an essay on free will

  • Politics & Social Sciences

Kindle app logo image

Download the free Kindle app and start reading Kindle books instantly on your smartphone, tablet, or computer - no Kindle device required .

Read instantly on your browser with Kindle for Web.

Using your mobile phone camera - scan the code below and download the Kindle app.

QR code to download the Kindle App

Image Unavailable

An Essay on Free Will

  • To view this video download Flash Player

Follow the author

Peter Van Inwagen

An Essay on Free Will y First edition

  • ISBN-10 8198249249
  • ISBN-13 978-0198246244
  • Edition y First edition
  • Publisher Oxford University Press
  • Publication date August 25, 1983
  • Language English
  • Print length 272 pages
  • See all details

Amazon First Reads | Editors' picks at exclusive prices

Similar items that may deliver to you quickly

Four Views on Free Will

Product details

  • ASIN ‏ : ‎ 0198246242
  • Publisher ‏ : ‎ Oxford University Press; y First edition (August 25, 1983)
  • Language ‏ : ‎ English
  • Hardcover ‏ : ‎ 272 pages
  • ISBN-10 ‏ : ‎ 8198249249
  • ISBN-13 ‏ : ‎ 978-0198246244
  • Item Weight ‏ : ‎ 8.1 ounces
  • #1,722 in Free Will & Determinism Philosophy
  • #12,727 in Philosophy (Books)
  • #208,792 in Unknown

About the author

Peter van inwagen.

Discover more of the author’s books, see similar authors, read author blogs and more

Customer reviews

Customer Reviews, including Product Star Ratings help customers to learn more about the product and decide whether it is the right product for them.

To calculate the overall star rating and percentage breakdown by star, we don’t use a simple average. Instead, our system considers things like how recent a review is and if the reviewer bought the item on Amazon. It also analyzed reviews to verify trustworthiness.

  • Sort reviews by Top reviews Most recent Top reviews

An essay on free will

By peter van inwagen.

  • 2 Want to read
  • 0 Currently reading
  • 0 Have read

An essay on free will by Peter Van Inwagen

My Reading Lists:

Use this Work

Create a new list

My book notes.

My private notes about this edition:

Download Options

Check nearby libraries

  • Library.link

Buy this book

This edition doesn't have a description yet. Can you add one ?

Previews available in: English

Showing 3 featured editions. View all 3 editions?

Add another edition?

Book Details

Published in.

Oxford [Oxfordshire], New York

Edition Notes

Includes bibliographical references and index.

Classifications

The physical object, community reviews (0).

  • Created April 1, 2008
  • 14 revisions

Wikipedia citation

Copy and paste this code into your Wikipedia page. Need help ?

SEP home page

  • Table of Contents
  • Random Entry
  • Chronological
  • Editorial Information
  • About the SEP
  • Editorial Board
  • How to Cite the SEP
  • Special Characters
  • Advanced Tools
  • Support the SEP
  • PDFs for SEP Friends
  • Make a Donation
  • SEPIA for Libraries
  • Entry Contents

Bibliography

Academic tools.

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

The term “free will” has emerged over the past two millennia as the canonical designator for a significant kind of control over one’s actions. Questions concerning the nature and existence of this kind of control (e.g., does it require and do we have the freedom to do otherwise or the power of self-determination?), and what its true significance is (is it necessary for moral responsibility or human dignity?) have been taken up in every period of Western philosophy and by many of the most important philosophical figures, such as Plato, Aristotle, Augustine, Aquinas, Descartes, and Kant. (We cannot undertake here a review of related discussions in other philosophical traditions. For a start, the reader may consult Marchal and Wenzel 2017 and Chakrabarti 2017 for overviews of thought on free will, broadly construed, in Chinese and Indian philosophical traditions, respectively.) In this way, it should be clear that disputes about free will ineluctably involve disputes about metaphysics and ethics. In ferreting out the kind of control at stake in free will, we are forced to consider questions about (among others) causation, laws of nature, time, substance, ontological reduction vs emergence, the relationship of causal and reasons-based explanations, the nature of motivation and more generally of human persons. In assessing the significance of free will, we are forced to consider questions about (among others) rightness and wrongness, good and evil, virtue and vice, blame and praise, reward and punishment, and desert. The topic of free will also gives rise to purely empirical questions that are beginning to be explored in the human sciences: do we have it, and to what degree?

Here is an overview of what follows. In Section 1 , we acquaint the reader with some central historical contributions to our understanding of free will. (As nearly every major and minor figure had something to say about it, we cannot begin to cover them all.) As with contributions to many other foundational topics, these ideas are not of ‘merely historical interest’: present-day philosophers continue to find themselves drawn back to certain thinkers as they freshly engage their contemporaries. In Section 2 , we map the complex architecture of the contemporary discussion of the nature of free will by dividing it into five subtopics: its relation to moral responsibility; the proper analysis of the freedom to do otherwise; a powerful, recent argument that the freedom to do otherwise (at least in one important sense) is not necessary for moral responsibility; ‘compatibilist’ accounts of sourcehood or self-determination; and ‘incompatibilist’ or ‘libertarian’ accounts of source and self-determination. In Section 3 , we consider arguments from experience, a priori reflection, and various scientific findings and theories for and against the thesis that human beings have free will, along with the related question of whether it is reasonable to believe that we have it. Finally, in Section 4 , we survey the long-debated questions involving free will that arise in classical theistic metaphysics.

1.1 Ancient and Medieval Period

1.2 modern period and twentieth century, 2.1 free will and moral responsibility, 2.2 the freedom to do otherwise, 2.3 freedom to do otherwise vs. sourcehood accounts, 2.4 compatibilist accounts of sourcehood, 2.5 libertarian accounts of sourcehood, 3.1 arguments against the reality of free will, 3.2 arguments for the reality of free will, 4.1 free will and god’s power, knowledge, and goodness, 4.2 god’s freedom, other internet resources, related entries, 1. major historical contributions.

One finds scholarly debate on the ‘origin’ of the notion of free will in Western philosophy. (See, e.g., Dihle (1982) and, in response Frede (2011), with Dihle finding it in St. Augustine (354–430 CE) and Frede in the Stoic Epictetus (c. 55–c. 135 CE).) But this debate presupposes a fairly particular and highly conceptualized concept of free will, with Dihle’s later ‘origin’ reflecting his having a yet more particular concept in view than Frede. If, instead, we look more generally for philosophical reflection on choice-directed control over one’s own actions, then we find significant discussion in Plato and Aristotle (cf. Irwin 1992). Indeed, on this matter, as with so many other major philosophical issues, Plato and Aristotle give importantly different emphases that inform much subsequent thought.

In Book IV of The Republic , Plato posits rational, spirited, and appetitive aspects to the human soul. The wise person strives for inner ‘justice’, a condition in which each part of the soul plays its proper role—reason as the guide, the spirited nature as the ally of reason, exhorting oneself to do what reason deems proper, and the passions as subjugated to the determinations of reason. In the absence of justice, the individual is enslaved to the passions. Hence, freedom for Plato is a kind of self-mastery, attained by developing the virtues of wisdom, courage, and temperance, resulting in one’s liberation from the tyranny of base desires and acquisition of a more accurate understanding and resolute pursuit of the Good (Hecht 2014).

While Aristotle shares with Plato a concern for cultivating virtues, he gives greater theoretical attention to the role of choice in initiating individual actions which, over time, result in habits, for good or ill. In Book III of the Nicomachean Ethics , Aristotle says that, unlike nonrational agents, we have the power to do or not to do, and much of what we do is voluntary, such that its origin is ‘in us’ and we are ‘aware of the particular circumstances of the action’. Furthermore, mature humans make choices after deliberating about different available means to our ends, drawing on rational principles of action. Choose consistently well (poorly), and a virtuous (vicious) character will form over time, and it is in our power to be either virtuous or vicious.

A question that Aristotle seems to recognize, while not satisfactorily answering, is whether the choice an individual makes on any given occasion is wholly determined by his internal state—perception of his circumstances and his relevant beliefs, desires, and general character dispositions (wherever on the continuum between virtue and vice he may be)—and external circumstances. He says that “the man is the father of his actions as of children”—that is, a person’s character shapes how she acts. One might worry that this seems to entail that the person could not have done otherwise—at the moment of choice, she has no control over what her present character is—and so she is not responsible for choosing as she does. Aristotle responds by contending that her present character is partly a result of previous choices she made. While this claim is plausible enough, it seems to ‘pass the buck’, since ‘the man is the father’ of those earlier choices and actions, too.

We note just a few contributions of the subsequent centuries of the Hellenistic era. (See Bobzien 1998.) This period was dominated by debates between Epicureans, Stoics, and the Academic Skeptics, and as it concerned freedom of the will, the debate centered on the place of determinism or of fate in governing human actions and lives. The Stoics and the Epicureans believed that all ordinary things, human souls included, are corporeal and governed by natural laws or principles. Stoics believed that all human choice and behavior was causally determined, but held that this was compatible with our actions being ‘up to us’. Chrysippus ably defended this position by contending that your actions are ‘up to you’ when they come about ‘through you’—when the determining factors of your action are not external circumstances compelling you to act as you do but are instead your own choices grounded in your perception of the options before you. Hence, for moral responsibility, the issue is not whether one’s choices are determined (they are) but in what manner they are determined. Epicurus and his followers had a more mechanistic conception of bodily action than the Stoics. They held that all things (human soul included) are constituted by atoms, whose law-governed behavior fixes the behavior of everything made of such atoms. But they rejected determinism by supposing that atoms, though law-governed, are susceptible to slight ‘swerves’ or departures from the usual paths. Epicurus has often been understood as seeking to ground the freedom of human willings in such indeterministic swerves, but this is a matter of controversy. If this understanding of his aim is correct, how he thought that this scheme might work in detail is not known. (What little we know about his views in this matter stem chiefly from the account given in his follower Lucretius’s six-book poem, On the Nature of Things . See Bobzien 2000 for discussion.)

A final notable figure of this period was Alexander of Aphrodisias , the most important Peripatetic commentator on Aristotle. In his On Fate , Alexander sharply criticizes the positions of the Stoics. He goes on to resolve the ambiguity in Aristotle on the question of the determining nature of character on individual choices by maintaining that, given all such shaping factors, it remains open to the person when she acts freely to do or not to do what she in fact does. Many scholars see Alexander as the first unambiguously ‘libertarian’ theorist of the will (for more information about such theories see section 2 below).

Augustine (354–430) is the central bridge between the ancient and medieval eras of philosophy. His mature thinking about the will was influenced by his early encounter with late classical Neoplatonist thought, which is then transformed by the theological views he embraces in his adult Christian conversion, famously recounted in his Confessions . In that work and in the earlier On the Free Choice of the Will , Augustine struggles to draw together into a coherent whole the doctrines that creaturely misuse of freedom, not God, is the source of evil in the world and that the human will has been corrupted through the ‘fall’ from grace of the earliest human beings, necessitating a salvation that is attained entirely through the actions of God, even as it requires, constitutively, an individual’s willed response of faith. The details of Augustine’s positive account remain a matter of controversy. He clearly affirms that the will is by its nature a self-determining power—no powers external to it determine its choice—and that this feature is the basis of its freedom. But he does not explicitly rule out the will’s being internally determined by psychological factors, as Chrysippus held, and Augustine had theological reasons that might favor (as well as others that would oppose) the thesis that all things are determined in some manner by God. Scholars divide on whether Augustine was a libertarian or instead a kind of compatibilist with respect to metaphysical freedom. (Macdonald 1999 and Stump 2006 argue the former, Baker 2003 and Couenhoven 2007 the latter.) It is clear, however, that Augustine thought that we are powerfully shaped by wrongly-ordered desires that can make it impossible for us to wholeheartedly will ends contrary to those desires, for a sustained period of time. This condition entails an absence of something more valuable, ‘true freedom’, in which our wills are aligned with the Good, a freedom that can be attained only by a transformative operation of divine grace. This latter, psychological conception of freedom of will clearly echoes Plato’s notion of the soul’s (possible) inner justice.

