thereof—and for dispute about whether there is anything which Godel says which really supports the attribution of anything like Axiom 7 to him. Moreover, I continue to maintain that the first of my attempted parodies does succeed against the argument if Axiom 7 is not included. And I would also add that the second of my attempted parodies fails for reasons which have nothing to do with considerations which pertain to the new Axiom 7. But I do not propose to take up these issues here. For a minor change to the second of the attempted parodies from my earlier paper does provide a successful parody of the argument which includes Axiom 7; consequently, Gettings’ further claims about the invulnerability of Godel’s argument to Gaunilist strategies are mistaken.
This paper is an extension of the dialogue developed in Matthews, G. and L. R. Baker 2010. The Ontological Argument Simplified. Analysis 70: 210-12. While their dialogue ends with the ontological arguer triumphant, my extension of their... more
This paper is an extension of the dialogue developed in Matthews, G. and L. R. Baker 2010. The Ontological Argument Simplified. Analysis 70: 210-12. While their dialogue ends with the ontological arguer triumphant, my extension of their dialogue gives the laurels to the Fool.
Maydole (2009) writes: "Ontological arguments are captivating. They convince some people but not others. Our purpose here was not to convince, but simply to show that some ontological arguments are sound, do not beg the question, and are... more
Maydole (2009) writes:

"Ontological arguments are captivating. They convince some people but not others. Our purpose here was not to convince, but simply to show that some ontological arguments are sound, do not beg the question, and are insulated from extant parodies. Yet good logic does convince sometimes. Other times, something else is needed." (586)

On the way I understand these matters, any argument that is both sound and non-question-begging ought to be considered successful (and no argument that is subject to successful parody is sound). Be that as it may, what I propose to do here is to critically examine the ontological arguments of Maydole (2009). I claim that close examination reveals that these arguments are both question-begging and subject to successful parodies—and hence that nothing else is needed to permit non-believers to dismiss these arguments with good conscience.
Maydole (2003) presents a formal derivation of the claim that there is exactly one supreme being. He claims that this derivation is ‘arguably sound’ (311), but acknowledges that it has ‘premises, presuppositions and inference rules’ that... more
Maydole (2003) presents a formal derivation of the claim that there is exactly one supreme being. He claims that this derivation is ‘arguably sound’ (311), but acknowledges that it has ‘premises, presuppositions and inference rules’ that ‘can and, perhaps, should be challenged’ (311). In the last part of his paper, Maydole tries to address some of the potential challenges to his argument, but nonetheless allows that it would be ‘philosophically arrogant’ to claim that the argument is ‘an honest-to-god demonstration of the existence of God’ (311). Even so, he concludes that, ‘rather than being a cause for despair, this shortage can hopefully serve as an invitation to further philosophical disputation’ (311).

I’m happy to accept the extended invitation. In my view, it is quite clear that no one—theist or non-theist—should suppose that Maydole’s argument is sound. Moreover, I think, there is no serious prospect of patching Maydole’s argument to produce a successful argument for the conclusion that there is exactly one supreme being, i.e. an argument that gives reasonable people who do not already suppose that there is a supreme being a reason to accept the conclusion that there is such a being. If there is a supreme being, then sound arguments with that conclusion are a dime a dozen—and, likewise, if there is no supreme being, then sound arguments with that conclusion are equally a dime a dozen: so there is a nice question to be addressed about the distinctive virtues that Maydole might claim for his argument, given his own acknowledgement that it is not successful.
In Proslogion II, Anselm provides the following argument: But surely when this same Fool hears what I am speaking about, namely ‘something-than-which-nothing-greater-can-be-thought’, he understands what he hears, and what he understands... more
In Proslogion II, Anselm provides the following argument:

But surely when this same Fool hears what I am speaking about, namely ‘something-than-which-nothing-greater-can-be-thought’, he understands what he hears, and what he understands is in his mind, even if he does not understand that it actually exists. … Even the Fool, then, is forced to agree that something-than-which-nothing-greater-can-be-thought exists in the mind, since he understands this when he hears it, and whatever is understood is in the mind.

The conclusion of this argument—viz. that that than which no greater can be conceived exists in the understanding (or mind)—is then the first premise in what is standardly said to be Anselm’s ontological argument. In this paper, I propose to look carefully at the argument contained in the above-cited passage.

On first sight, it is tempting to set out the argument contained in the above passage in the following way:

1. When the fool hears the words ‘that than which no greater can be conceived’, he understands what he hears. (Premise)
2. Whatever is understood exists in the understanding. (Premise)
3. (Hence) That than which no greater can be conceived exists in the understanding. (From 1 and 2.)

Thus understood, the argument is problematic. The initial premise refers to ‘the fool’, and to what ‘the fool’ hears and understands. But the conclusion simply refers to ‘the understanding’, as does the second premise. On any plausible reading, then, the argument seems simply to be invalid.

In order to restore validity, there are two natural possible moves. On the one hand, we might set out the argument in a way that consistently maintains reference to ‘the fool’:

1. When the fool hears the words ‘that than which no greater can be conceived’, he understands what he hears. (Premise)
2. Whatever the fool understands exists in his understanding. (Premise)
3. (Hence) That than which no greater can be conceived exists in the understanding of the fool. (From 1 and 2.)

On the other hand, we might set out the argument in a way that consistently omits all mention of ‘the fool’:

1. When the words ‘that than which no greater can be conceived’ are heard, they are understood. (Premise)
2. Whatever is understood exists in the understanding. (Premise)
3. (Hence) That than which no greater can be conceived exists in the understanding. (From 1 and 2.)

The second rendition of the argument is at least potentially ambiguous. Read one way, the second rendition of the argument is merely a kind of generalisation of the first rendition, and so involves no more problematic commitments than the first rendition:

1. At least some people who hear the words ‘that than which no greater can be conceived’ understand what they hear. (Premise)
2. Whatever is understood by someone exists in the understanding of that one. (Premise)
3. (Hence) That than which no greater can be conceived exists in the understandings of at least some people. (From 1 and 2.)

However, read another way, the second rendition of the argument involves a further, at least potentially problematic, commitment beyond the commitments incurred in acceptance of the first rendition—for, on this second reading, we suppose that ‘the understanding’ refers to something that is independent of any particular human individual. On this second reading, at least roughly speaking, the conclusion of the argument could be true even if there were no human individuals, nor any other kinds of cognitive agents.
A modal theistic argument is a proof of the existence of God which makes use of the premise that God is a being who exists in every possible world. Such arguments have been advanced by Alvin Plantinga, and more recently by Brian Leftow.... more
A modal theistic argument is a proof of the existence of God which makes use of the premise that God is a being who exists in every possible world. Such arguments have been advanced by Alvin Plantinga, and more recently by Brian Leftow. In this paper, I provide a general ground for objecting to all modal theistic arguments. Moreover, I suggest that there is something important about recent conceptions of modality which can be learned from these arguments.
Bruce Langtry (1999) argues against the general objection to ontological arguments presented in my book Ontological Arguments and Belief in God. I am no longer sure that that general objection is correctly expressed in my book—and,... more
Bruce Langtry (1999) argues against the general objection to ontological arguments presented in my book Ontological Arguments and Belief in God. I am no longer sure that that general objection is correctly expressed in my book—and, indeed, I am no longer confident that there is such a general objection to be given—but I also do not think that Langtry’s criticisms of that objection are quite right. What I propose to do here is the following: first, to briefly rehearse the general objection to ontological arguments given in my book; second, to briefly recapitulate Langtry’s criticisms of this general objection; third, to explain why I think that Langtry’s criticisms are ineffective; and fourth to air some doubts of my own about the argument which I originally defended.
ABSTRACT Millican (Mind 113(451):437–476, 2004) claims to have detected ‘the one fatal flaw in Anselm’s ontological argument.’ I argue that there is more than one important flaw in the position defended in Millican (Mind 113(451):437–476,... more
ABSTRACT Millican (Mind 113(451):437–476, 2004) claims to have detected ‘the one fatal flaw in Anselm’s ontological argument.’ I argue that there is more than one important flaw in the position defended in Millican (Mind 113(451):437–476, 2004). First, Millican’s reconstruction of Anselm’s argument does serious violence to the original text. Second, Millican’s generalised objection fails to diagnose any flaw in a vast range of ontological arguments. Third, there are independent reasons for thinking that Millican’s generalised objection is unpersuasive.
Pruss (2010) offers a novel defence of possibility premises in modal ontological arguments. In particular, he offers a new way of arguing that these possibility premises are ‘probably true’. I propose to argue that Pruss’s defence is... more
Pruss (2010) offers a novel defence of possibility premises in modal ontological arguments. In particular, he offers a new way of arguing that these possibility premises are ‘probably true’. I propose to argue that Pruss’s defence is unconvincing: in the end, it probably amounts to nothing more than an expression of prejudice against worldviews that reject that claim that God is essentially omnipotent, essentially omniscient, essentially perfectly good, essentially the creator of all else, and necessarily existent.
Mereological ontological arguments are -- as the name suggests -- ontological arguments which draw on the resources of mereology, i.e. the theory of the part-whole relation. An instance of arguments of this kind is the following: 1. I... more
Mereological ontological arguments are -- as the name suggests -- ontological arguments which draw on the resources of mereology, i.e. the theory of the part-whole relation.