Thomas Aquinas (1225–1274) attempted to synthesize major strands of Aristotle’s systematic philosophy with Christian theology, and so Aquinas begins his complex discussion of human action and choice by agreeing with Aristotle that creatures such as ourselves who are endowed with both intellect and will are hardwired to will certain general ends ordered to the most general goal of goodness. Will is rational desire: we cannot move towards that which does not appear to us at the time to be good. Freedom enters the picture when we consider various means to these ends and move ourselves to activity in pursuit of certain of them. Our will is free in that it is not fixed by nature on any particular means, and they generally do not appear to us either as unqualifiedly good or as uniquely satisfying the end we wish to fulfill. Furthermore, what appears to us to be good can vary widely—even, over time, intra-personally. So much is consistent with saying that in a given total circumstance (including one’s present beliefs and desires), one is necessitated to will as one does. For this reason, some commentators have taken Aquinas to be a kind of compatibilist concerning freedom and causal or theological determinism. In his most extended defense of the thesis that the will is not ‘compelled’ ( DM 6), Aquinas notes three ways that the will might reject an option it sees as attractive: (i) it finds another option more attractive, (ii) it comes to think of some circumstance rendering an alternative more favorable “by some chance circumstance, external or internal”, and (iii) the person is momentarily disposed to find an alternative attractive by virtue of a non-innate state that is subject to the will (e.g., being angry vs being at peace). The first consideration is clearly consistent with compatibilism. The second at best points to a kind of contingency that is not grounded in the activity of the will itself. And one wanting to read Aquinas as a libertarian might worry that his third consideration just passes the buck: even if we do sometimes have an ability to directly modify perception-coloring states such as moods, Aquinas’s account of will as rational desire seems to indicate that we will do so only if it seems to us on balance to be good to do so. Those who read Aquinas as a libertarian point to the following further remark in this text: “Will itself can interfere with the process [of some cause’s moving the will] either by refusing to consider what attracts it to will or by considering its opposite: namely, that there is a bad side to what is being proposed…” (Reply to 15; see also DV 24.2). For discussion, see MacDonald (1998), Stump (2003, ch. 9) and especially Hoffman & Michon (2017), which offers the most comprehensive analysis of relevant texts to date.

John Duns Scotus (1265/66–1308) was the stoutest defender in the medieval era of a strongly libertarian conception of the will, maintaining on introspective grounds that will by its very nature is such that “nothing other than the will is the total cause” of its activity ( QAM ). Indeed, he held the unusual view that not only up to but at the very instant that one is willing X , it is possible for one to will Y or at least not to will X . (He articulates this view through the puzzling claim that a single instant of time comprises two ‘instants of nature’, at the first but not the second of which alternative possibilities are preserved.) In opposition to Aquinas and other medieval Aristotelians, Scotus maintained that a precondition of our freedom is that there are two fundamentally distinct ways things can seem good to us: as practically advantageous to us or as according with justice. Contrary to some popular accounts, however, Scotus allowed that the scope of available alternatives for a person will be more or less constricted. He grants that we are not capable of willing something in which we see no good whatsoever, nor of positively repudiating something which appears to us as unqualifiedly good. However, in accordance with his uncompromising position that nothing can be the total cause of the will other than itself, he held that where something does appear to us as unqualifiedly good (perfectly suited both to our advantage and justice)—viz., in the ‘beatific vision’ of God in the afterlife—we still can refrain from willing it. For discussion, see John Duns Scotus, §5.2 .

The problem of free will was an important topic in the modern period, with all the major figures wading into it (Descartes 1641 [1988], 1644 [1988]; Hobbes 1654 [1999], 1656 [1999]; Spinoza 1677 [1992]; Malebranche 1684 [1993]; Leibniz 1686 [1991]; Locke 1690 [1975]; Hume 1740 [1978], 1748 [1975]; Edwards 1754 [1957]; Kant 1781 [1998], 1785 [1998], 1788 [2015]; Reid 1788 [1969]). After less sustained attention in the 19th Century (most notable were Schopenhauer 1841 [1999] and Nietzsche 1886 [1966]), it was widely discussed again among early twentieth century philosophers (Moore 1912; Hobart 1934; Schlick 1939; Nowell-Smith 1948, 1954; Campbell 1951; Ayer 1954; Smart 1961). The centrality of the problem of free will to the various projects of early modern philosophers can be traced to two widely, though not universally, shared assumptions. The first is that without belief in free will, there would be little reason for us to act morally. More carefully, it was widely assumed that belief in an afterlife in which a just God rewards and punishes us according to our right or wrong use of free will was key to motivating us to be moral (Russell 2008, chs. 16–17). Life before death affords us many examples in which vice is better rewarded than virtue and so without knowledge of a final judgment in the afterlife, we would have little reason to pursue virtue and justice when they depart from self-interest. And without free will there can be no final judgement.

The second widely shared assumption is that free will seems difficult to reconcile with what we know about the world. While this assumption is shared by the majority of early modern philosophers, what specifically it is about the world that seems to conflict with freedom differs from philosopher to philosopher. For some, the worry is primarily theological. How can we make sense of contingency and freedom in a world determined by a God who must choose the best possible world to create? For some, the worry was primarily metaphysical. The principle of sufficient reason—roughly, the idea that every event must have a reason or cause—was a cornerstone of Leibniz’s and Spinoza’s metaphysics. How does contingency and freedom fit into such a world? For some, the worry was primarily scientific (Descartes). Given that a proper understanding of the physical world is one in which all physical objects are governed by deterministic laws of nature, how does contingency and freedom fit into such a world? Of course, for some, all three worries were in play in their work (this is true especially of Leibniz).

Despite many disagreements about how best to solve these worries, there were three claims that were widely, although not universally, agreed upon. The first was that free will has two aspects: the freedom to do otherwise and the power of self-determination. The second is that an adequate account of free will must entail that free agents are morally responsible agents and/or fit subjects for punishment. Ideas about moral responsibility were often a yard stick by which analyses of free will were measured, with critics objecting to an analysis of free will by arguing that agents who satisfied the analysis would not, intuitively, be morally responsible for their actions. The third is that compatibilism—the thesis that free will is compatible with determinism—is true. (Spinoza, Reid, and Kant are the clear exceptions to this, though some also see Descartes as an incompatibilist [Ragland 2006].)

Since a detailed discussion of these philosophers’ accounts of free will would take us too far afield, we want instead to focus on isolating a two-step strategy for defending compatibilism that emerges in the early modern period and continued to exert considerable force into the early twentieth century (and perhaps is still at work today). Advocates of this two-step strategy have come to be known as “classical compatibilists”. The first step was to argue that the contrary of freedom is not determinism but external constraint on doing what one wants to do. For example, Hobbes contends that liberty is “the absence of all the impediments to action that are not contained in the nature and intrinsical quality of the agent” (Hobbes 1654 [1999], 38; cf. Hume 1748 [1975] VIII.1; Edwards 1754 [1957]; Ayer 1954). This idea led many compatibilists, especially the more empiricist-inclined, to develop desire- or preference-based analyses of both the freedom to do otherwise and self-determination. An agent has the freedom to do otherwise than \(\phi\) just in case if she preferred or willed to do otherwise, she would have done otherwise (Hobbes 1654 [1999], 16; Locke 1690 [1975]) II.xx.8; Hume 1748 [1975] VIII.1; Moore 1912; Ayer 1954). The freedom to do otherwise does not require that you are able to act contrary to your strongest motivation but simply that your action be dependent on your strongest motivation in the sense that had you desired something else more strongly, then you would have pursued that alternative end. (We will discuss this analysis in more detail below in section 2.2.) Similarly, an agent self-determines her \(\phi\)-ing just in case \(\phi\) is caused by her strongest desires or preferences at the time of action (Hobbes 1654 [1999]; Locke 1690 [1975]; Edwards 1754 [1957]). (We will discuss this analysis in more detail below in section 2.4.) Given these analyses, determinism seems innocuous to freedom.

The second step was to argue that any attempt to analyze free will in a way that putatively captures a deeper or more robust sense of freedom leads to intractable conundrums. The most important examples of this attempt to capture a deeper sense of freedom in the modern period are Immanuel Kant (1781 [1998], 1785 [1998], 1788 [2015]) and Thomas Reid (1788 [1969]) and in the early twentieth century C. A. Campbell (1951). These philosophers argued that the above compatibilist analyses of the freedom to do otherwise and self-determination are, at best, insufficient for free will, and, at worst, incompatible with it. With respect to the classical compatibilist analysis of the freedom to do otherwise, these critics argued that the freedom to do otherwise requires not just that an agent could have acted differently if he had willed differently, but also that he could have willed differently. Free will requires more than free action. With respect to classical compatibilists’ analysis of self-determination, they argued that self-determination requires that the agent—rather than his desires, preferences, or any other mental state—cause his free choices and actions. Reid explains:

I consider the determination of the will as an effect. This effect must have a cause which had the power to produce it; and the cause must be either the person himself, whose will it is, or some other being…. If the person was the cause of that determination of his own will, he was free in that action, and it is justly imputed to him, whether it be good or bad. But, if another being was the cause of this determination, either producing it immediately, or by means and instruments under his direction, then the determination is the act and deed of that being, and is solely imputed to him. (1788 [1969] IV.i, 265)

Classical compatibilists argued that both claims are incoherent. While it is intelligible to ask whether a man willed to do what he did, it is incoherent to ask whether a man willed to will what he did:

For to ask whether a man is at liberty to will either motion or rest, speaking or silence, which he pleases, is to ask whether a man can will what he wills , or be pleased with what he is pleased with? A question which, I think, needs no answer; and they who make a question of it must suppose one will to determine the acts of another, and another to determine that, and so on in infinitum . (Locke 1690 [1975] II.xx.25; cf. Hobbes 1656 [1999], 72)

In response to libertarians’ claim that self-determination requires that the agent, rather than his motives, cause his actions, it was objected that this removes the agent from the natural causal order, which is clearly unintelligible for human animals (Hobbes 1654 [1999], 38). It is important to recognize that an implication of the second step of the strategy is that free will is not only compatible with determinism but actually requires determinism (cf. Hume 1748 [1975] VIII). This was a widely shared assumption among compatibilists up through the mid-twentieth century.

Spinoza’s Ethics (1677 [1992]) is an important departure from the above dialectic. He endorses a strong form of necessitarianism in which everything is categorically necessary as opposed to the conditional necessity embraced by most compatibilists, and he contends that there is no room in such a world for divine or creaturely free will. Thus, Spinoza is a free will skeptic. Interestingly, Spinoza is also keen to deny that the nonexistence of free will has the dire implications often assumed. As noted above, many in the modern period saw belief in free will and an afterlife in which God rewards the just and punishes the wicked as necessary to motivate us to act morally. According to Spinoza, so far from this being necessary to motivate us to be moral, it actually distorts our pursuit of morality. True moral living, Spinoza thinks, sees virtue as its own reward (Part V, Prop. 42). Moreover, while free will is a chimera, humans are still capable of freedom or self-determination. Such self-determination, which admits of degrees on Spinoza’s view, arises when our emotions are determined by true ideas about the nature of reality. The emotional lives of the free persons are ones in which “we desire nothing but that which must be, nor, in an absolute sense, can we find contentment in anything but truth. And so in so far as we rightly understand these matters, the endeavor of the better part of us is in harmony with the order of the whole of Nature” (Part IV, Appendix). Spinoza is an important forerunner to the many free will skeptics in the twentieth century, a position that continues to attract strong support (see Strawson 1986; Double 1992; Smilansky 2000; Pereboom 2001, 2014; Levy 2011; Waller 2011; Caruso 2012; Vilhauer 2012. For further discussion see the entry skepticism about moral responsibility ).

It is worth observing that in many of these disputes about the nature of free will there is an underlying dispute about the nature of moral responsibility. This is seen clearly in Hobbes (1654 [1999]) and early twentieth century philosophers’ defenses of compatibilism. Underlying the belief that free will is incompatible with determinism is the thought that no one would be morally responsible for any actions in a deterministic world in the sense that no one would deserve blame or punishment. Hobbes responded to this charge in part by endorsing broadly consequentialist justifications of blame and punishment: we are justified in blaming or punishing because these practices deter future harmful actions and/or contribute to reforming the offender (1654 [1999], 24–25; cf. Schlick 1939; Nowell-Smith 1948; Smart 1961). While many, perhaps even most, compatibilists have come to reject this consequentialist approach to moral responsibility in the wake of P. F. Strawson’s 1962 landmark essay ‘Freedom and Resentment’ (though see Vargas (2013) and McGeer (2014) for contemporary defenses of compatibilism that appeal to forward-looking considerations) there is still a general lesson to be learned: disputes about free will are often a function of underlying disputes about the nature and value of moral responsibility.

2. The Nature of Free Will

As should be clear from this short discussion of the history of the idea of free will, free will has traditionally been conceived of as a kind of power to control one’s choices and actions. When an agent exercises free will over her choices and actions, her choices and actions are up to her . But up to her in what sense? As should be clear from our historical survey, two common (and compatible) answers are: (i) up to her in the sense that she is able to choose otherwise, or at minimum that she is able not to choose or act as she does, and (ii) up to her in the sense that she is the source of her action. However, there is widespread controversy both over whether each of these conditions is required for free will and if so, how to understand the kind or sense of freedom to do otherwise or sourcehood that is required. While some seek to resolve these controversies in part by careful articulation of our experiences of deliberation, choice, and action (Nozick 1981, ch. 4; van Inwagen 1983, ch. 1), many seek to resolve these controversies by appealing to the nature of moral responsibility. The idea is that the kind of control or sense of up-to-meness involved in free will is the kind of control or sense of up-to-meness relevant to moral responsibility (Double 1992, 12; Ekstrom 2000, 7–8; Smilansky 2000, 16; Widerker and McKenna 2003, 2; Vargas 2007, 128; Nelkin 2011, 151–52; Levy 2011, 1; Pereboom 2014, 1–2). Indeed, some go so far as to define ‘free will’ as ‘the strongest control condition—whatever that turns out to be—necessary for moral responsibility’ (Wolf 1990, 3–4; Fischer 1994, 3; Mele 2006, 17). Given this connection, we can determine whether the freedom to do otherwise and the power of self-determination are constitutive of free will and, if so, in what sense, by considering what it takes to be a morally responsible agent. On these latter characterizations of free will, understanding free will is inextricably linked to, and perhaps even derivative from, understanding moral responsibility. And even those who demur from this claim regarding conceptual priority typically see a close link between these two ideas. Consequently, to appreciate the current debates surrounding the nature of free will, we need to say something about the nature of moral responsibility.