An instance of arguments of this kind is the following:

1. I exist. (Premise, contingent a priori)
2. (Hence) Some -- i.e. least one -- thing exists. (From 1)
3. Whenever some things exist, there is some thing of which they are all parts. (Premise, from mereology)
4. (Hence) There is exactly one thing of which every thing is a part. (From 2, 3)
5. The unique thing of which every thing is a part is God. (Definition, pantheism)
6. (Hence) God exists. (From 4, 5)

The status of premise 1 is controversial: friends of two-dimensional modal logic (and others) will be reluctant to grant that the proposition that I exist is both contingent and knowable a priori  (even by me). Instead, they will insist that all that I know a priori is that the sentence “I exist” expresses some true proposition or other when I token it. But, of course, even that will suffice for the purposes of the argument. Provided that I know a priori that the sentence “I exist” expresses some true singular proposition or other -- i.e. some proposition or other which contains an individual -- then I have an a priori guarantee that there are some individuals, and so I am entitled to assert 2. Of course, it will remain true that there are some people who refuse to accept 2: consider, for example, those ontological nihilists who think that the proper logical form of every sentence can be given in a feature-placing language.
However, many people will be prepared to grant that we can know a priori
that there are at least some individuals -- and that it enough to sustain interest in our argument to this point.

The status of premise 3 is also controversial: there are various reasons why one might be inclined to reject it. However, it is important to be clear about exactly what the premise says. Note, in particular, that it does not say that, whenever some things exist, there is some thing which is the mereological sum of those things. Rather, what it says is that, whenever some things exist, there is some thing of which all of those things are parts -- i.e. the thing completely overlaps each of the parts, but the parts together need not completely overlap the thing. Of course, given the mereological claim about sums, the weaker claim follows: so friends of unrestricted mereological composition will certainly be happy with 3. But one could subscribe to 3 on independent grounds: one might think, for example, that it is just impossible for there to be two things which are not both parts of a single, more-inclusive, thing. Again, there will be people who are not prepared to accept 3. But, for now, it seems reasonable to suppose that there will be lots of people who are quite happy with it. (We shall have more to say about 3 later.) The inference of 4 from 3 looks distinctly suspicious. Indeed, it seems to have the form of the quantifier-exchange fallacy which moves from ∀∃ to ∃∀. However, we can patch this. What we need to suppose is that we can talk unrestrictedly about every thing. Now, consider all things. If premise 3 is correct, then it does indeed follow that there is some thing of which every thing is a part. (By ‘part’, I mean ‘proper or improper part’ of course.) Moreover, it is then extremely plausible to suggest that there can only be one such thing: in order to deny this, one would need to deny the uniqueness of composition (a course which is possible, but, at least prima facie, quite unattractive). Of course, some people will not be happy with the claim that we can talk unrestrictedly about every thing. Among the reasons which might be given for this unhappiness, perhaps the most important is the suggestion that unrestricted quantification leads to paradox. However, it is important to bear in mind that we are talking about quantification over individuals here. Whether one supposes that there are finitely many, or countably many, or continuum many, or Beth-2 many, or even proper class many, individuals, it is hard to see how any contradiction can arise from this assumption. Of course, there are other objections which one might make to the totality assumption. However, it again seems reasonable to suppose that there will be lots of people who are quite happy with it. (Once more, we shall return to this assumption later.)

On the basis of the above considerations, it seems reasonable to suggest that there will be lots of people -- including lots of people who do not count themselves as having any kinds of religious beliefs -- who will be happy with the argument to 4. Or, perhaps better, there will be lots of people -- including lots of people who do not count themselves as having any kinds of religious beliefs -- who will be prepared to accept the following argument at least as far as 5:

1. I exist. (Premise, contingent a priori)
2. Some things -- i.e. at least one -- exist. (From 1)
3. If some things exist, then there are some things which are all of the things that exist. (Premise, from the meaning of ‘all’.)
4. Whenever some things exist, there is some thing of which they are all parts. (Premise, from mereology)
5. There is exactly one thing of which every thing is a part. (From 3, 4)
6. The unique thing of which every thing is a part is God. (Definition, pantheism)
7. Hence God exists. (From 5, 6)

In other words, there will be lots of people who are happy to allow -- on more or less a  priori grounds -- that there is exactly one thing of which every thing is a part. So, for these people, the important question will be whether the thing of which every thing is a part deserves to be called ‘God’. If this thing does deserve the name, then pantheism is vindicated; if this thing does not deserve the name, then -- presumably -- pantheism (or, at least, this kind of pantheism) is simply a mistake
First, I suggest that it is possible to make some further improvements upon the Gödelian ontological arguments that Pruss develops. Then, I argue that it is possible to parody Pruss’s Gödelian ontological arguments in a way that shows... more
First, I suggest that it is possible to make some further improvements upon the Gödelian ontological arguments that Pruss develops. Then, I argue that it is possible to parody Pruss’s Gödelian ontological arguments in a way that shows that they make no contribution towards ‘lowering the probability of atheism and raising the probability of theism’. I conclude with some remarks about ways in which the arguments of this paper can be extended to apply to the whole family of Gödelian ontological arguments.

The outline of my paper is as follows. In the first part, I construct an ontological argument which, I claim, is an improvement upon the arguments that Pruss produces. In the second part, I argue that this ontological argument—and, by extension, each of the arguments that Pruss develops in his paper—is plainly not a successful argument. In particular, I point out that there are, after all, effective parodies of Pruss’s arguments. While I don’t have space to develop the further argument in this paper, I claim that the preceding observations can be developed into a general critique of Gödelian ontological arguments.
There is now a considerable secondary literature on Godel’s ontological arguments; in particular, interested readers should consult Sobel (1987), Anderson (1990) and Adams (1995). In this note, I wish to draw attention to an objection to... more
There is now a considerable secondary literature on Godel’s ontological arguments; in particular, interested readers should consult Sobel (1987), Anderson (1990) and Adams (1995). In this note, I wish to draw attention to an objection to these arguments which has hitherto gone unnoticed. This objection does not depend upon fine details of the formulation of the arguments; I arbitrarily choose to develop the objection in connection with the formulation provided by Anderson.
In “To Bet The Impossible Bet”, Harmon Holcomb III argues that “[W]e should not accept [the terms of Pascal’s Wager] since they violate the preconditions of it making sense to bet. The wager is useless because of a structural breakdown in... more
In “To Bet The Impossible Bet”, Harmon Holcomb III argues that “[W]e should not accept [the terms of Pascal’s Wager] since they violate the preconditions of it making sense to bet. The wager is useless because of a structural breakdown in the conditions that determine the relations between the act of betting and the payoff matrix options. ... A bet is genuine only if there is a chance of success in reaping the bet’s rewards. Any chance vanishes if the wagerer’s act of betting something (what is bet) on something (a betting option) puts the same thing in both roles: we cannot be forced to bet by believing or not believing on the options of believing or not believing. Neither an irresistable temptation nor a bribe, the wager is an inconceivable bet. No question of its prudence arises, any more than that of how best to square the circle.”