It is now widely accepted that there are different species of moral responsibility. It is common (though not uncontroversial) to distinguish moral responsibility as answerability from moral responsibility as attributability from moral responsibility as accountability (Watson 1996; Fischer and Tognazzini 2011; Shoemaker 2011. See Smith (2012) for a critique of this taxonomy). These different species of moral responsibility differ along three dimensions: (i) the kind of responses licensed toward the responsible agent, (ii) the nature of the licensing relation, and (iii) the necessary and sufficient conditions for licensing the relevant kind of responses toward the agent. For example, some argue that when an agent is morally responsible in the attributability sense, certain judgments about the agent—such as judgments concerning the virtues and vices of the agent—are fitting , and that the fittingness of such judgments does not depend on whether the agent in question possessed the freedom to do otherwise (cf. Watson 1996).

While keeping this controversy about the nature of moral responsibility firmly in mind (see the entry on moral responsibility for a more detailed discussion of these issues), we think it is fair to say that the most commonly assumed understanding of moral responsibility in the historical and contemporary discussion of the problem of free will is moral responsibility as accountability in something like the following sense:

An agent \(S\) is morally accountable for performing an action \(\phi\) \(=_{df.}\) \(S\) deserves praise if \(\phi\) goes beyond what can be reasonably expected of \(S\) and \(S\) deserves blame if \(\phi\) is morally wrong.

The central notions in this definition are praise , blame , and desert . The majority of contemporary philosophers have followed Strawson (1962) in contending that praising and blaming an agent consist in experiencing (or at least being disposed to experience (cf. Wallace 1994, 70–71)) reactive attitudes or emotions directed toward the agent, such as gratitude, approbation, and pride in the case of praise, and resentment, indignation, and guilt in the case of blame. (See Sher (2006) and Scanlon (2008) for important dissents from this trend. See the entry on blame for a more detailed discussion.) These emotions, in turn, dispose us to act in a variety of ways. For example, blame disposes us to respond with some kind of hostility toward the blameworthy agent, such as verbal rebuke or partial withdrawal of good will. But while these kinds of dispositions are essential to our blaming someone, their manifestation is not: it is possible to blame someone with very little change in attitudes or actions toward the agent. Blaming someone might be immediately followed by forgiveness as an end of the matter.

By ‘desert’, we have in mind what Derk Pereboom has called basic desert :

The desert at issue here is basic in the sense that the agent would deserve to be blamed or praised just because she has performed the action, given an understanding of its moral status, and not, for example, merely by virtue of consequentialist or contractualist considerations. (2014, 2)

As we understand desert, if an agent deserves blame, then we have a strong pro tanto reason to blame him simply in virtue of his being accountable for doing wrong. Importantly, these reasons can be outweighed by other considerations. While an agent may deserve blame, it might, all things considered, be best to forgive him unconditionally instead.

When an agent is morally responsible for doing something wrong, he is blame worthy : he deserves hard treatment marked by resentment and indignation and the actions these emotions dispose us toward, such as censure, rebuke, and ostracism. However, it would seem unfair to treat agents in these ways unless their actions were up to them . Thus, we arrive at the core connection between free will and moral responsibility: agents deserve praise or blame only if their actions are up to them—only if they have free will. Consequently, we can assess analyses of free will by their implications for judgments of moral responsibility. We note that some might reject the claim that free will is necessary for moral responsibility (e.g., Frankfurt 1971; Stump 1988), but even for these theorists an adequate analysis of free will must specify a sufficient condition for the kind of control at play in moral responsibility.

In what follows, we focus our attention on the two most commonly cited features of free will: the freedom to do otherwise and sourcehood. While some seem to think that free will consists exclusively in either the freedom to do otherwise (van Inwagen 2008) or in sourcehood (Zagzebski 2000), many philosophers hold that free will involves both conditions—though philosophers often emphasize one condition over the other depending on their dialectical situation or argumentative purposes (cf. Watson 1987). In what follows, we will describe the most common characterizations of these two conditions.

For most newcomers to the problem of free will, it will seem obvious that an action is up to an agent only if she had the freedom to do otherwise. But what does this freedom come to? The freedom to do otherwise is clearly a modal property of agents, but it is controversial just what species of modality is at stake. It must be more than mere possibility : to have the freedom to do otherwise consists in more than the mere possibility of something else’s happening. A more plausible and widely endorsed understanding claims the relevant modality is ability or power (Locke 1690 [1975], II.xx; Reid 1788 [1969], II.i–ii; D. Locke 1973; Clarke 2009; Vihvelin 2013). But abilities themselves seem to come in different varieties (Lewis 1976; Horgan 1979; van Inwagen 1983, ch. 1; Mele 2003; Clarke 2009; Vihvelin 2013, ch. 1; Franklin 2015; Cyr and Swenson 2019; Hofmann 2022; Whittle 2022), so a claim that an agent has ‘the ability to do otherwise’ is potentially ambiguous or indeterminate; in philosophical discussion, the sense of ability appealed to needs to be spelled out. A satisfactory account of the freedom to do otherwise owes us both an account of the kind of ability in terms of which the freedom to do otherwise is analyzed, and an argument for why this kind of ability (as opposed to some other species) is the one constitutive of the freedom to do otherwise. As we will see, philosophers sometimes leave this second debt unpaid.

The contemporary literature takes its cue from classical compatibilism’s recognized failure to deliver a satisfactory analysis of the freedom to do otherwise. As we saw above, classical compatibilists (Hobbes 1654 [1999], 1656 [1999]; Locke 1690 [1975]; Hume 1740 [1978], 1748 [1975]; Edwards 1754 [1957]; Moore 1912; Schlick 1939; Ayer 1954) sought to analyze the freedom to do otherwise in terms of a simple conditional analysis of ability:

Simple Conditional Analysis: An agent \(S\) has the ability to do otherwise if and only if, were \(S\) to choose to do otherwise, then \(S\) would do otherwise.

Part of the attraction of this analysis is that it obviously reconciles the freedom to do otherwise with determinism. While the truth of determinism entails that one’s action is inevitable given the past and laws of nature, there is nothing about determinism that implies that if one had chosen otherwise, then one would not do otherwise.

There are two problems with the Simple Conditional Analysis . The first is that it is, at best, an analysis of free action, not free will (cf. Reid 1788 [1969]; Chisholm 1966; 1976, ch. 2; Lehrer 1968, 1976). It only tells us when an agent has the ability to do otherwise, not when an agent has the ability to choose to do otherwise. One might be tempted to think that there is an easy fix along the following lines:

Simple Conditional Analysis*: An agent \(S\) has the ability to choose otherwise if and only if, were \(S\) to desire or prefer to choose otherwise, then \(S\) would choose otherwise.

The problem is that we often fail to choose to do things we want to choose, even when it appears that we had the ability to choose otherwise (one might think the same problem attends the original analysis). Suppose that, in deciding how to spend my evening, I have a desire to choose to read and a desire to choose to watch a movie. Suppose that I choose to read. By all appearances, I had the ability to choose to watch a movie. And yet, according to the Simple Conditional Analysis* , I lack this freedom, since the conditional ‘if I were to desire to choose to watch a movie, then I would choose to watch a movie’ is false. I do desire to choose to watch a movie and yet I do not choose to watch a movie. It is unclear how to remedy this problem. On the one hand, we might refine the antecedent by replacing ‘desire’ with ‘strongest desire’ (cf. Hobbes 1654 [1999], 1656 [1999]; Edwards 1754 [1957]). The problem is that this assumes, implausibly, that we always choose what we most strongly desire (for criticisms of this view see Reid 1788 [1969]; Campbell 1951; Wallace 1999; Holton 2009). On the other hand, we might refine the consequent by replacing ‘would choose to do otherwise’ with either ‘would probably choose to do otherwise’ or ‘might choose to do otherwise’. But each of these proposals is also problematic. If ‘probably’ means ‘more likely than not’, then this revised conditional still seems too strong: it seems possible to have the ability to choose otherwise even when one’s so choosing is unlikely. If we opt for ‘might’, then the relevant sense of modality needs to be spelled out.

Even if there are fixes to these problems, there is a yet deeper problem with these analyses. There are some agents who clearly lack the freedom to do otherwise and yet satisfy the conditional at the heart of these analyses. That is, although these agents lack the freedom to do otherwise, it is, for example, true of them that if they chose otherwise, they would do otherwise. Picking up on an argument developed by Keith Lehrer (1968; cf. Campbell 1951; Broad 1952; Chisholm 1966), consider an agoraphobic, Luke, who, when faced with the prospect of entering an open space, is subject not merely to an irresistible desire to refrain from intentionally going outside, but an irresistible desire to refrain from even choosing to go outside. Given Luke’s psychology, there is no possible world in which he suffers from his agoraphobia and chooses to go outside. It may well nevertheless be true that if Luke chose to go outside, then he would have gone outside. After all, any possible world in which he chooses to go outside will be a world in which he no longer suffers (to the same degree) from his agoraphobia, and thus we have no reason to doubt that in those worlds he would go outside as a result of his choosing to go outside. The same kind of counterexample applies with equal force to the conditional ‘if \(S\) desired to choose otherwise, then \(S\) would choose otherwise’.

While simple conditional analyses admirably make clear the species of ability to which they appeal, they fail to show that this species of ability is constitutive of the freedom to do otherwise. Agents need a stronger ability to do otherwise than characterized by such simple conditionals. Some argue that the fundamental source of the above problems is the conditional nature of these analyses (Campbell 1951; Austin 1961; Chisholm 1966; Lehrer 1976; van Inwagen 1983, ch. 4). The sense of ability relevant to the freedom to do otherwise is the ‘all-in sense’—that is, holding everything fixed up to the time of the decision or action—and this sense, so it is argued, can only be captured by a categorical analysis of the ability to do otherwise:

Categorical Analysis: An agent \(S\) has the ability to choose or do otherwise than \(\phi\) at time \(t\) if and only if it was possible, holding fixed everything up to \(t\), that \(S\) choose or do otherwise than \(\phi\) at \(t\).

This analysis gets the right verdict in Luke’s case. He lacks the ability to do otherwise than refrain from choosing to go outside, according to this analysis, because there is no possible world in which he suffers from his agoraphobia and yet chooses to go outside. Unlike the above conditional analyses, the Categorical Analysis requires that we hold fixed Luke’s agoraphobia when considering alternative possibilities.

If the Categorical Analysis is correct, then free will is incompatible with determinism. According to the thesis of determinism, all deterministic possible worlds with the same pasts and laws of nature have the same futures (Lewis 1979; van Inwagen 1983, 3). Suppose John is in deterministic world \(W\) and refrains from raising his hand at time \(t\). Since \(W\) is deterministic, it follows that any possible world \(W^*\) that has the same past and laws up to \(t\) must have the same future, including John’s refraining from raising his hand at \(t\). Therefore, John lacked the ability, and thus freedom, to raise his hand.

This argument, carefully articulated in the late 1960s and early 1970s by Carl Ginet (1966, 1990) and Peter van Inwagen (1975, 1983) and refined in important ways by John Martin Fischer (1994), has come to be known as the Consequence Argument. van Inwagen offers the following informal statement of the argument:

If determinism is true, then our acts are the consequences of the laws of nature and events in the remote past. But it is not up to us what went on before we were born [i.e., we do not have the ability to change the past], and neither is it up to us what the laws of nature are [i.e., we do not have the ability to break the laws of nature]. Therefore, the consequences of these things (including our present acts) are not up to us. (van Inwagen 1983, 16; cf. Fischer 1994, ch. 1)

Like the Simple Conditional Analysis , a virtue of the Categorical Analysis is that it spells out clearly the kind of ability appealed to in its analysis of the freedom to do otherwise, but like the Simple Conditional Analysis , critics have argued that the sense of ability it captures is not the sense at the heart of free will. The objection here, though, is not that the analysis is too permissive or weak, but rather that it is too restrictive or strong.

While there have been numerous different replies along these lines (e.g., Lehrer 1980; Slote 1982; Watson 1986. See the entry on arguments for incompatibilism for a more extensive discussion of and bibliography for the Consequence Argument), the most influential of these objections is due to David Lewis (1981). Lewis contended that van Inwagen’s argument equivocated on ‘is able to break a law of nature’. We can distinguish two senses of ‘is able to break a law of nature’:

(Weak Thesis) I am able to do something such that, if I did it, a law of nature would be broken.