I disagree. If there is a structural breakdown in the formulation of Pascal’s Wager, it comes with the introduction of infinite utilities; otherwise, the wager is a straightforward application of decision theory in the calculation of the expected utilities of alternate (courses of) actions. Moreover, I disagree with Holcomb’s further claim that “If it is a possible bet, the wager argument works; if it is not a possible bet, it doesn’t work”. Subject to the possible qualification involving infinite utilities, Pascal’s Wager is “a possible bet”; but there are numerous reasons why it is not a very good one. In particular: (i) as it stands, the Wager gives infinite utility to every course of action which is open to those who are giving consideration to the wager; (ii) there are many alternative formulations of the payoff matrix in Pascal’s Wager, and not all of these formulations give the same result; and (iii) there are many alternative wagers (most involving alternatives to Pascal’s God), and consideration of these wagers undermines the apparent cogency of Pascal’s argument. Holcomb claims that these criticisms rest on a mistaken logicist methodology; however, I shall argue that this suggestion rests on a misunderstanding of the nature of the distinction between theoretical and practical reasoning.

I begin with this last point.
In Pascal's Wager: A Study Of Practical Reasoning In Philosophical Theology, Nicholas Rescher aims to show that, contrary to received philosophical opinion, Pascal's Wager argument is "the vehicle of a fruitful and valuable insight -- one... more
In Pascal's Wager: A Study Of Practical Reasoning In Philosophical Theology, Nicholas Rescher aims to show that, contrary to received philosophical opinion, Pascal's Wager argument is "the vehicle of a fruitful and valuable insight -- one which not only represents a milestone in the development of an historically important tradition of thought but can still be seen as making an instructive contribution to philosophical theology". In particular, Rescher argues that one only needs to adopt a correct perspective in order to see that Pascal's Wager argument is a good argument. Moreover, there seems to be a certain amount of contemporary support for Rescher's claim that Pascal's Wager argument can be seen to be a good argument when properly construed. However, despite this recent trend to adopt a more sympathetic stance towards Pascal's Wager argument, I propose to defend the traditional view that Pascal's Wager argument is almost entirely worthless -- at least from the theological standpoint. (No doubt, it has historical significance from the standpoint of decision theory; but that's a separate matter.)

This paper is divided into two sections. I begin, in section I, by outlining the defence of Pascal's Wager argument which is given by Rescher in Pascal's Wager. Then, in section II, I explain why this defence fails.
In Analysis 46.3, June 1986, Timothy Williamson claims to have produced an instance of a 'deeply contingent a priori truth which in no way turns on indexicals' ([3], p.113). However, I want to suggest that his argument does not establish... more
In Analysis 46.3, June 1986, Timothy Williamson claims to have produced an instance of a 'deeply contingent a priori truth which in no way turns on indexicals' ([3], p.113). However, I want to suggest that his argument does not establish this claim.
Semantic minimalisms come in many varieties. Local semantic minimalisms offer minimalist analyses of a restricted range of semantic expressions, whereas global semantic minimalisms offer minimalist analyses across the board. Local... more
Semantic minimalisms come in many varieties. Local semantic minimalisms offer minimalist analyses of a restricted range of semantic expressions, whereas global semantic minimalisms offer minimalist analyses across the board. Local semantic minimalisms for selected classes of expressions themselves come in many varieties.

Consider truth predicates. Minimalist analyses of truth predicates may involve commitment to some of the following claims: (i) truth “predicates” are not genuine predicates -- either because the truth “predicate” disappears under paraphrase or translation into deep structure, or because the truth “predicate” is shown to have a non-predicative function by performative or expressivist analysis, or because truth “predicates” must be traded in for predicates of the form “true-in-L”; (ii) truth predicates express ineligible, non-natural, gerrymandered properties; (iii) truth predicates express metaphysically lightweight properties; (iv) truth predicates have thin conceptual roles; (v) truth predicates express properties with no hidden essence; (vi) truth predicates express properties which have no causal or explanatory role in canonical formulations of fundamental theories.

Behind this diversity, there is some unity. In the case of truth predicates, it seems plausible to suggest that almost all minimalists hold: (i) that there is no interesting or important difference between the claim that p and the claim that ‘p’ is true; and (ii) that the disquotational property of truth predicates, manifested in the truth-schema ‘p’ is true iff p, is the only theoretically interesting or important primitive property of those predicates. However, it is dubious that this ‘minimalist core’ characterises a single position which forms a useful subject for philosophical investigation; rather, further commitments must be added in order to provide precise targets for scrutiny. And what goes for truth goes for other semantic notions as well: ‘minimalism’ covers such a multitude of sins that nothing but confusion is generated by generalisations which fail to attend to the particularities of each sin.

In this paper, I propose to investigate that kind of minimalist view which is elaborated and defended by Crispin Wright in his book Truth and Objectivity and subsequent publications. I shall be particularly concerned to argue that no satisfactory position emerges from those writings; however, I shall also suggest that it is quite unclear what patches could be applied ¬¬– i.e. I shall suggest that Wright’s minimalism belongs to a family of views none of which is satisfactory.

Each of the members of this minimalist family is committed to the following four, admittedly somewhat vague, theses:

(1)  Variation in Semantic Width: Semantic predicates vary in width – so, e.g., truth and (maybe?) truth-aptness are thicker, or more robust, for some areas of discourse than they are for others;

(2)  Minimal Truth: Minimal truth does not make demands which go much beyond the demands of warranted assertability -- e.g. in appropriately restricted domains, superassertability  coincides with truth;

(3)  Minimal Truth-Aptness: A sufficient condition for (minimal) truth-aptness for sentences is provided by satisfaction of certain constraints on fit substituends in the truth-schema, viz: (a) they must have the correct syntax -- roughly, having that declarative form which fits them for embedding by negations, conditionals, verbs of propositional attitude, etc. -- and (b) they must be disciplined by norms of appropriate utterance.  ;

(4)  Objectivity Reconceived: Acceptance of (2) and (3) forces a reconceptualisation or refiguration of the nature of local debates between realists and anti-realists, since non-cognitivisms are ruled out, and error theories are extraordinarily difficult to justify; thus, e.g. in the case of ethics, (2) and (3) entail that ethical non-cognitivisms are ruled out, and that Mackie’s error theory is utterly implausible.

One strategy which one might take in arguing against the family of views is to argue against a particular thesis from this list. I shall begin with some suggestions about the construction of cases against each thesis (Part 1); however, these suggestions will be primarily in the nature of preliminary skirmishes conducted with a view to clarifying some of the important vocabulary involved.  Another strategy which one might take in arguing against the family of views is to argue that there is something unsatisfactory about the package of theses as a whole. This is the form of argument which I wish to pursue, in connection with questions about reference (Part 2). In particular, I shall argue that no distinctive and plausible view about singular term reference and predicate reference coheres with, or emerges from, the package of theses.

The intuitive thought which motivates the argument is this: If truth and truth-aptness are (relatively) easy to achieve -- as our minimalists suppose -- then so is reference. But (relative) ease of reference threatens an explosion of ontological, ideological and theoretical commitments. If one is concerned to avoid this explosion -- as plausibly one ought to be, and as Wright explicitly is -- then one must make use of strategies which, I claim, are the tools-in-trade of error theorists and/or non-cognitivists. But then it follows that the acceptance of thesis (4) is a mistake -- those techniques which are supposed to be available to rule out commitments to, say, fictional entities (Part 3), can be adapted to make room for error theories and/or non-cognitivisms in domains such as ethics (Part 4). So our minimalists should renounce their kind of minimalism -- something in the package of theses must be given up.
This paper is a critque of Quinean arguments against direct reference. In particular, it criticises the argument developed by Steven Wagner in "California Semantics meets the Great Fact".
In ‘Do I Have To Be Here Now?’, Professor C. J. F. Williams argues against Kaplan’s claim, that the sentence ‘I am here now’ expresses a contingent but analytically true proposition in the mouth of any contingently existing speaker, on... more
In ‘Do I Have To Be Here Now?’, Professor C. J. F. Williams argues against Kaplan’s claim, that the sentence ‘I am here now’ expresses a contingent but analytically true proposition in the mouth of any contingently existing speaker, on the grounds that ‘the sentence is not a proposition’.

Following Geach, Williams professes to be unhappy with the suggestion that propositions are abstract entities expressed by sentences; instead, he prefers to say that some sentences are propositions. However, there are surely further options. In particular, it seems that one could choose to construe talk of propositions as talk about, e.g., the conditions under which utterances of sentences are equivalent – in extension, or intension, or character, or whatever – without supposing that this talk commits one to any extra entities. Indeed, given objects and properties – both of which Williams admits into his ontology – one only needs some of the resources of set theory in order to generate the semantic theory to which Kaplan commits himself. Since Williams himself provides no clue about how he would construct a semantic theory to encode the information which Kaplan’s theory encodes, this seems to be a point in favour of Kaplan’s view.

Suppose we set these kinds of considerations aside. The point of this note is to suggest that, even then, Williams’ position faces a severe internal difficulty, which arises from his assumption that a sentence is a proposition only if it contains referential devices which operate independently of context of utterance.
In Naming And Necessity, Saul Kripke suggests that there may be contingent propositions which are knowable a priori (by creatures like us), and that there may be necessary propositions which are knowable only a posteriori (by creatures... more
In Naming And Necessity, Saul Kripke suggests that there may be contingent propositions which are knowable a priori (by creatures like us), and that there may be necessary propositions which are knowable only a posteriori (by creatures like us). As an example of a contingent proposition which is knowable a priori (by creatures like us) Kripke offers: the proposition which is the content of the sentence The length of stick S at to is one metre, where stick S is the standard metre, and to is the time at which the reference of the expression "one metre" is fixed for one via reference to (the length of) S. And, as an example of a necessary proposition which is knowable only a posteriori (by creatures like us) Kripke suggests: the proposition which is the content of the sentence Hesperus is Phosphorus.

In Frege’s Puzzle, Nathan Salmon argues that, although the general claims which Kripke makes are correct, the examples which he gives do not carry conviction. On the one hand, Salmon argues that, although there are necessary propositions which are knowable only a posteriori (by creatures like us), the sentence Hesperus is Phosphorus actually has as its content a necessary proposition which is knowable a priori (by creatures like us). And, on the other hand, Salmon argues that although there are contingent propositions which are knowable a priori (by creatures like us), the sentence The length of stick S at to is one metre actually has as its content a contingent proposition which is only knowable a posteriori (by creatures like us).

In this paper, I propose to argue: (i) that if Salmon's case against Kripke's examples is good, then it can be extended to show that there are no examples of the contingent a priori; and (ii) that if Salmon were to grant that Kripke's examples are good, while maintaining the rest of his views, then he would very likely be committed to the conclusion that all singular propositions -- or, at least, all suitably existentially
circumscribed singular propositions -- can be known a priori. I shall conclude with some suggestions about how Salmon's views ought to be amended.
These comments, on the paper by Branden Thornhill-Miller and Peter Millican and on the critique of that paper by Janusz Salamon, divide into four sections. In the first two sections, I briefly sketch some of the major themes from the... more
These comments, on the paper by Branden Thornhill-Miller and Peter Millican and on the critique of that paper by Janusz Salamon, divide into four sections. In the first two sections, I briefly sketch some of the major themes from the paper by Thornhill-Miller and Millican, and then from the critique by Salamon. In the final two sections, I provide some critical thoughts on Salamon's objections to Thornhill-Miller and Millican, and then on the leading claims made by Thornhill-Miller and Millican. I find much to commend, but also some things to dispute, in both papers. As is so often the way, I shall focus on areas of disagreement. (1) Thornhill-Miller and Millican argue that rationality requires a retreat from 'first-order religion'. Their argument has two main prongs: (a) 'The Common Core/Diversity Dilemma'; and (b) 'The Normal/Objective Dilemma'. (2-5) The Common Core/Diversity Dilemma has two horns: (A) in so far as religious phenomena point towards specific aspects of particular religions, their diversity and mutual opposition undermines their evidential force; and (B) in so far as such phenomena involve a 'common core' of similarity, they point towards a proximate common cause for these phenomena that is natural rather than supernatural. (20) Thornhill-Miller and Millican argue that (A) is supported by, for example, considerations about medical miracles and intercessory prayer (21-3); and that (B) is supported by psychometric studies and considerations concerning near death experiences, meditative and introvertive religious experiences, hypersensitive agency detection devices, and theories of mind (23-31). Moreover, Thornhill-Miller and Millican also argue that considerations about egocentric bias, confirmation bias, and needs for significance and social cohesion point to proximate naturalistic explanations of the fact that religions are so assertive and persistent in their claims to special authority in the face of obvious disagreement from so many competing faiths. (32-7) The Normal/Objective Dilemma is really a question: if the psychological causes of religious belief are associated with normal, healthy mental functioning and various positive (individual and social) outcomes, should these rationally weigh with us more heavily than objective epistemological considerations would allow? (40) Thornhill-Miller and Millican note that there are individual and social benefits of religious belief-enhanced happiness, increased longevity, improved recovery from addiction, deepened in-group trust, heightened in-group empathy, greater in-group cohesion, and so forth (37-41)-as well as individual and social costs of religious belief-intensified out-group conflict, increased insularity, greater xenophobia, heightened prejudices, and so on (41-3). Thornhill-Miller and Millican take the view that the in-group benefits are outweighed by the out-group damage; in their view, there may be no greater threat to humanity than intergroup conflict motivated by exclusivist and other-worldly religious thinking. (41) However, Thornhill-Miller and Millican also note that the very naturalness of religion makes it very doubtful that we can simply replace it with other things that deliver the same goods that it delivers: humanity is deeply immersed in well-established religious traditions whose rituals have evolved to fit human needs. (45)
Oppy (2006) contains an extensive discussion of the understanding and application of a conception of the infinite that is fundamentally mathematical: questions about the mathematically infinite are questions about the cardinality of... more
Oppy (2006) contains an extensive discussion of the understanding and application of a conception of the infinite that is fundamentally mathematical: questions about the mathematically infinite are questions about the cardinality of collections, or the divisibility of time and space, or about the magnitude of measurable properties, and so forth. But—following the discussion of Anaximander and other early philosophers—we might wonder whether the discussion of the mathematically infinite really does exhaust the discussion of the infinite. Is it the case that the concept of the infinite is, in all essentials, the concept of the mathematically infinite—or is it rather the case that the concept of the infinite is importantly ambiguous in such a way that we can also discern something that might properly be called a non-mathematical conception of the infinite? In order to address this question, I shall largely follow the lead of Sweeney (1992), which surveys the range of attributions of ‘infinity’ to monotheistic gods.
Alan Keightley, in Wittgenstein, Grammar and God, identifies a group of ‘devout Wittgensteinians’ including some of Wittgenstein’s students, such as Rush Rhees and Norman Malcolm, and some of those who have been closely associated with... more
Alan Keightley, in Wittgenstein, Grammar and God, identifies a group of ‘devout Wittgensteinians’ including some of Wittgenstein’s students, such as Rush Rhees and Norman Malcolm, and some of those who have been closely associated with the work of Wittgenstein’s pupils, including Peter Winch and D.Z. Phillips. These and like-minded thinkers have come to form what might be called ‘the Wittgensteinian school of philosophy of religion’. One feature of the writings of members of this ‘school’ is that they make use of a theoretical vocabulary that is derived from the writings—and, in particular, the later writings—of Wittgenstein: ‘language game’, ‘form of life’, ‘grammatical observation’, ‘philosophical grammar’, ‘depth grammar’, and the like. While there are differences in the views of the members of the ‘school’, there are also many important similarities. In particular, members of the ‘school’ are agreed that religion constitutes a ‘form of life’, and that ‘the religious language game’ has a unique ‘grammar’ that is utterly misrepresented in standard philosophical discussions of religion and religious belief.