(Strong Thesis) I am able to do something such that, if I did it, it would constitute a law of nature’s being broken or would cause a law of nature to be broken.

If we are committed to the Categorical Analysis , then those desiring to defend compatibilism seem to be committed to the sense of ability in ‘is able to break a law of nature’ along the lines of the strong thesis. Lewis agrees with van Inwagen that it is “incredible” to think humans have such an ability (Lewis 1981, 113), but maintains that compatibilists need only appeal to the ability to break a law of nature in the weak sense. While it is absurd to think that humans are able to do something that is a violation of a law of nature or causes a law of nature to be broken, there is nothing incredible, so Lewis claimed, in thinking that humans are able to do something such that if they did it, a law of nature would be broken. In essence, Lewis is arguing that incompatibilists like van Inwagen have failed to adequately motivate the restrictiveness of the Categorical Analysis .

Some incompatibilists have responded to Lewis by contending that even the weak ability is incredible (van Inwagen 2004). But there is a different and often overlooked problem for Lewis: the weak ability seems to be too weak. Returning to the case of John’s refraining from raising his hand, Lewis maintains that the following three propositions are consistent:

One might think that (ii) and (iii) are incompatible with (i). Consider again Luke, our agoraphobic. Suppose that his agoraphobia affects him in such a way that he will only intentionally go outside if he chooses to go outside, and yet his agoraphobia makes it impossible for him to make this choice. In this case, a necessary condition for Luke’s intentionally going outside is his choosing to go outside. Moreover, Luke is not able to choose or cause himself to choose to go outside. Intuitively, this would seem to imply that Luke lacks the freedom to go outside. But this implication does not follow for Lewis. From the fact that Luke is able to go outside only if he chooses to go outside and the fact that Luke is not able to choose to go outside, it does not follow , on Lewis’s account, that Luke lacks the ability to go outside. Consequently, Lewis’s account fails to explain why Luke lacks the ability to go outside (cf. Speak 2011). (For other important criticisms of Lewis, see Ginet [1990, ch. 5] and Fischer [1994, ch. 4].)

While Lewis may be right that the Categorical Analysis is too restrictive, his argument, all by itself, doesn’t seem to establish this. His argument is successful only if (a) he can provide an alternative analysis of ability that entails that Luke’s agoraphobia robs him of the ability to go outside and (b) does not entail that determinism robs John of the ability to raise his hand (cf. Pendergraft 2010). Lewis must point out a principled difference between these two cases. As should be clear from the above, the Simple Conditional Analysis is of no help. However, some recent work by Michael Smith (2003), Kadri Vihvelin (2004; 2013), and Michael Fara (2008) have attempted to fill this gap. What unites these theorists—whom Clarke (2009) has called the ‘new dispositionalists’—is their attempt to appeal to recent advances in the metaphysics of dispositions to arrive at a revised conditional analysis of the freedom to do otherwise. The most perspicuous of these accounts is offered by Vihvelin (2004), who argues that an agent’s having the ability to do otherwise is solely a function of the agent’s intrinsic properties. (It is important to note that Vihvelin [2013] has come to reject the view that free will consists exclusively in the kind of ability analyzed below.) Building on Lewis’s work on the metaphysics of dispositions, she arrives at the following analysis of ability:

Revised Conditional Analysis of Ability : \(S\) has the ability at time \(t\) to do \(X\) iff, for some intrinsic property or set of properties \(B\) that \(S\) has at \(t\), for some time \(t'\) after \(t\), if \(S\) chose (decided, intended, or tried) at \(t\) to do \(X\), and \(S\) were to retain \(B\) until \(t'\), \(S\)’s choosing (deciding, intending, or trying) to do \(X\) and \(S\)’s having \(B\) would jointly be an \(S\)-complete cause of \(S\)’s doing \(X\). (Vihvelin 2004, 438)

Lewis defines an ‘\(S\)-complete cause’ as “a cause complete insofar as havings of properties intrinsic to [\(S\)] are concerned, though perhaps omitting some events extrinsic to [\(S\)]” (cf. Lewis 1997, 156). In other words, an \(S\)-complete cause of \(S\)’s doing \(\phi\) requires that \(S\) possess all the intrinsic properties relevant to \(S\)’s causing \(S\)’s doing \(\phi\). This analysis appears to afford Vihvelin the basis for a principled difference between agoraphobics and merely determined agents. We must hold fixed an agent’s phobias since they are intrinsic properties of agents, but we need not hold fixed the laws of nature because these are not intrinsic properties of agents. (It should be noted that the assumption that intrinsic properties are wholly separable from the laws of nature is disputed by ‘dispositional essentialists.’ See the entry on metaphysics of causation .) Vihvelin’s analysis appears to be restrictive enough to exclude phobics from having the freedom to do otherwise, but permissive enough to allow that some agents in deterministic worlds have the freedom to do otherwise.

But appearances can be deceiving. The new dispositionalist claims have received some serious criticism, with the majority of the criticisms maintaining that these analyses are still too permissive (Clarke 2009; Whittle 2010; Franklin 2011b). For example, Randolph Clarke argues that Vihvelin’s analysis fails to overcome the original problem with the Simple Conditional Analysis . He writes, “A phobic agent might, on some occasion, be unable to choose to A and unable to A without so choosing, while retaining all that she would need to implement such a choice, should she make it. Despite lacking the ability to choose to A , the agent might have some set of intrinsic properties B such that, if she chose to A and retained B , then her choosing to A and her having B would jointly be an agent-complete cause of her A -ing” (Clarke 2009, p. 329).

The Categorical Analysis , and thus incompatibilism about free will and determinism, remains an attractive option for many philosophers precisely because it seems that compatibilists have yet to furnish an analysis of the freedom to do otherwise that implies that phobics clearly lack the ability to choose or do otherwise that is relevant to moral responsibility and yet some merely determined agents have this ability.

Some have tried to avoid these lingering problems for compatibilists by arguing that the freedom to do otherwise is not required for free will or moral responsibility. What matters for an agent’s freedom and responsibility, so it is argued, is the source of her action—how her action was brought about. The most prominent strategy for defending this move appeals to ‘Frankfurt-style cases’. In a ground-breaking article, Harry Frankfurt (1969) presented a series of thought experiments intended to show that it is possible that agents are morally responsible for their actions and yet they lack the ability to do otherwise. While Frankfurt (1971) took this to show that moral responsibility and free will come apart—free will requires the ability to do otherwise but moral responsibility does not—if we define ‘free will’ as ‘the strongest control condition required for moral responsibility’ (cf. Wolf 1990, 3–4; Fischer 1994, 3; Mele 2006, 17), then if Frankfurt-style cases show that moral responsibility does not require the ability to do otherwise, then they also show that free will does not require the ability to do otherwise. Let us consider this challenge in more detail.

Here is a representative Frankfurt-style case:

Imagine, if you will, that Black is a quite nifty (and even generally nice) neurosurgeon. But in performing an operation on Jones to remove a brain tumor, Black inserts a mechanism into Jones’s brain which enables Black to monitor and control Jones’s activities. Jones, meanwhile, knows nothing of this. Black exercises this control through a sophisticated computer which he has programmed so that, among other things, it monitors Jones’s voting behavior. If Jones were to show any inclination to vote for Bush, then the computer, through the mechanism in Jones’s brain, intervenes to ensure that he actually decides to vote for Clinton and does so vote. But if Jones decides on his own to vote for Clinton, the computer does nothing but continue to monitor—without affecting—the goings-on in Jones’s head. (Fischer 2006, 38)

Fischer goes on to suppose that Jones “decides to vote for Clinton on his own”, without any interference from Black, and maintains that in such a case Jones is morally responsible for his decision. Fischer draws two interrelated conclusions from this case. The first, negative conclusion, is that the ability to do otherwise is not necessary for moral responsibility. Jones is unable to refrain from deciding to vote for Clinton, and yet, so long as Jones decides to vote for Clinton on his own, his decision is free and one for which he is morally responsible. The second, positive conclusion, is that freedom and responsibility are functions of the actual sequence . What matters for an agent’s freedom and moral responsibility is not what might have happened, but how his action was actually brought about. What matters is not whether the agent had the ability to do otherwise, but whether he was the source of his actions.

The success of Frankfurt-style cases is hotly contested. An early and far-reaching criticism is due to David Widerker (1995), Carl Ginet (1996), and Robert Kane (1996, 142–43). According to this criticism, proponents of Frankfurt-style cases face a dilemma: either these cases assume that the connection between the indicator (in our case, the absence of Jones’s showing any inclination to decide to vote for Bush) and the agent’s decision (here, Jones’s deciding to vote for Clinton) is deterministic or not. If the connection is deterministic, then Frankfurt-style cases cannot be expected to convince incompatibilists that the ability to do otherwise is not necessary for moral responsibility and/or free will, since Jones’s action will be deterministically brought about by factors beyond his control, leading incompatibilists to conclude that Jones is not morally responsible for his decision. But if the connection is nondeterministic, then it is possible even in the absence of showing any inclination to decide to vote for Bush, that Jones decides to vote for Bush, and so he retains the ability to do otherwise. Either way Frankfurt-style cases fail to show that Jones is both morally responsible for his decision and yet is unable to do otherwise.

While some have argued that even Frankfurt-style cases that assume determinism are effective (see, e.g., Fischer 1999, 2010, 2013 and Haji and McKenna 2004 and for criticisms of this approach, see Goetz 2005, Palmer 2005, 2014, Widerker and Goetz 2013, and Cohen 2017), the majority of proponents of Frankfurt-style cases have attempted to revise these cases so that they are explicitly nondeterministic and yet still show that the agent was morally responsible even though he lacked the ability to do otherwise—or, at least that he lacked any ability to do otherwise that could be relevant to grounding the agent’s moral responsibility (see, e.g., Mele and Robb 1998, 2003, Pereboom 2001, 2014, McKenna 2003, Hunt 2005, and for criticisms of these cases see Ginet 2002, Timpe 2006, Widerker 2006, Franklin 2011c, Moya 2011, Palmer 2011, 2013, Robinson 2014, Capes 2016, Capes and Swenson 2017, and Elzein 2017).

Supposing that Frankfurt-style cases are successful, what exactly do they show? In our view, they show neither that free will and moral responsibility do not require an ability to do otherwise in any sense nor that compatibilism is true. Frankfurt-style cases are of clear help to the compatibilists’ position (though see Speak 2007 for a dissenting opinion). The Consequence Argument raises a powerful challenge to the cogency of compatibilism. But if Frankfurt-style cases are successful, agents can act freely in the sense relevant to moral responsibility while lacking the ability to do otherwise in the all-in sense. This allows compatibilists to concede that the all-in ability to do otherwise is incompatible with determinism, and yet insist that it is irrelevant to the question of the compatibility of determinism with moral responsibility (and perhaps even free will, depending on how we define this) (cf. Fischer 1987, 1994. For a challenge to the move from not strictly necessary to irrelevant, see O’Connor [2000, 20–22] and in reply, Fischer [2006, 152–56].). But, of course, showing that an argument for the falsity of compatibilism is irrelevant does not show that compatibilism is true. Indeed, many incompatibilists maintain that Frankfurt-style cases are successful and defend incompatibilism not via the Consequence Argument, but by way of arguments that attempt to show that agents in deterministic worlds cannot be the ‘source’ of their actions in the way that moral responsibility requires (Stump 1999; Zagzebski 2000; Pereboom 2001, 2014). Thus, if successful, Frankfurt-style cases would be at best the first step in defending compatibilism. The second step must offer an analysis of the kind of sourcehood constitutive of free will that entails that free will is compatible with determinism (cf. Fischer 1982).

Furthermore, while proponents of Frankfurt-style cases often maintain that these cases show that no ability to do otherwise is necessary for moral responsibility (“I have employed the Frankfurt-type example to argue that this sense of control [i.e. the one required for moral responsibility] need not involve any alternative possibilities” [Fischer 2006, p. 40; emphasis ours]), we believe that this conclusion overreaches. At best, Frankfurt-style cases show that the ability to do otherwise in the all-in sense —in the sense defined by the Categorical Analysis —is not necessary for free will or moral responsibility (cf. Franklin 2015). To appreciate this, let us assume that in the above Frankfurt-style case Jones lacks the ability to do otherwise in the all-in sense: there is no possible world in which we hold fixed the past and laws and yet Jones does otherwise, since all such worlds include Black and his preparations for preventing Jones from doing otherwise should Jones show any inclination. Even if this is all true, it should take only a little reflection to recognize that in this case Jones is able to do otherwise in certain weaker senses we might attach to that phrase, and compatibilists in fact still think that the ability to do otherwise in some such senses is necessary for free will and moral responsibility. Consequently, even though Frankfurt-style cases have, as a matter of fact, moved many compatibilists away from emphasizing ability to do otherwise to emphasizing sourcehood, we suggest that this move is best seen as a weakening of the ability-to-do-otherwise condition on moral responsibility (but see Cyr 2017 and Kittle 2019 for criticisms of this claim). (A potentially important exception to this claim is Sartorio [2016], who appealing to some controversial ideas in the metaphysics of causation appears to argue that no sense of the ability to do otherwise is necessary for control in the sense at stake for moral responsibility, but instead what matters is whether the agent is the cause of the action. We simply note that Sartorio’s account of causation is a modal one [see especially Sartorio (2016, 94–95, 132–37)] and thus it is far from clear that her account of freedom and responsibility is really an exception.)

In this section, we will assume that Frankfurt-style cases are successful in order to consider two prominent compatibilist attempts to construct analyses of the sourcehood condition (though see the entry on compatibilism for a more systematic survey of compatibilist theories of free will). The first, and perhaps most popular, compatibilist model is a reasons-responsiveness model. According to this model, an agent’s action \(\phi\) is free just in case the agent or manner in which the action is brought about is responsive to the reasons available to the agent at the time of action. While compatibilists develop this kind of account in different ways, the most detailed proposal is due to John Martin Fischer (1994, 2006, 2010, 2012; Fischer and Ravizza 1998. For similar compatibilist treatments of reasons-responsiveness, see Wolf 1990, Wallace 1994, Haji 1998, Nelkin 2011, McKenna 2013, Vargas 2013, Sartorio 2016). Fischer and Ravizza argue that an agent’s action is free and one for which he is morally responsible only if the mechanism that issued in the action is moderately reasons-responsive (Fischer and Ravizza 1998, ch. 3). By ‘mechanism’, Fischer and Ravizza simply mean “the way the action was brought about” (38). One mechanism they often discuss is practical deliberation. For example, in the case of Jones discussed above, his decision to vote for Clinton on his own was brought about by the process of practical deliberation. What must be true of this process, this mechanism, for it to be moderately reasons-responsive? Fischer and Ravizza maintain that moderate reasons-responsiveness consists in two conditions: reasons-receptivity and reasons-reactivity. A mechanism’s reasons-receptivity depends on the agent’s cognitive capacities, such as being capable of understanding moral reasons and the implications of their actions (69–73). The second condition is more important for us in the present context. A mechanism’s reasons-reactivity depends on how the mechanism would react given different reasons for action. Fischer and Ravizza argue that the kind of reasons-reactivity at stake is weak reasons-reactivity, where this merely requires that there is some possible world in which the laws of nature remain the same, the same mechanism operates, there is a sufficient reason to do otherwise, and the mechanism brings about the alternative action in response to this sufficient reason (73–76). On this analysis, while Jones, due to the activity of Black, lacks the ‘all-in’ sense of the ability to do otherwise, he is nevertheless morally responsible for deciding to vote for Clinton because his action finds its source in Jones’s practical deliberation that is moderately reasons-responsive.

Fischer and Ravizza’s theory of freedom and responsibility has shifted the focus of much recent debate to questions of sourcehood. Moreover, one might argue that this theory is a clear improvement over classical compatibilism with respect to handling cases of phobia. By focusing on mechanisms, Fischer and Ravizza can argue that our agoraphobic Luke is not morally responsible for deciding to refrain from going outside because the mechanism that issues in this action—namely his agoraphobia—is not moderately reasons-responsive. There is no world with the same laws of nature as our own, this mechanism operates, and yet it reacts to a sufficient reason to go outside. No matter what reasons there are for Luke to go outside, when acting on this mechanism, he will always refrain from going outside (cf. Fischer 1987, 74).

Before turning to our second compatibilist model, it is worth noting that it would be a mistake to think that Fischer and Ravizza’s account is a sourcehood account to the exclusion of the ability to do otherwise in any sense. As we have just seen, Fischer and Ravizza place clear modal requirements on mechanisms that issue in actions with respect to which agents are free and morally responsible. Indeed, this should be clear from the very idea of reasons-responsiveness. Whether one is responsive depends not merely on how one does respond, but also on how one would respond. Thus, any account that makes reasons-responsiveness an essential condition of free will is an account that makes the ability to do otherwise, in some sense, necessary for free will (Fischer [2018] concedes this point, though, as noted above, the reader should consider Sartorio [2016] as a potential counterexample to this claim).

The second main compatibilist model of sourcehood is an identification model. Accounts of sourcehood of this kind lay stress on self-determination or autonomy: to be the source of her action the agent must self-determine her action. Like the contemporary discussion of the ability to do otherwise, the contemporary discussion of the power of self-determination begins with the failure of classical compatibilism to produce an acceptable definition. According to classical compatibilists, self-determination simply consists in the agent’s action being determined by her strongest motive. On the assumption that some compulsive agents’ compulsions operate by generating irresistible desires to act in certain ways, the classical compatibilist analysis of self-determination implies that these compulsive actions are self-determined. While Hobbes seems willing to accept this implication (1656 [1999], 78), most contemporary compatibilists concede that this result is unacceptable.

Beginning with the work of Harry Frankfurt (1971) and Gary Watson (1975), many compatibilists have developed identification accounts of self-determination that attempt to draw a distinction between an agent’s desires or motives that are internal to the agent and those that are external. The idea is that while agents are not (or at least may not be) identical to any motivations (or bundle of motivations), they are identified with a subset of their motivations, rendering these motivations internal to the agent in such a way that any actions brought about by these motivations are self -determined. The identification relation is not an identity relation, but something weaker (cf. Bratman 2000, 39n12). What the precise nature of the identification relation is and to which attitudes an agent stands in this relation is hotly disputed. Lippert-Rasmussen (2003) helpfully divides identification accounts into two main types. The first are “authority” accounts, according to which agents are identified with attitudes that are authorized to speak for them (368). The second are authenticity accounts, according to which agents are identified with attitudes that reveal who they truly are (368). (But see Shoemaker 2015 for an ecumenical account of identification that blends these two accounts.) Proposed attitudes to which agents are said to stand in the identification relation include higher-order desires (Frankfurt 1971), cares or loves (Frankfurt 1993, 1994; Shoemaker 2003; Jaworska 2007; Sripada 2016), self-governing policies (Bratman 2000), the desire to make sense of oneself (Velleman 1992, 2009), and perceptions (or judgments) of the good (or best) (Watson 1975; Stump 1988; Ekstrom 1993; Mitchell-Yellin 2015).

The distinction between internal and external motivations allows identification theorists to enrich classical compatibilists’ understanding of constraint, while remaining compatibilists about free will and determinism. According to classical compatibilists, the only kind of constraint is external (e.g., broken cars and broken legs), but addictions and phobias seem just as threatening to free will. Identification theorists have the resources to concede that some constraints are internal. For example, they can argue that our agoraphobic Luke is not free in refraining from going outside even though this decision was caused by his strongest desires because he is not identified with his strongest desires. On compatibilist identification accounts, what matters for self-determination is not whether our actions are determined or undetermined, but whether they are brought about by motives with which the agent is identified: exercises of the power of self-determination consists in an agent’s actions being brought about, in part, by an agent’s motives with which she is identified. (It is important to note that while we have distinguished reasons-responsive accounts from identification accounts, there is nothing preventing one from combing both elements in a complete analysis of free will.)

Even if these reasons-responsive and identification compatibilist accounts of sourcehood might successfully side-step the Consequence Argument, they must come to grips with a second incompatibilist argument: the Manipulation Argument. The general problem raised by this line of argument is that whatever proposed compatibilist conditions for an agent \(S\)’s being free with respect to, and morally responsible for, some action \(\phi\), it will seem that agents can be manipulated into satisfying these conditions with respect to \(\phi\) and, yet, precisely because they are manipulated into satisfying these conditions, their freedom and responsibility seem undermined. The two most influential forms of the Manipulation Argument are Pereboom’s Four-case Argument (2001, ch. 4; 2014, ch. 4) and Mele’s Zygote Argument (2006, ch. 7. See Todd 2010, 2012 for developments of Mele’s argument). As the structure of Mele’s version is simpler, we will focus on it.

Imagine a goddess Diana who creates a zygote \(Z\) in Mary in some deterministic world. Suppose that Diana creates \(Z\) as she does because she wants Jones to be murdered thirty years later. From her knowledge of the laws of nature in her world and her knowledge of the state of the world just prior to her creating \(Z\), she knows that a zygote with precisely \(Z\)’s constitution located in Mary will develop into an agent Ernie who, thirty years later, will murder Jones as a result of his moderately reasons-responsive mechanism and on the basis of motivations with which he is identified (whatever those might be). Suppose Diana succeeds in her plan and Ernie murders Jones as a result of her manipulation.

Many judge that Ernie is not morally responsible for murdering Jones even though he satisfies both the reasons-responsive and identification criteria. There are two possible lines of reply open to compatibilists. On the soft-line reply, compatibilists attempt to show that there is a relevant difference between manipulated agents such as Ernie and agents who satisfy their account (McKenna 2008, 470). For example, Fischer and Ravizza propose a second condition on sourcehood: in addition to a mechanism’s being moderately reasons-responsive, an agent is morally responsible for the output of such a mechanism only if the agent has come to take responsibility for the mechanism, where an agent has taken responsibility for a mechanism \(M\) just in case (i) she believes that she is an agent when acting from \(M\), (ii) she believes that she is an apt target for blame and praise for acting from \(M\), and (iii) her beliefs specified in (i) and (ii) are “based, in an appropriate way, on [her] evidence” (Fischer and Ravizza 1998, 238). The problem with this reply is that we can easily imagine Diana creating Ernie so that his murdering Jones is a result not only of a moderately reasons-responsive mechanism, but also a mechanism for which he has taken responsibility. On the hard-line reply, compatibilists concede that, despite initial appearances, the manipulated agent is free and morally responsible and attempt to ameliorate the seeming counterintuitiveness of this concession (McKenna 2008, 470–71). Here compatibilists might point out that the idea of being manipulated is worrisome only so long as the manipulators are interfering with an agent’s development. But if the manipulators simply create a person, and then allow that person’s life to unfold without any further inference, the manipulators’ activity is no threat to freedom (McKenna 2008; Fischer 2011; Sartorio 2016, ch. 5). (For other responses to the Manipulation Argument, see Kearns 2012; Sripada 2012; McKenna 2014.)

Despite these compatibilist replies, to some the idea that the entirety of a free agent’s life can be determined, and in this way controlled, by another agent will seem incredible. Some take the lesson of the Manipulation Argument to be that no compatibilist account of sourcehood or self-determination is satisfactory. True sourcehood—the kind of sourcehood that can actually ground an agent’s freedom and responsibility—requires, so it is argued, that one’s action not be causally determined by factors beyond one’s control.

Libertarians, while united in endorsing this negative condition on sourcehood, are deeply divided concerning which further positive conditions may be required. It is important to note that while libertarians are united in insisting that compatibilist accounts of sourcehood are insufficient, they are not committed to thinking that the conditions of freedom spelled out in terms either of reasons-responsiveness or of identification are not necessary. For example, Stump (1988, 1996, 2010) builds a sophisticated libertarian model of free will out of resources originally developed within Frankfurt’s identification model (see also Ekstrom 1993, 2000; Franklin 2014) and nearly all libertarians agree that exercises of free will require agents to be reasons-responsive (e.g., Kane 1996; Clarke 2003, chs. 8–9; Franklin 2018, ch. 2). Moreover, while this section focuses on libertarian accounts of sourcehood, we remind readers that most (if not all) libertarians think that the freedom to do otherwise is also necessary for free will and moral responsibility.

There are three main libertarian options for understanding sourcehood or self-determination: non-causal libertarianism (Ginet 1990, 2008; McCann 1998; Lowe 2008; Goetz 2009; Pink 2017; Palmer 2021), event-causal libertarianism (Wiggins 1973; Kane 1996, 1999, 2011, 2016; Mele 1995, chs. 11–12; 2006, chs. 4–5; 2017; Ekstrom 2000, 2019; Clarke 2003, chs. 2–6; Franklin 2018), and agent-causal libertarianism (Reid 1788 [1969]; Chisholm 1966, 1976; Taylor 1966; O’Connor 2000; Clarke 1993; 1996; 2003, chs. 8–10; Griffith 2010; Steward 2012). Non-causal libertarians contend that exercises of the power of self-determination need not (or perhaps even cannot) be caused or causally structured. According to this view, we control our volition or choice simply in virtue of its being ours—its occurring in us. We do not exert a special kind of causality in bringing it about; instead, it is an intrinsically active event, intrinsically something we do . While there may be causal influences upon our choice, there need not be, and any such causal influence is wholly irrelevant to understanding why it occurs. Reasons provide an autonomous, non-causal form of explanation. Provided our choice is not wholly determined by prior factors, it is free and under our control simply in virtue of being ours. Non-causal views have failed to garner wide support among libertarians since, for many, self- determination seems to be an essentially causal notion (cf. Mele 2000 and Clarke 2003, ch. 2). This dispute hinges on the necessary conditions on the concept of causal power, and relatedly on whether power simpliciter admits causal and non-causal variants. For discussion, see O’Connor (2021).