Critics of ‘the Wittgensteinian school of philosophy of religion’ are apt to complain that the crucial theoretical vocabulary involved in these formulations is, at best, “obscure and ambiguous”. While the key terms are undeniably theoretical—they are,
after all, the terms of art of a particular philosophical movement—there is nowhere that they have ever been given a satisfactory explication. Moreover—and perhaps more importantly—even if we suppose that we can give sufficient sense to this theoretical vocabulary, it seems clear that there are competing theories, couched in terms of alternative vocabularies, that should be preferred because of their greater explanatory scope, explanatory power, advancement of understanding, and so forth.
This five volume work is an edited history of Western Philosophy of Religion. The 106 chapters cover many -- though, of course, not all -- of the major figures in the history of western philosophy of religion.
Reading Philosophy of Religion combines a diverse selection of classical and contemporary texts in philosophy of religion with insightful commentaries. Offers a unique presentation through a combination of text and interactive commentary... more
Reading Philosophy of Religion combines a diverse selection of classical and contemporary texts in philosophy of religion with insightful commentaries. Offers a unique presentation through a combination of text and interactive commentary Provides a mix of classic and contemporary texts, including some not anthologized elsewhere Includes writings from thinkers such as Aquinas, Boethius, Hume, Plantinga and Putnam Divided into sections which examine religious language, the existence of God, reason, argument and belief, divine properties, and religious pluralism
I begin with a very brief, very general outline of the history of Western philosophical treatments of theism. I then discuss contributions to this history from pagans, Jews, Muslims, Christians, and non-believers. Finally, I make some... more
I begin with a very brief, very general outline of the history of Western philosophical treatments of theism. I then discuss contributions to this history from pagans, Jews, Muslims, Christians, and non-believers. Finally, I make some comments about the range of conceptions of God that is evident in this history.
Mark Nelson thinks that Bertrand Russell’s well-known criticisms of St. Thomas Aquinas turn on Russell’s acceptance of a highly implausible epistemic principle (DAM), and that my previous objection to this claim depends upon the... more
Mark Nelson thinks that Bertrand Russell’s well-known criticisms of St. Thomas Aquinas turn on Russell’s acceptance of a highly implausible epistemic principle (DAM), and that my previous objection to this claim depends upon the attribution to Russell of an even more implausible Insincerity Objection. While I agree that Russell’s criticisms do not turn on the Insincerity Objection, I argue that my previous rejection of the attribution of (DAM) to Russell is well-justified: there is a plausible reading of Russell that requires neither (DAM) nor the Insincerity Objection.
Mark Nelson cites Russell on Aquinas: There is little of the true philosophic spirit in Aquinas. He does not, like the Platonic Socrates, set out to follow wherever the argument may lead. He is not engaged in an inquiry, the result of... more
Mark Nelson cites Russell on Aquinas:

There is little of the true philosophic spirit in Aquinas. He does not, like the Platonic Socrates, set out to follow wherever the argument may lead. He is not engaged in an inquiry, the result of which it is impossible to know in advance. Before he begins to philosophise, he already knows the truth; it is declared in the Catholic faith. If he can find apparently rational arguments for some parts of the faith, so much the better; if he cannot, he need only fall back on revelation. The finding of arguments for a conclusion given in advance is not philosophy, but special pleading. I cannot therefore, feel that he deserves to be put on a level with the best philosophers either of Greek or of modern times.

He then goes on to claim that “like many of Russell’s pronouncements, this is breathtakingly supercilious and unfair”. Later, he adds that Russell’s “dismissal” of Aquinas is “sniffy”.

I think that there is probably about as much justice in the claim that it is Nelson himself who is “breathtakingly supercilious and unfair”—and “dismissive” and “sniffy”—in his treatment of this passage from Russell. While it is unclear how far one ought to agree with Russell’s assessment of Aquinas’ standing as a philosopher—and while it is not entirely clear exactly what grounds Russell has for the assessment which he makes—I think that a good case can be made for the claim that Nelson completely misrepresents the position which Russell develops in the final five paragraphs of his chapter on Aquinas. The aim of the present paper is to make this case.
Judge William Overton’s Memorandum Opinion in the case of Rev. Bill McLean et. al. v. The Arkansas Board of Education et. al. occasioned a small flurry of papers by philosophers of science debating the merits of the judicial decision.... more
Judge William Overton’s Memorandum Opinion in the case of Rev. Bill McLean et. al. v. The Arkansas Board of Education et. al. occasioned a small flurry of papers by philosophers of science debating the merits of the judicial decision. Almost twenty years have now elapsed since the conclusion of the trial, and the time seems ripe for a retrospective assessment of the merits of the judge’s decision, and of the criticisms of that decision which were made by the philosophers of science in question. Although I think that there wasn’t much wrong with the judge’s opinion, I propose to argue that the claims made by the various philosophers were also substantially correct. This might sound surprising, given the heat generated by the discussion. However, it seems to me that what happened was a classic case of misunderstanding: people talked straight past each other because they placed different construals on the key terms involved.
This book is a detailed examination of a wide range of arguments for the existence of God. It contains chapters on: the theory of argumentation; ontological arguments; cosmological arguments; teleological arguments; Pascal's wager;... more
This book is a detailed examination of a wide range of arguments for the existence of God. It contains chapters on: the theory of argumentation; ontological arguments; cosmological arguments; teleological arguments; Pascal's wager; arguments for design; and a miscellany of "lesser" arguments.
This five volume work is an edited history of Western Philosophy of Religion. The 106 chapters cover many -- though, of course, not all -- of the major figures in the history of western philosophy of religion.
In this chapter, I provide a chronological survey of Plantinga’s changing conceptions of the project of natural theology, and of the ways in which those conceptions of the project of natural theology interact with his major philosophical... more
In this chapter, I provide a chronological survey of Plantinga’s changing conceptions of the project of natural theology, and of the ways in which those conceptions of the project of natural theology interact with his major philosophical concerns. In his earliest works, Plantinga has a very clear and strict conception of the project of natural theology, and he argues very clearly (and correctly) that that project fails. In his middle works, Plantinga has a tolerably clear and slightly less strict conception of the project of natural theology, and he argues—in my view unsuccessfully—that this project succeeds. In his later works, Plantinga has a much less clear and less strict conception of the project of natural theology, and it is much harder to determine whether there is any merit in the claims that he makes for natural theology as thus conceived.
This book sketches what I take to be the best argument about the existence of God. I don't claim that the argument is successful; I do claim that there is no alternative but to consider arguments like the one that is sketched here.
This edited work collects together new essays on a range of important topics in philosophy of religion. The volume aims to be a reasonably comprehensive introduction to contemporary philosophy of religion. It divides into a number of... more
This edited work collects together new essays on a range of important topics in philosophy of religion. The volume aims to be a reasonably comprehensive introduction to contemporary philosophy of religion. It divides into a number of sections: theoretical orientations; conceptions of divinity; metaphysics; epistemology; ethics; politics; and science.
This book provides an opinionated introduction to philosophy of religion. After a general introduction, the work divides into a discussion of epistemology, metaphysics, and ethics.
Despite the evident difficulties involved in the making of generalisations about religion and philosophy, we have recently witnessed a flood of ‘new atheist’ attacks on religion and religious belief in the name of philosophy and reason.... more
Despite the evident difficulties involved in the making of generalisations about religion and philosophy, we have recently witnessed a flood of ‘new atheist’ attacks on religion and religious belief in the name of philosophy and reason. In the works of Richard Dawkins, Sam Harris, Christopher Hitchens, and other like-minded ‘new atheist’ authors, we find declarations that religion is an enemy of philosophy and reason, a hangover from our barbaric past that needs to be extinguished with extreme prejudice.

While these authors are particularly disturbed by the recent rise (and excesses) of militant Islam and evangelical Christianity, the scope of their critiques extends to all manifestations of religious belief: in their view, there can be no such thing as reasonable religious faith.

In Oppy (2006), I argue at length for the claim that there can be reasonable difference of opinion on the question whether there exists a standardly conceived monotheistic god. Moreover, I argue—at even greater length—for the claim that there are no successful arguments about the existence of standardly conceived monotheistic gods: no arguments that ought to persuade those who have reasonable views about the existence of standardly conceived monotheistic gods to change their minds. Given this history, it should not be surprising that I am somewhat at odds with the opinions expressed by the ‘new atheists’. While I agree with them that we should not think that religious faith is, in itself, a virtue—and while I also agree with them that we should not think that religious faith must be respected merely because it is religious faith—I think that it is a mistake to suppose that there cannot be reasonable religious belief. If we suppose that it is true, by definition, that faith requires no justification and brooks no argument, then it seems to me that there can be religious belief that is not religious faith. Alternatively, if we suppose that religious belief entails religious faith, then we should reject the idea that it is simply true by definition that faith requires no justification and brooks no argument.
Given the diversity of systems of religious and philosophical beliefs to which I adverted earlier, it seems to me that it is incredible to suppose that there are no religious believers who are reasonable in their religious beliefs, at least by any ordinary standards of reasonableness. Moreover, I think that my own interactions with religious believers bears out the claim that, not only can there be religious believers who are, by any ordinary standards of reasonableness, reasonable in their religious beliefs, but also that there actually are religious believers who are, by any ordinary standards of reasonableness, reasonable in their religious beliefs. Of course, to say this much is not to say very much: even David Hume insisted that, while he would naturally conclude that a man was a scoundrel on hearing him making religious affirmations, nonetheless, some of the best men that he knew were religious believers. Hence, perhaps, it might seem that a straightforward accommodation between my views about reasonable religious belief and the views adopted by the ‘new atheists’ could be reached by way of (acceptance of) the claim that, by ordinary standards of reasonableness, there are very few reasonable religious believers.