Most libertarians endorse an event-causal or agent-causal account of sourcehood. Both these accounts maintain that exercises of the power of self-determination consist partly in the agent’s bringing about her choice or action, but they disagree on how to analyze an agent’s bringing about her choice . While event-causal libertarianism admits of different species, at the heart of this view is the idea that self-determining an action requires, at minimum, that the agent cause the action and that an agent’s causing his action is wholly reducible to mental states and other events involving the agent nondeviantly causing his action. Consider an agent’s raising his hand. According to the event-causal model at its most basic level, an agent’s raising his hand consists in the agent’s causing his hand to rise and his causing his hand to rise consists in apt mental states and events involving the agent—such as the agent’s desire to ask a question and his belief that he can ask a question by raising his hand— nondeviantly causing his hand to rise. (The nondeviance clause is required since it seems possible that an event be brought about by one’s desires and beliefs and yet not be self-determined, or even an action for that matter, due to the unusual causal path leading from the desires and beliefs to action. Imagine a would-be accomplice of an assassin believes that his dropping his cigarette is the signal for the assassin to shoot his intended victim and he desires to drop his cigarette and yet this belief and desire so unnerve him that he accidentally drops his cigarette. While the event of dropping the cigarette is caused by a relevant desire and belief it does not seem to be self-determined and perhaps is not even an action [cf. Davidson 1973].) To fully spell out this account, event-causal libertarians must specify which mental states and events are apt (cf. Brand 1979)—which mental states and events are the springs of self-determined actions—and what nondeviance consists in (cf. Bishop 1989). (We note that this has proven very difficult, enough so that some take the problem to spell doom for event-causal theories of action. Such philosophers [e.g., Taylor 1966 and Sehon 2005] take agential power to be conceptually and/or ontologically primitive and understand reasons explanations of action in irreducibly teleological terms. See Stout 2010 for a brisk survey of discussions of this topic.) For ease, in what follows we will assume that apt mental states are an agent’s reasons that favor the action.

Event-causal libertarians, of course, contend that self-determination requires more than nondeviant causation by agents’ reasons: for it is possible that agents’ actions in deterministic worlds are nondeviantly caused by apt mental states and events. Self-determination requires nondeterministic causation, in a nondeviant way, by an agent’s reasons. While historically many have thought that nondeterministic causation is impossible (Hobbes 1654 [1999], 1656 [1999]; Hume 1740 [1978], 1748 [1975]), with the advent of quantum physics and, from a very different direction, an influential essay by G.E.M. Anscombe (1971), it is now widely assumed that nondeterministic (or probabilistic) causation is possible. There are two importantly different ways to understand nondeterministic causation: as the causation of probability or as the probability of causation. Under the causation of probability model, a nondeterministic cause \(C\) causes (or causally contributes to) the objective probability of the outcome’s occurring rather than the outcome itself. On this account, \(S\)’s reasons do not cause his decision but there being a certain antecedent objective probability of its occurring, and the decision itself is uncaused. On the competing probability of causation model, a nondeterministic cause \(C\) causes the outcome of a nondeterministic process. Given that \(C\) is a nondeterministic cause of the outcome, it was possible given the exact same past and laws of nature that \(C\) not cause the outcome (perhaps because it was possible that some other event cause some other outcome)—the probability of this causal transaction’s occurring was less than \(1\). Given that event-causal libertarians maintain that self-determined actions, and thus free actions, must be caused, they are committed to the probability of causation model of nondeterministic causation (cf. Franklin 2018, 25–26). (We note that Balaguer [2010] is skeptical of the above distinction, and it is thus unclear whether he should best be classified as a non-causal or event-causal libertarian, though see Balaguer [2014] for evidence that it is best to treat him as a non-causalist.) Consequently, according to event-causal libertarians, when an agent \(S\) self-determines his choice \(\phi\), then \(S\)’s reasons \(r_1\) nondeterministically cause (in a nondeviant way) \(\phi\), and it was possible, given the past and laws, that \(r_1\) not have caused \(\phi\), but rather some of \(S\)’s other reasons \(r_2\) nondeterministically caused (in a nondeviant way) a different action \(\psi\).

Agent-causal libertarians contend that the event-causal picture fails to capture self-determination, for it fails to accord the agent with a power to settle what she does. Pereboom offers a forceful statement of this worry:

On an event-causal libertarian picture, the relevant causal conditions antecedent to the decision, i.e., the occurrence of certain agent-involving events, do not settle whether the decision will occur, but only render the occurrence of the decision about \(50\%\) probable. In fact, because no occurrence of antecedent events settles whether the decision will occur, and only antecedent events are causally relevant, nothing settles whether the decision will occur. (Pereboom 2014, 32; cf. Watson 1987, 1996; Clarke 2003 [ch. 8], 2011; Griffith 2010; Shabo 2011, 2013; Steward 2012 [ch. 3]; and Schlosser 2014); and for critical assessment, see Clarke 2019.

On the event-causal picture, the agent’s causal contribution to her actions is exhausted by the causal contribution of her reasons, and yet her reasons leave open which decisions she will make, and this seems insufficient for self-determination.

But what more must be added? Agent-causal libertarians maintain that self-determination requires that the agent herself play a causal role over and above the causal role played by her reasons. Some agent-causal libertarians deny that an agent’s reasons play any direct causal role in bringing about an agent’s self-determined actions (Chisholm 1966; O’Connor 2000, ch. 5), whereas others allow or even require that self-determined actions be caused in part by the agent’s reasons (Clarke 2003, ch. 9; Steward 2012, ch. 3). But all agent-causal libertarians insist that exercises of the power of self-determination do not reduce to nondeterministic causation by apt mental states: agent-causation does not reduce to event-causation.

Agent-causal libertarianism seems to capture an aspect of self-determination that neither the above compatibilists accounts nor event-causal libertarian accounts capture. (Some compatibilists even accept this and try to incorporate agent-causation into a compatibilist understanding of free will. See Markosian 1999, 2012; Nelkin 2011.) These accounts reduce the causal role of the self to states and events to which the agent is not identical (even if he is identified with them). But how can self -determination of my actions wholly reduce to determination of my actions by things other than the self? Richard Taylor nicely expresses this intuition: “If I believe that something not identical to myself was the cause of my behavior—some event wholly external to myself, for instance, or even one internal to myself, such as a nerve impulse, volition, or whatnot—then I cannot regard the behavior as being an act of mine, unless I further believed that I was the cause of that external or internal event” (1974, 55; cf. Franklin 2016).

Despite its powerful intuitive pull for some, many have argued that agent-causal libertarianism is obscure or even incoherent. The stock objection used to be that the very idea of agent-causation—causation by agents that is not reducible to causation by mental states and events involving the agent—is incoherent, but this objection has become less common due to pioneering work by Chisholm (1966, 1976), Taylor (1974), O’Connor (2000, 2011), Clarke (2003), and Steward 2012, ch. 8). More common objections now concern, first, how to understand the relationship between agent-causation and an agent’s reasons (or motivations in general), and, second, the empirical adequacy of agent-causal libertarianism. With respect to the first worry, it is widely assumed that the only (or at least best) way to understand reasons-explanation and motivational influence is within a causal account of reasons, where reasons cause our actions (Davidson 1963; Mele 1992). If agent-causal libertarians accept that self-determined actions, in addition to being agent-caused, must also be caused by agents’ reasons that favored those actions, then agent-causal libertarians need to explain how to integrate these causes (for a detailed attempt to do just this, see Clarke 2003, ch. 8). Given that these two causes seem distinct, is it not possible that the agent cause his decision to \(\phi\) and yet the agent’s reasons simultaneously cause an incompatible decision to \(\psi\)? If agent-causal libertarians side-step this difficult question by denying that reasons cause action, then they must explain how reasons can explain and motivate action without causing it; and this has turned out to be no easy task. (For more general attempts to understand reasons-explanation and motivation within a non-causal framework see Schueler 1995, 2003; Sehon 2005). For further discussion see the entry on incompatibilist (nondeterministic) theories of free will .

Finally, we note that some recent philosophers have questioned the presumed difference between event- and agent-causation by arguing that all causation is object or substance causation. They argue that the dominant tendency to understand ‘garden variety’ causal transactions in the world as relations between events is an unfortunate legacy of David Hume’s rejection of substance and causation as basic metaphysical categories. On the competing metaphysical picture of the world, the event or state of an object’s having some property such as mass is its having a causal power, which in suitable circumstances it exercises to bring about a characteristic effect. Applied to human agents in an account of free will, the account suggests a picture on which an agent’s having desires, beliefs, and intentions are rational powers to will particular courses of action, and where the agent’s willing is not determined in any one direction, she wills freely. An advantage for the agent-causalist who embraces this broader metaphysics is ‘ideological’ parsimony. For different developments and defenses of this approach, see Lowe (2008), Swinburne (2013), and O’Connor (2021); and for reason to doubt that a substance-causal metaphysics helps to allay skepticism concerning free will, see Clarke and Reed (2015).

3. Do We Have Free Will?

Most philosophers theorizing about free will take themselves to be attempting to analyze a near-universal power of mature human beings. But as we’ve noted above, there have been free will skeptics in both ancient and (especially) modern times. (Israel 2001 highlights a number of such skeptics in the early modern period.) In this section, we summarize the main lines of argument both for and against the reality of human freedom of will.

There are both a priori and empirical arguments against free will (See the entry on skepticism about moral responsibility ). Several of these start with an argument that free will is incompatible with causal determinism, which we will not rehearse here. Instead, we focus on arguments that human beings lack free will, against the background assumption that freedom and causal determinism are incompatible.

The most radical a priori argument is that free will is not merely contingently absent but is impossible. Nietzsche 1886 [1966] argues to this effect, and more recently it has been argued by Galen Strawson (1986, ch. 2; 1994, 2002). Strawson associates free will with being ‘ultimately morally responsible’ for one’s actions. He argues that, because how one acts is a result of, or explained by, “how one is, mentally speaking” (\(M\)), for one to be responsible for that choice one must be responsible for \(M\). To be responsible for \(M\), one must have chosen to be \(M\) itself—and that not blindly, but deliberately, in accordance with some reasons \(r_1\). But for that choice to be a responsible one, one must have chosen to be such as to be moved by \(r_1\), requiring some further reasons \(r_2\) for such a choice. And so on, ad infinitum . Free choice requires an impossible infinite regress of choices to be the way one is in making choices.

There have been numerous replies to Strawson’s argument. Mele (1995, 221ff.) argues that Strawson misconstrues the locus of freedom and responsibility. Freedom is principally a feature of our actions, and only derivatively of our characters from which such actions spring. The task of the theorist is to show how one is in rational, reflective control of the choices one makes, consistent with there being no freedom-negating conditions. While this seems right, when considering those theories that make one’s free control to reside directly in the causal efficacy of one’s reasons (such as compatibilist reasons-responsive accounts or event-causal libertarianism), it is not beside the point to reflect on how one came to be that way in the first place and to worry that such reflection should lead one to conclude that true responsibility (and hence freedom) is undermined, since a complete distal source of any action may be found external to the agent. Clarke (2003, 170–76) argues that an effective reply may be made by indeterminists, and, in particular, by nondeterministic agent-causal theorists. Such theorists contend that (i) aspects of ‘how one is, mentally speaking’, fully explain an agent’s choice without causally determining it and (ii) the agent himself causes the choice that is made (so that the agent’s antecedent state, while grounding an explanation of the action, is not the complete causal source of it). Since the agent’s exercise of this power is causally undetermined, it is not true that there is a sufficient ‘ultimate’ source of it external to the agent. Finally, Mele (2006, 129–34, and 2017, 212–16) and O’Connor (2009b) suggest that freedom and moral responsibility come in degrees and grow over time, reflecting the fact that ‘how one is, mentally speaking’ is increasingly shaped by one’s own past choices. Furthermore, some choices for a given individual may reflect more freedom and responsibility than others, which may be the kernel of truth behind Strawson’s sweeping argument. (For discussion of the ways that nature, nurture, and contingent circumstances shape our behavior and raise deep issues concerning the extent of our freedom and responsibility, see Levy 2011 and Russell 2017, chs. 10–12.)