But I do not believe this accommodating claim either. It seems to me that, at least by ordinary standards of reasonableness, there are many reasonable religious believers: many religious moderates who are appalled by (for example) the excesses of militant Islam and evangelical Christianity, and whose religious beliefs give no comfort to terrorism and religious extremism. Harris (2005:148) notes, inter alia, that the religious beliefs of the Jains would lead them to condemn the excesses of militant Islam and evangelical Christianity; and surely the same thing goes for the religious beliefs of High Church Anglicans and members of other ‘liberal’ Christian denominations. It seems to me that there is nothing in the religious beliefs of those who belong to ‘liberal’ Christian denominations that requires them to give even tacit support to the excesses of militant Islam and evangelical Christianity. Here, Harris (2005:20) strongly disagrees:

"While moderation in religion may seem a reasonable position to stake out in light of all that we have (and have not) learned about the universe, it offers no bulwark against religious extremism and religious violence. From the perspective of those seeking to live by the letter of the texts, the religious moderate is nothing more than a failed fundamentalist. He is, in all likelihood, going to wind up in hell with the rest of the unbelievers. The problem that religious moderation poses for us all is that it does not permit anything very critical to be said about religious literalism."

I think that Harris is just wrong about this. It seems to me to be evidently true that some ‘religious moderates’ have (quite rightly) had very critical things to say about religious literalism; and Harris does damage to the wider struggle against religious ‘enthusiasm’ in failing to recognise the contribution that ‘religious moderates’ can and do make to this wider struggle. Given that ‘religious moderates’ can be—and are—effective and vociferous critics of religious ‘enthusiasm’, we should have no part of the ‘new atheist’ attack on ‘religious moderates’.

Of course, even if it is agreed that the ‘new atheists’ go too far in the claims that they make about the rationality of religious belief in the case of ‘religious moderates’, it might still be firmly insisted that the ‘new atheists’ are right in the claims that they make about the rationality of those given over to the excesses of militant Islam and evangelical Christianity. Here, once more, is Harris (2005:236)

"There are days when almost every headline in the morning papers attests to the social costs of religious faith, and the nightly news seems miraculously broadcast from the fourteenth century. One spectacle of religious hysteria follows fast upon the next. Sanctimonious eruptions announcing the death of the pope (a man who actively opposed condom use in sub-Saharan Africa and shielded frocked child molesters from secular justice) are soon followed by other outbursts of religious lunacy. At the time of writing, Muslims in several countries are rioting over a report that US interrogators desecrated a copy of the Koran. … Such perfect visions of unreason have been punctuated by the more ordinary trespasses of faith: daily reports of pious massacres in Iraq, of evangelical ravings about the evils of a secular judiciary, of widespread religious coercion in the US Air Force, or efforts in at least twenty states to redefine science to include supernatural explanations of the origin of life, of devout pharmacists refusing to fill prescriptions for birth control, of movie theatres refusing to show documentaries that report the actual age of the earth, and on and on and onward … to the fifteenth century."

Can’t we at least agree with the ‘new atheists’ that the beliefs of evangelical Christians—concerning the evils of a secular judiciary, or the desirability of the establishment of theocracy, or the age of the earth, or the impossibility of Darwinian evolution, or the infallibility of literally interpreted scripture, or the harm done by the distribution of condoms in sub-Saharan Africa, or …—are plainly not of a kind that could be entertained by any reasonable person, given ordinary standards of reasonableness in the formation and maintenance of beliefs?

One difficulty that we face here is that we need to have some way of taking into account considerations about the information (or evidence) that is available to people. Given that it is plainly a rational strategy for young children to believe what they are told by their parents, it is not hard to understand how it can be that intelligent people come to have radically false beliefs even though they are more or less reasonable in the way that they form those beliefs. Furthermore, it may well be unreasonable to expect people to be able to overcome those radically false beliefs unless they are open to instruction by experts with access to information and evidence to which those people have not hitherto been exposed. Thus, rather than insist that the targeted beliefs of evangelical Christians are irrational, we might well do better to insist that they are rather the product of ignorance: what is required, for the correction of these beliefs, is not that the people in question become more rational, but rather that they become better informed. Or, better still, we should say that, while the targeted beliefs of evangelical Christians may often be the product of unreasonableness, they may also be the products of ignorance (or of some mixture of unreasonableness and ignorance).

Suppose, then, that we formulate our question in the following way: are the targeted beliefs of evangelical Christians of a kind that can be reasonably entertained by reasonable, thoughtful, reflective, well-educated and well-informed people, given ordinary standards of reasonableness, thoughtfulness, etc? It is clear that Dawkins, Harris, Hitchens, and the other ‘new atheists’ will say ‘No!’ The question that I wish to take up in this paper is whether we should agree with the ‘new atheists’ on this point. In particular, I am interested in what we should say about well-established and well-regarded professional philosophers—people like Robert Koons, and William Lane Craig—who hold at least some of the targeted views. But, before we turn to a discussion of what we should say about these professional philosophers, we need to say something about their intellectual background, the wider cultural climate, and the beliefs that they in fact hold.
Many recent semantic theories have involved explicit acceptance of the following two theses: 1. DIRECT REFERENCE (DR): The utterance of a simple sentence containing names or demonstratives normally expresses a "singular proposition" -- a... more
Many recent semantic theories have involved explicit acceptance of the following two theses:

1. DIRECT REFERENCE (DR): The utterance of a simple sentence containing names or demonstratives normally expresses a "singular proposition" -- a proposition that contains as constituents the individuals referred to, and not any descriptions or conditions on them.

2. SEMANTIC INNOCENCE (SI): The utterances of the embedded sentences in belief reports express just the propositions they would if not embedded, and these propositions are the contents of the ascribed beliefs.

Such theories face a well-known difficulty: they seem to conflict with a third, and apparently obviously true, thesis:

3. OPACITY (O): Substitution of coreferring names and demonstratives in belief reports does not always preserve the truth of those reports.

In order to meet this difficulty, two different strategies have been proposed.

A. Conventional Implicature: The first suggestion, adopted by Salmon and Soames, is to claim that O is false: despite appearances, the substitution of coreferring names and demonstratives in belief reports does preserve the truth of those reports.

Now, of course, this suggestion leaves us with a puzzle, namely: why do we ordinarily suppose, and speak as if it were the case, that O is true? Here, Salmon and Soames suggest that belief reports carry conventional (or generalised) implicatures which can change under substitution of coreferring embedded names and demonstratives.
Neither Soames nor Salmon has given any details about the nature of these implicatures. However, they do say that what gets implicated is information about the mode of presentation under which a singular proposition is grasped by a subject. Consequently, it seems hard to resist the suggestion that what they require is a compositional theory according to which implicated modes of presentation associated with sentences are composed of implicated modes of presentation associated with the words which make up those sentences.

But then the question naturally arises why it should not be supposed that these allegedly implicated modes of presentation are actually part of the semantic content of belief reports. Given that we need a theory which associates modes of presentation with words, won’t all of this talk about implicated modes of presentation be just a pointless and unmotivated complication in the theory?

B. Unarticulated constituents: The second suggestion, adopted by Crimmins and Perry, is to deny that the theses DR, SI, and O are inconsistent. This response seems most unpromising. Consider the following quasi-logical principle which connects the notions of truth and semantic value:

4. FUNDAMENTAL SEMANTIC PRINCIPLE (FSP): If the substitution of expression E1 for expression E2 in sentence S (in context C) leads to a change in (literal) truth-value, then this change is due to the semantic values of E1 and E2 (in context C)

I take it that this is more or less a definition of what it is to be a semantic value: semantic values are whatever it is that words contribute to the determination of the literal truth-values of sentences in which they occur (upon particular occasions of utterance, or more generally, of tokening of those sentences). Moreover, I take it that it is obvious that FSP entails that if the substitution of a name or demonstrative E1 for a coreferring name or demonstrative E2 in a sentence S (in a context C) leads to a change in the literal truth-value of S, then it follows -- contrary to DR, or SI, or both -- that E1 and E2 do not have the same semantic content.

Not surprisingly, Crimmins and Perry are at least tacitly committed to the rejection of FSP. In their words, their view is as follows:

"It is very common in natural languages for a statement to exploit unarticulated constituents. When we consider the conditions under which such a statement is true, we find it expresses a proposition which has more constituents in it than can be traced to expressions in the sentence that was spoken. Each constituent of the content that is not itself the content of some expression in the sentence, is an unarticulated constituent of the content of the statement. ... The important principle to be learned is that a change in the wording can precipitate a change in propositional constituents, even when the words do not stand for constituents."

In other words, the "important principle" to which Crimmins and Perry wish to draw attention is that the substitution of an expression E1 for an expression E2 in a sentence S (in a context C) can lead to a change in the semantic content of the sentence S (in the context C) even though E1 and E2 have exactly the same semantic content -- and this is the denial of FSP.

In order to illustrate the notion of "unarticulated constituency", Crimmins and Perry consider the example of utterances of:

(1) It's raining.

As they note, if one utters (1), one will be understood to be claiming that it is raining at the time of one's utterance at some place which is indicated by features of the context of utterance. (Often this place will be the place of one's utterance, but it needn't be.) Moreover, there is no (surface) expression in (1) which has the place in question as its content.

However, there are two points to note about this example. Firstly, this example does nothing to support the principle that a change in wording can precipitate a change in propositional content even when the words do not stand for the constituents. Rather, this example supports the principle that there can be a change in propositional constituents when there is no change in wording even in the case of sentences which contain no indexical expressions. (In other words: sentences can exhibit an indexicality which is not derived from the indexicality of their component expressions.)

Secondly, and more importantly, the genuine principle which can be derived from cases like (1) does nothing to support the account of belief reports which Crimmins and Perry wish to defend. In their view, the semantic contents of

(2a) Scott believes that Hesperus rises in the morning.

and

(2b) Scott believes that Phosphorus rises in the morning.

in a given context C, may differ because the context contributes different unarticulated constituents to (2a) and (2b). But it is incredible to suppose that cases like (1) lend any credence to such a view. In the case of (1), there is no word which can plausibly be connected to the place which (allegedly) forms part of the semantic content of (1). However, in the case of (2a) and (2b) there are obvious candidate words -- namely, "Hesperus" and "Phosphorus" -- which could be semantically associated with the (allegedly) unarticulated constituents of the semantic contents of (2a) and (2b). So why suppose that these constituents of the semantic contents of (2a) and (2b) are not (parts of) the semantic contents of the words "Hesperus" and "Phosphorus"?

This mystery is deepened when we note that Crimmins and Perry claim that "the whole utterance, the context and the words uttered, are relevant to identifying the unarticulated constituent". The point in the first example seemed to be that, since there is no word in (1) which could have the place in question as its semantic value, it is necessary to suppose that the place in question is an unarticulated semantic constituent; but now we are told that in (2a) and (2b) the words "Hesperus" and "Phosphorus" are "relevant to identifying" certain constituents of the semantic contents of (2a) and (2b), and yet that these words can't have those constituents as (parts of) their semantic values.

This time, I take it that the obvious question to ask is: Why not suppose that the semantic constituents which Crimmins and Perry claim are unarticulated semantic constituents in propositional attitude reports are actually (parts of) the semantic contents of the names and demonstratives which appear in those reports? Isn't all this talk of unarticulated constituents of belief reports a pointless and unmotivated complication in the theory?

In sum, then, the crucial question for the two types of semantic theory which I have discussed centres on the thesis of semantic innocence. Crimmins and Perry claim that it is "well-motivated by many considerations in the philosophy of language". However, I cannot see that this is so; rather, it seems to me that this thesis is completely unmotivated, and that, in virtue of the above considerations, it is obvious that semantic theory would be better off without it. The point of the rest of this note is to explain why this is so.
Nathan Salmon (1986a)(1986b)(1989) and Scott Soames (1987)(1988) defend a neo-Russellian account of the semantics of propositional attitude ascriptions in which senses and modes of presentation play no part. They also defend an... more
Nathan Salmon (1986a)(1986b)(1989) and Scott Soames (1987)(1988) defend a neo-Russellian account of the semantics of propositional attitude ascriptions in which senses and modes of presentation play no part. They also defend an under-developed pragmatic theory in which senses and modes of presentation do play an important role.

Graeme Forbes (1987a)(1987b) defends a neo-Fregean account of the semantics of propositional attitude ascriptions in which senses and modes of presentation play a crucial part. He contends that the view which is defended by Salmon and Soames is not really an alternative to the view which he defends. (Hence, I suppose, his view is not an alternative to their view, either.)

There are two issues which I wish to take up. First, in sections I-V, I explore the question whether there are really any important differences between the view defended by Salmon and Soames, and the view defended by Forbes. The conclusion which I defend is that there is very little which distinguishes between them.
Second, in sections VI-X, I take up the somewhat neglected question of how theories of propositional attitude ascriptions in the Frege-Russell tradition ought to construe the notions of sense and mode of presentation.
This dissertation presents a semantic theory for sentences of the form "S believes that A is F"--where "S" is a singular term which denotes a person, "A" is a non-empty proper name, and "F" is an unstructured predicate--according to which... more
This dissertation presents a semantic theory for sentences of the form "S believes that A is F"--where "S" is a singular term which denotes a person, "A" is a non-empty proper name, and "F" is an unstructured predicate--according to which these sentences express relations between the person which "S" denotes and a quadruple which consists of: (i) the entity denoted by "A"; (ii) a mode of presentation of the entity denoted by "A" to the person denoted by "S"; (iii) the property which is denoted by "F"; and (iv) a mode of presentation of the property denoted by "F" to the person denoted by "S". The major innovation in the theory is that the modes of presentation mentioned in (ii) and (iv) are claimed to be associated with the name "A" and the predicate "F" in a context-dependent way. This feature enables the theory to avoid Saul Kripke's well-known objections to theories in which modes of presentation are taken to be context-independent entities.