A second family of arguments against free will contend that, in one way or another, nondeterministic theories of freedom entail either that agents lack control over their choices or that the choices cannot be adequately explained. These arguments are variously called the ‘Mind’, ‘Rollback’, or ‘Luck’ argument, with the latter admitting of several versions. (For statements of such arguments, see van Inwagen 1983, ch. 4; 2000; Haji 2001; Mele 2006; Shabo 2011, 2013, 2020; Coffman 2015). We note that some philosophers advance such arguments not as parts of a general case against free will, but merely as showing the inadequacy of specific accounts of free will [see, e.g., Griffith 2010].) They each describe imagined cases—individual cases, or comparison of intra- or inter-world duplicate antecedent conditions followed by diverging outcomes—designed to elicit the judgment that the occurrence of a choice that had remained unsettled given all prior causal factors can only be a ‘matter of chance’, ‘random’, or ‘a matter of luck’. Such terms have been imported from other contexts and have come to function as quasi-technical, unanalyzed concepts in these debates, and it is perhaps more helpful to avoid such proxies and to conduct the debates directly in terms of the metaphysical notion of control and epistemic notion of explanation. Where the arguments question whether an undetermined agent can exercise appropriate control over the choice he makes, proponents of nondeterministic theories often reply that control is not exercised prior to, but at the time of the choice—in the very act of bringing it about (see, e.g., Clarke 2005 and O’Connor 2007). Where the arguments question whether undetermined choices can be adequately explained, the reply often consists in identifying a form of explanation other than the form demanded by the critic—a ‘noncontrastive’ explanation, perhaps, rather than a ‘contrastive’ explanation, or a species of contrastive explanation consistent with indeterminism (see, e.g., Kane 1999; Clarke, 2003, ch. 8; and Franklin 2011a; 2018, ch. 5).

We now consider empirical arguments against human freedom. Some of these stem from the physical sciences (while making assumptions concerning the way physical phenomena fix psychological phenomena) and others from neuroscience and psychology.

It used to be common for philosophers to argue that there is empirical reason to believe that the world in general is causally determined, and since human beings are parts of the world, they are too. Many took this to be strongly confirmed by the spectacular success of Isaac Newton’s framework for understanding the universe as governed everywhere by fairly simple, exceptionless laws of motion. But the quantum revolution of the early twentieth century has made that ‘clockwork universe’ image at least doubtful at the level of basic physics. While quantum mechanics has proven spectacularly successful as a framework for making precise and accurate predictions of certain observable phenomena, its implications for the causal structure of reality is still not well understood, and there are competing indeterministic and deterministic interpretations. See the entry on quantum mechanics for detailed discussion.) It is possible that indeterminacy on the small-scale, supposing it to be genuine, ‘cancels out’ at the macroscopic scale of birds and buildings and people, so that behavior at this scale is virtually deterministic. But this idea, once common, is now being challenged empirically, even at the level of basic biology. Furthermore, the social, biological, and medical sciences, too, are rife with merely statistical generalizations. Plainly, the jury is out on all these inter-theoretic questions. But that is just a way to say that current science does not decisively support the idea that everything we do is pre-determined by the past, and ultimately by the distant past, wholly out of our control. For discussion, see Balaguer (2009), Koch (2009), Roskies (2014), Ellis (2016).

Maybe, then, we are subject to myriad causal influences, but the sum total of these influences doesn’t determine what we do, they only make it more or less likely that we’ll do this or that. Now some of the a priori no-free-will arguments above center on nondeterministic theories according to which there are objective antecedent probabilities associated with each possible choice outcome. Why objective probabilities of this kind might present special problems beyond those posed by the absence of determinism has been insufficiently explored to date. (For brief discussion, see Vicens 2016 and O’Connor 2016.) But one philosopher who argues that there is reason to hold that our actions, if undetermined, are governed by objective probabilities and that this fact calls into question whether we act freely is Derk Pereboom (2001, ch. 3; 2014, ch. 3). Pereboom notes that our best physical theories indicate that statistical laws govern isolated, small-scale physical events, and he infers from the thesis that human beings are wholly physically composed that such statistical laws will also govern all the physical components of human actions. Finally, Pereboom maintains that agent-causal libertarianism offers the correct analysis of free will. He then invites us to imagine that the antecedent probability of some physical component of an action occurring is \(0.32\). If the action is free while not violating the statistical law, then, in a scenario with a large enough number of instances, this action would have to be freely chosen close to \(32\) percent of the time. This leads to the problem of “wild coincidences”:

if the occurrence of these physical components were settled by the choices of agent-causes, then their actually being chosen close to 32 percent of the time would amount to a coincidence no less wild than the coincidence of possible actions whose physical components have an antecedent probability of about 0.99 being chosen, over large enough number of instances, close to 99 percent of the time. The proposal that agent-caused free choices do not diverge from what the statistical laws predict for the physical components of our actions would run so sharply counter to what we would expect as to make it incredible. (2014, 67)

Clarke (2010) questions the implicit assumption that free agent-causal choices should be expected not to conform to physical statistical laws, while O’Connor (2009a) challenges the more general assumption that freedom requires that agent-causal choices not be governed by statistical laws of any kind, as they plausibly would be if the relevant psychological states/powers are strongly emergent from physical states of the human brain. Finally, Runyan 2018 argues that Pereboom’s case rests on an implausible empirical assumption concerning the evolution of objective probabilities concerning types of behavior over time.

Pereboom’s empirical basis for free will skepticism is very general. Others see support for free will skepticism from specific findings and theories in the human sciences. They point to evidence that we can be unconsciously influenced in the choices we make by a range of factors, including ones that are not motivationally relevant; that we can come to believe that we chose to initiate a behavior that in fact was artificially induced; that people subject to certain neurological disorders will sometimes engage in purposive behavior while sincerely believing that they are not directing them. Finally, a great deal of attention has been given to the work of neuroscientist Benjamin Libet (2002). Libet conducted some simple experiments that seemed to reveal the existence of ‘preparatory’ brain activity (the ‘readiness potential’) shortly before a subject engages in an ostensibly spontaneous action. (Libet interpreted this activity as the brain’s ‘deciding’ what to do before we are consciously settled on a course of action.) Wegner (2002) surveys all of these findings (some of which are due to his own work as a social psychologist) and argues on their basis that the experience of conscious willing is ‘an illusion’. For criticism of such arguments, see Mele (2009); Nahmias (2014); Mudrik et al. (2022); and several contributions to Maoz and Sinnott-Armstrong (2022). Libet’s interpretation of the readiness potential has come in for severe criticism. After extensive subsequent study, neuroscientists are uncertain what it signifies. For thorough review of the evidence, see Schurger et al. (2021).

While Pereboom and others point to these empirical considerations in defense of free will skepticism, other philosophers see them as reasons to favor a more modest free will agnosticism (Kearns 2015) or to promote revisionism about the ‘folk idea of free will’ (Vargas 2013; Nichols 2015).

If one is a compatibilist, then a case for the reality of free will requires evidence for our being effective agents who for the most part are aware of what we do and why we are doing it. If one is an incompatibilist, then the case requires in addition evidence for causal indeterminism, occurring in the right locations in the process leading from deliberation to action. Many think that we already have third-personal ‘neutral’ scientific evidence for much of human behavior’s satisfying modest compatibilist requirements, such as Fischer and Ravizza’s reasons-responsiveness account. However, given the immaturity of social science and the controversy over whether psychological states ‘reduce’ in some sense to underlying physical states (and what this might entail for the reality of mental causation), this claim is doubtful. A more promising case for our satisfying (at least) compatibilist requirements on freedom is that effective agency is presupposed by all scientific inquiry and so cannot rationally be doubted (which fact is overlooked by some of the more extreme ‘willusionists’ such as Wegner).

However, effective intervention in the world (in scientific practice and elsewhere) does not (obviously) require that our behavior be causally undetermined, so the ‘freedom is rationally presupposed’ argument cannot be launched for such an understanding of freedom. Instead, incompatibilists usually give one of the following two bases for rational belief in freedom (both of which can be given by compatibilists, too).

First, philosophers have long claimed that we have introspective evidence of freedom in our experience of action, or perhaps of consciously attended or deliberated action. Augustine and Scotus, discussed earlier, are two examples among many. In recent years, philosophers have been more carefully scrutinizing the experience of agency and a debate has emerged concerning its contents, and in particular whether it supports an indeterministic theory of human free action. For discussion, see Deery et al. (2013), Guillon (2014), Horgan (2015), and Bayne (2017).

Second, philosophers (e.g., Reid 1788 [1969], Swinburne 2013) sometimes claim that our belief in the reality of free will is epistemically basic, or reasonable without requiring independent evidential support. Most philosophers hold that some beliefs have that status, on pain of our having no justified beliefs whatever. It is controversial, however, just which beliefs do because it is controversial which criteria a belief must satisfy to qualify for that privileged status. It is perhaps necessary that a basic belief be ‘instinctive’ (unreflectively held) for all or most human beings; that it be embedded in regular experience; and that it be central to our understanding of an important aspect of the world. Our belief in free will seems to meet these criteria, but whether they are sufficient is debated. (O’Connor 2019 proposes that free will belief is epistemically basic but defeasible.) Other philosophers defend a variation on this stance, maintaining instead that belief in the reality of moral responsibility is epistemically basic, and that since moral responsibility entails free will, or so it is claimed, we may infer the reality of free will (see, e.g., van Inwagen 1983, 206–13).

4. Theological Wrinkles

A large portion of Western philosophical work on free will has been written within an overarching theological framework, according to which God is the ultimate source, sustainer, and end of all else. Some of these thinkers draw the conclusion that God must be a sufficient, wholly determining cause for everything that happens; all of them suppose that every creaturely act necessarily depends on the explanatorily prior, cooperative activity of God. It is also commonly presumed by philosophical theists that human beings are free and responsible (on pain of attributing evil in the world to God alone, and so impugning His perfect goodness). Hence, those who believe that God is omni-determining typically are compatibilists with respect to freedom and (in this case) theological determinism. Edwards (1754 [1957]) is a good example. But those who suppose that God’s sustaining activity (and special activity of conferring grace) is only a necessary condition on the outcome of human free choices need to tell a more subtle story, on which omnipotent God’s cooperative activity can be (explanatorily) prior to a human choice and yet the outcome of that choice be settled only by the choice itself. For important medieval discussions—the apex of philosophical reflection on theological concerns—see the relevant portions of Al-Ghazali IP , Aquinas BW and Scotus QAM . Three positions (given in order of logical strength) on God’s activity vis-à-vis creaturely activity were variously defended by thinkers of this area: mere conservationism, concurrentism, and occasionalism. These positions turn on subtle distinctions, which have recently been explored by Freddoso (1988), Kvanvig and McCann (1991), Koons (2002), Grant (2016 and 2019), and Judisch (2016).

Many suppose that there is a challenge to human freedom stemming not only from God’s perfect power but also from his perfect knowledge. A standard argument for the incompatibility of free will and causal determinism has a close theological analogue. Recall van Inwagen’s influential formulation of the ‘Consequence Argument’:

If determinism is true, then our acts are the consequences of the laws of nature and events in the remote past. But it is not up to us what went on before we were born, and neither is it up to us what the laws of nature are. Therefore, the consequences of these things (including our present acts) are not up to us. (van Inwagen 1983, 16)

And now consider an argument that turns on God’s comprehensive and infallible knowledge of the future:

If infallible divine foreknowledge is true, then our acts are the (logical) consequences of God’s beliefs in the remote past. (Since God cannot get things wrong, his believing that something will be so entails that it will be so.) But it is not up to us what beliefs God had before we were born, and neither is it up to us that God’s beliefs are necessarily true. Therefore, the consequences of these things (including our present acts) are not up to us.

An excellent discussion of these arguments in tandem and attempts to point to relevant disanalogies between causal determinism and infallible foreknowledge may be found in the introduction to Fischer (1989). See also the entry on foreknowledge and free will.

Another issue concerns how knowledge of God, the ultimate Good, would impact human freedom. Many philosophical theologians, especially the medieval Aristotelians, were drawn to the idea that human beings cannot but will that which they take to be an unqualified good. (As noted above, Duns Scotus is an exception to this consensus, as were Ockham and Suarez subsequently, but their dissent is limited.) Hence, if there is an afterlife, in which humans ‘see God face to face,’ they will inevitably be drawn to Him. Following Pascal, Murray (1993, 2002) argues that a good God would choose to make His existence and character less than certain for human beings, for the sake of preserving their freedom. (He will do so, the argument goes, at least for a period of time in which human beings participate in their own character formation.) If it is a good for human beings that they freely choose to respond in love to God and to act in obedience to His will, then God must maintain an ‘epistemic distance’ from them lest they be overwhelmed by His goodness or power and respond out of necessity, rather than freedom. (See also the other essays in Howard-Snyder and Moser 2002.)

If it is true that God withholds our ability to be certain of his existence for the sake of our freedom, then it is natural to conclude that humans will lack freedom in heaven. And it is anyways common to traditional Jewish, Christian, and Muslim theologies to maintain that humans cannot sin in heaven. Even so, traditional Christian theology at least maintains that human persons in heaven are free. What sort of freedom is in view here, and how does it relate to mundane freedom? Two good recent discussions of these questions are Pawl and Timpe (2009) and Tamburro (2017).