Furthermore, the inclusion of the mode of presentation of the property denoted by "F" (in addition to the mode of presentation of the entity denoted by "A") allows the theory to provide a solution to Stephen Schiffer's recent objections to "propositional" analysis of belief ascriptions. The dissertation contains extensions of this basic analysis to cover a wide range of more complex kinds of sentences (e.g. sentences in which "A" is replaced by a demonstrative, a pronoun, a definite description, an empty proper name; sentences in which "believes" is replaced by a variety of other "attitude" verbs; etc.) Finally, this extended theory is applied to a wide range of well-known philosophical puzzles--e.g. the analysis of attitudes de se, the analysis of (alleged cases of) the contingent a priori, Kripke's puzzle about belief, linguistic arguments for eliminative materialism, etc.--and is shown to provide various insightful analyses.
Critical Review of Weisberger (1999)
Critical Review of Craig and Smith (1993)
Review of Laurence and MacDonald (1998)
Review of Nagel (2010)
This is an enthusiastic review of Helge Kragh's very nice book on the history of the development of cosmology in the twentieth century.
Review of Dale Jacquette's 1996 book on Meinongian Logic.
Review of *The Science of God* by Gerald Schroeder, 1997
Review of God and Design, edited by Neil Manson, 2003
Extended critical review of *Logic and Theism*, by Jordan Howard Sobel, 2004
This is an enthusiastic review of Helge Kragh's very nice book on the history of the development of cosmology in the twentieth century.
Review of John Earman's fabulous *Bangs, Crunches, Shrieks, Whispers*.
A stellar line-up of leading philosophers from around the world offer new treatments of a topic which has long been central to philosophical debate, and in which there has recently been a surge of interest. The a priori is the category of... more
A stellar line-up of leading philosophers from around the world offer new treatments of a topic which has long been central to philosophical debate, and in which there has recently been a surge of interest. The a priori is the category of knowledge that is supposed to be independent of experience. The contributors offer a variety of approaches to the a priori and examine its role in different areas of philosophical inquiry. The editors' introduction offers an ideal way into the discussions.
ABSTRACT Book Information Naturalism: A Critical Analysis. Edited by William Lane Craig and J.P. Moreland. Routledge. London. 2000. Pp. xv + 286. £60.00.
... I won't mention Moser's essay again.) Part II contains essays by Stephen Davis (`The Ontological Argument'); William Craig (`The Cosmological Argument'); Robin Collins (`The Teleological Argument'); Paul Copan... more
... I won't mention Moser's essay again.) Part II contains essays by Stephen Davis (`The Ontological Argument'); William Craig (`The Cosmological Argument'); Robin Collins (`The Teleological Argument'); Paul Copan (`The Moral Argument'); Doug Geivett (`The Evidential Value of ...
ABSTRACT Kevin Harrelson's book commences with the following words: This book provides a philosophical analysis of the several debates concerning the "ontological argument" from the middle of the seventeenth to... more
ABSTRACT Kevin Harrelson's book commences with the following words: This book provides a philosophical analysis of the several debates concerning the "ontological argument" from the middle of the seventeenth to the beginning of the nineteenth century. My aim in writing it was twofold. First, I wished to provide a detailed and comprehensive account of the history of these debates, which I perceived to be lacking in the scholarly literature. Second, I wanted also to pursue a more philosophically interesting question concerning the apparent unassailability of ontological arguments. In pursuit of this latter problem, the driving question that my account addresses is "why has this argument, or kind of argument, been such a constant in otherwise diverse philosophical contexts and periods?" I think that there is no doubt that Harrelson succeeds in the first of these aims. He has, indeed, produced a detailed scholarly account of the history of debates about ontological arguments from the middle of the seventeenth century to the beginning of the nineteenth. His history is engaging and interesting, covering a wide range of authors with diverse philosophical orientations: Descartes, Arnauld, Caterus, Gassendi, Hobbes, Mersenne, More, Geulincx, Cudworth, Locke, Clarke, Malebranche, Huet, Spinoza, Leibniz, Wolff, Baumgarten, Eberhard, Crusius, Kant, Mendelssohn, and Hegel, among others. It seems to me that anyone who works on the treatment of ontological arguments in this period is bound to profit from Harrelson's study. I think that it is less clear that Harrelson succeeds in the second of his aims. Indeed, it is not clear to me that he ever really succeeds in properly clarifying it. While this brief review is hardly the place to set out detailed considerations, perhaps I can point to just one of the difficulties I find in Harrelson's discussion, a difficulty that arises in connection with the various ways in which he talks about "argument," "demonstration," and "proof." Harrelson tells us, for example, that "the ontological arguments of Descartes, Hegel, et al. stand and fall with a fairly well-defined set of metaphysical, psychological and theological claims to which the arguments are wedded" (19), that "acceptance or rejection of [Spinoza's] ontological argument involves the acceptance or rejection of an entire philosophy" (135), and that "Hegel cannot demonstrate the existence of God in any sense of 'demonstrate' that involves convincing someone who initially rejects the conclusion" (220). I wonder whether there is any acceptable sense of 'argument' that can accommodate these—and many similar—claims. There are various places where Harrelson clearly supposes that particular syllogisms are examples of ontological arguments. For instance, he represents the argument from the First Replies as follows: Premise 1: That which we clearly understand to belong to the true and immutable nature, or essence, or form of something, can be truly asserted of that thing. Premise 2: But once we have made a sufficiently careful investigation into what God is, we clearly and distinctly understand that (necessary) existence belongs to his true and immutable nature. Conclusion: Hence we can now truly assert of God that he does exist. But, if—as seems evidently correct—this is the kind of thing that is properly called an "ontological argument," then it is hard to understand how it could be correct to say that it is the kind of thing that could "stand and fall with a fairly well-defined set of metaphysical, psychological and theological claims," etc. Perhaps it is true that the premises of this argument stand and fall with a fairly well-defined set of metaphysical, psychological, and theological claims, etc.—but the same is no less true of the following argument: Premise: God exists. Conclusion: God exists. What is missing here—and what I find myself unable to supply—is some sense in which the argumentative virtues of things like the First Replies syllogism depend upon an entire philosophy, stand and fall with particular metaphysical, psychological, and theological claims, and so forth. Surely, one is tempted to think, if the premises of an argument "stand and fall with a fairly well-defined set of metaphysical, psychological and theological claims," then it is plainly not the case that that argument is "apparently unassailable"—certainly not if, as in the...
... Horaires d'ouverture du magasin: Du mardi au vendredi, de 09h30 à 18h30 sans interruption Le Samedi, de 10h00 à 18h00 sans interruption. Plus d'un million de titres à notre catalogue ! »Afficher votre panier«. ... 14.... more
... Horaires d'ouverture du magasin: Du mardi au vendredi, de 09h30 à 18h30 sans interruption Le Samedi, de 10h00 à 18h00 sans interruption. Plus d'un million de titres à notre catalogue ! »Afficher votre panier«. ... 14. O'Shaughnessy on the Stream of Consciousness. 15. ...
ABSTRACT Bruce Langtry's ‘God, the Best and Evil’ is a fine contribution to the literature. Here, I review the contents of the book, and then provide some critical remarks that, as fas as I know, have not been made elsewhere. In... more
ABSTRACT Bruce Langtry's ‘God, the Best and Evil’ is a fine contribution to the literature. Here, I review the contents of the book, and then provide some critical remarks that, as fas as I know, have not been made elsewhere. In particular, I argue that his criticism of my formulations of logical arguments from evil (in my Arguing about Gods) is unsuccessful.
... Natural Kinds 311 JOHN DUPRE 47 Newton 320 RICHARD S. YVESTI ALL 48 Observation and Theory 325 PETER ACHINSTEIN 49 ... 61 Russell 408 PAUL J. HAGER 62 Scientific Change 413 DUDLEY SHAPERE 63 Scientific Methodology 423 GARY GUTTING 64... more
... Natural Kinds 311 JOHN DUPRE 47 Newton 320 RICHARD S. YVESTI ALL 48 Observation and Theory 325 PETER ACHINSTEIN 49 ... 61 Russell 408 PAUL J. HAGER 62 Scientific Change 413 DUDLEY SHAPERE 63 Scientific Methodology 423 GARY GUTTING 64 Simplicity 4 ...
Among challenges to Molinism, the challenge posed by divine prophecy of human free action has received insufficient attention. We argue that this challenge is a significant addition to the array of challenges that confront Molinism.
In this paper we describe a simple software system that allows students to practise their critical thinking skills by constructing argument maps of natural language arguments. As the students construct their maps of an argument, the... more
In this paper we describe a simple software system that allows students to practise their critical thinking skills by constructing argument maps of natural language arguments. As the students construct their maps of an argument, the system provides automatic, real time feedback ...
Philosophy in Australia and New Zealand has for some time now been experiencing something of a ‘golden age’. The richness of Australasia’s philosophical past, though less well known, should also not be forgotten. Australasian philosophy,... more
Philosophy in Australia and New Zealand has for some time now been experiencing something of a ‘golden age’. The richness of Australasia’s philosophical past, though less well known, should also not be forgotten. Australasian philosophy, although heavily indebted to overseas trends, includes much distinctive and highly original work. The Companion contains a wide range of articles contributed by prominent philosophers and scholars, and includes biographical essays on selected philosophers, topics and controversies, as well as shorter entries on associations, research centres, departments, journals, pedagogy, and international links. Important philosophical contributions made by those working outside of the academy are also included, along with philosophy’s recent inroads into the wider community – in primary and secondary schools, community-based forums and ‘philosophy cafés’. A Companion to Philosophy in Australia and New Zealand will provide scholars and the wider community, in Australia, New Zealand and beyond, with a greater appreciation of the philosophical heritage of this region, and will be a standard work of reference for many years to come. Note that an electronic version of this book is freely available from the publishers: http://www.publishing.monash.edu/books/cpanz.html
Among challenges to Molinism, the challenge posed by divine prophecy of human free action has received insufficient attention. We argue that this challenge is a significant addition to the array of challenges that confront Molinism.
Page 248. CHAPTER 11 God and Infinity: Directions for Future Research Graham Oppy Philosophical investigation–in particular, metaphysical investigation–is rarely advanced through the consultation of dictionaries. In the present ...