Finally, there is the question of the freedom of God himself. Perfect goodness is an essential, not acquired, attribute of God. God cannot lie or be in any way immoral in His dealings with His creatures (appearances notwithstanding). Unless we take the minority position on which this is a trivial claim, since whatever God does definitionally counts as good, this appears to be a significant, inner constraint on God’s freedom. Did we not contemplate immediately above that human freedom would be curtailed by our having an unmistakable awareness of what is in fact the Good? And yet is it not passing strange to suppose that God should be less than perfectly free?

One suggested solution to this puzzle takes as its point of departure the distinction noted in section 2.3 between the ability to do otherwise and sourcehood, proposing that the core metaphysical feature of freedom is being the ultimate source, or originator, of one’s choices. For human beings or any created persons who owe their existence to factors outside themselves, the only way their acts of will could find their ultimate origin in themselves is for such acts not to be determined by their character and circumstances. For if all my willings were wholly determined, then if we were to trace my causal history back far enough, we would ultimately arrive at external factors that gave rise to me, with my particular genetic dispositions. My motives at the time would not be the ultimate source of my willings, only the most proximate ones. Only by there being less than deterministic connections between external influences and choices, then, is it be possible for me to be an ultimate source of my activity, concerning which I may truly say, “the buck stops here.”

As is generally the case, things are different on this point in the case of God. As Anselm observed, even if God’s character absolutely precludes His performing certain actions in certain contexts, this will not imply that some external factor is in any way a partial origin of His willings and refrainings from willing. Indeed, this would not be so even if he were determined by character to will everything which He wills. God’s nature owes its existence to nothing. Thus, God would be the sole and ultimate source of His will even if He couldn’t will otherwise.

Well, then, might God have willed otherwise in any respect? The majority view in the history of philosophical theology is that He indeed could have. He might have chosen not to create anything at all. And given that He did create, He might have created any number of alternatives to what we observe. But there have been noteworthy thinkers who argued the contrary position, along with others who clearly felt the pull of the contrary position even while resisting it. The most famous such thinker is Leibniz (1710 [1985]), who argued that God, being both perfectly good and perfectly powerful, cannot fail to will the best possible world. Leibniz insisted that this is consistent with saying that God is able to will otherwise, although his defense of this last claim is notoriously difficult to make out satisfactorily. Many read Leibniz, malgré lui , as one whose basic commitments imply that God could not have willed other than He does in any respect.

One might challenge Leibniz’s reasoning on this point by questioning the assumption that there is a uniquely best possible Creation (an option noted by Adams 1987, though he challenges instead Leibniz’s conclusion based on it). One way this could be is if there is no well-ordering of worlds: some pairs of worlds are sufficiently different in kind that they are incommensurate with each other (neither is better than the other, nor are they equal) and no world is better than either of them. Another way this could be is if there is no upper limit on goodness of worlds: for every possible world God might have created, there are others (infinitely many, in fact) which are better. If such is the case, one might argue, it is reasonable for God to arbitrarily choose which world to create from among those worlds exceeding some threshold value of overall goodness.

However, William Rowe (2004) has countered that the thesis that there is no upper limit on goodness of worlds has a very different consequence: it shows that there could not be a morally perfect Creator! For suppose our world has an on-balance moral value of \(n\) and that God chose to create it despite being aware of possibilities having values higher than \(n\) that He was able to create. It seems we can now imagine a morally better Creator: one having the same options who chooses to create a better world. For critical replies to Rowe, see Almeida (2008, ch. 1), Kray (2010), and Zimmerman (2018).

Finally, Norman Kretzmann (1997, 220–25) has argued in the context of Aquinas’s theological system that there is strong pressure to say that God must have created something or other, though it may well have been open to Him to create any of a number of contingent orders. The reason is that there is no plausible account of how an absolutely perfect God might have a resistible motivation—one consideration among other, competing considerations—for creating something rather than nothing. (It obviously cannot have to do with any sort of utility, for example.) The best general understanding of God’s being motivated to create at all—one which in places Aquinas himself comes very close to endorsing—is to see it as reflecting the fact that God’s very being, which is goodness, necessarily diffuses itself. Perfect goodness will naturally communicate itself outwardly; God who is perfect goodness will naturally create, generating a dependent reality that imperfectly reflects that goodness. Wainwright (1996) discusses a somewhat similar line of thought in the Puritan thinker Jonathan Edwards. Alexander Pruss (2016), however, raises substantial grounds for doubt concerning this line of thought; O’Connor (2022) offers a rejoinder.

  • Adams, Robert, 1987. “Must God Create the Best?” in his The Virtue of Faith and Other Essays in Philosophical Theology , New York: Oxford University Press, 51–64.
  • Al-Ghazali, 2000. [ IP ] The Incoherence of the Philosophers , Michael E. Marmura (ed.), Provo: Brigham Young University Press.
  • Almeida, Michael, 2008. The Metaphysics of Perfect Beings , New York: Routledge.
  • Anscombe, G. E. M., 1971. Causality and Determination , New York: Cambridge University Press.
  • Aquinas, Thomas, 1945. [ BW ] Basic Writings of Saint Thomas Aquinas , 2 volumes, New York: Random House.
  • –––, 1993. [ SPW ] Selected Philosophical Writings , T. McDermott (ed.), Oxford: Oxford University Press.
  • –––, 2003. [ DM ] On Evil , R. Regan (trans.), B. Davies (ed.), New York: Oxford University Press.
  • Aristotle, 1985. [ NE ] Nicomachean Ethics , tr. Terence Irwin, Indianapolis: Hackett Publishing.
  • Augustine, 1993. [ FCW ] On the Free Choice of the Will , tr. Thomas Williams, Indianapolis: Hackett Publishing.
  • Austin, J. L., 1961. “Ifs and Cans,” in Philosophical Papers , J. O. Urmson and G. Warnock (eds.), Oxford: Clarendon Press, 153–80.
  • Ayer, A. J., 1954. “Freedom and Necessity,” in his Philosophical Essays , New York: St. Martin’s Press, 3–20.
  • Baker, Lynne Rudder, 2003. “Why Christians Should Not Be Libertarians: An Augustinian Challenge,” Faith and Philosophy , 20 (4): 460–478.
  • Balaguer, Mark, 2009. “Why There Are No Good Arguments for any Interesting Version of Determinism,” Synthese , 168: 1–21.
  • –––, 2010. Free Will as an Open Scientific Problem , Cambridge, MA: MIT Press.
  • –––, 2014. “Replies to McKenna, Pereboom, and Kane,” Philosophical Studies , 169: 71–92
  • Bayne, Tim, 2017. “Free Will and the Phenomenology of Agency,” in The Routledge Companion to Free Will , Kevin Timpe, Meghan Griffith, and Neil Levy, (eds.), Abingdon: Routledge, 633–44.
  • Bishop, John, 1989. Natural Agency: An Essay on the Causal Theory of Action , New York: Cambridge University Press.
  • Bobzien, Suzanne, 1998. Determinism and Freedom in Stoic Philosophy , New York: Oxford University Press.
  • –––, 2000. “Did Epicurus Discover the Free-Will Problem?” Oxford Studies in Ancient Philosophy , 19: 287–337.
  • Bramhall, John, [1655] 1999. “Bramhall’s Discourse of Liberty and Necessity,” in Hobbes and Bramhall on Liberty and Necessity , Vere Chappell (ed.), Cambridge: Cambridge University Press, 1–14.
  • Brand, Myles, 1979. “The Fundamental Question in Action Theory,” Noûs , 13: 131–51.
  • Bratman, Michael, 2000. “Reflection, Planning, and Temporally Extended Agency,” Philosophical Review , 109: 35–61.
  • –––, 2005. “Planning Agency, Autonomous Agency,” in Personal Autonomy , James Stacey Taylor (ed.), New York: Cambridge University Press.
  • Broad, C. D., 1952. “Determinism, Indeterminism, and Libertarianism,” in his Ethics and the History of Philosophy: Selected Essays , London: Routledge & Kegan Paul LTD.
  • Campbell, C. A., 1951. “Is ‘Freewill’ a Pseudo-Problem?” Mind , 60: 441–65.
  • Capes, Justin A., 2016. “Blameworthiness and Buffered Alternatives,” American Philosophical Quarterly , 53: 269–80.
  • Capes, Justin A. and Philip Swenson, 2017. “Frankfurt Cases: The Fine-grained Response Revisited,” Philosophical Studies , 174: 967–81.
  • Caruso, Gregg, 2012. Free Will and Consciousness: A Determinist Account of the Illusion of Free Will , Plymouth: Lexington Books.
  • Chakrabarti, Arindam, 2017. “Free Will and Freedom in Indian Philosophies,” in The Routledge Companion to Free Will , Kevin Timpe, Meghan Griffith, and Neil Levy (eds.), New York: Routledge, 389–404.
  • Chisholm, Roderick, 1966. “Freedom and Action,” in Freedom and Determinism , Keith Lehrer (ed.), New York: Random House, 11–40.
  • –––, 1976. Person and Object: A Metaphysical Study , La Salle, IL: Open Court,
  • Clarke, Randolph, 1993. “Toward a Credible Agent-Causal Account of Free Will,” Noûs , 27: 191–203.
  • –––, 1995. “Indeterminism and Control,” American Philosophical Quarterly , 32: 125–38.
  • –––, 1996. “Agent Causation and Event Causation in the Production of Free Action,” Philosophical Topics , 24: 19–48.
  • –––, 2003. Libertarian Accounts of Free Will , Oxford: Oxford University Press.
  • –––, 2005. “Agent Causation and the Problem of Luck,” Pacific Philosophical Quarterly , 86: 408–21.
  • –––, 2009. “Dispositions, Abilities to Act, and Free Will: The New Dispositionalism,” Mind , 118: 323–51.
  • –––, 2010. “Are We Free to Obey the Laws?” American Philosophical Quarterly , 47: 389–401.
  • –––, 2011. “Alternatives for Libertarians,” in The Oxford Handbook of Free Will , Robert Kane (ed.), 2nd edition, New York: Oxford University Press, 329–48.
  • –––, 2019. “Free Will, Agent Causation, and ‘Disappearing Agents’,” Noûs , 53 (1): 76–96.
  • Clarke, Randolph and Thomas Reed, 2015. “Free Will and Agential Powers,” Oxford Studies in Agency and Moral Responsibility , 3: 6–33.
  • Coffman, E. J., 2015. Luck: Its Nature and Significance for Human Knowledge and Agency , Basingstoke, UK: Palgrave Macmillan.
  • Cohen, Yishai, 2017. “Fischer’s Deterministic Frankfurt-style Argument,” Erkenntnis , 82: 121–40.
  • Couenhoven, Jesse, 2007. “Augustine’s Rejection of the Free-will Defence: An Overview of the Late Augustine’s Theodicy,” Religious Studies , 43 (3): 279–298.
  • Cyr, Taylor W., 2017.“Semicompatibilism: No Ability to Do Otherwise Required,” Philosophical Explorations , 20 (3): 308–21.
  • Cyr, Taylor W. and Philip Swenson, 2019. “Moral Responsibility without General Ability,” Philosophical Quarterly , 69 (274): 22–40.
  • Davidson, Donald, 1963. “Actions, Reasons, and Causes,” Journal of Philosophy , 60: 685–700.
  • –––, 1973. “Freedom to Act,” in Essays on Freedom of Action , Ted Honderich (ed.), New York: Routledge and Kagan Press, 63–81.
  • Deery, Oisin, Matthew Bedke, and Shaun Nichols, 2013. “Phenomenal Abilities: Incompatibilism and the Experience of Agency,” Oxford Studies in Agency and Responsibility , 1: 126–50.
  • Dennett, Daniel, 1984. Elbow Room: The Varieties of Free Will Worth Wanting , Cambridge, MA: MIT Press.
  • Descartes, René, 1641 [1988]. Meditations on First Philosophy , in Descartes: Selected Philosophical Writings , John Cottingham, Robert Stoothoff, and Dugald Murdoch (eds. and trans.), Cambridge: Cambridge University Press, 73–121.
  • –––, 1644 [1988]. Principles of Philosophy , in Descartes: Selected Philosophical Writings , John Cottingham, Robert Stoothoff, and Dugald Murdoch (eds. and trans.), Cambridge: Cambridge University Press, 160–212.
  • Donagan, Alan, 1985. Human Ends and Human Actions: An Exploration in St. Thomas’s Treatment , Milwaukee: Marquette University Press.
  • Dihle, Albrecht, 1982. The Theory of Will in Classical Antiquity , Berkeley: University of California Press.
  • Dilman, Ilham, 1999. Free Will: An Historical and Philosophical Introduction , London: Routledge.
  • Double, Richard, 1991. The Non-Reality of Free Will , New York: Oxford University Press.
  • Duns Scotus, John, 1986. “Questions on Aristotle’s Metaphysics IX, Q.15” [QAM] in Duns Scotus on the Will and Morality , Allan B. Wolter, O.F.M. (ed. and trans.), Washington: Catholic University of America Press.
  • –––, [1297–99] [1994]. Contingency and Freedom: Lectura I 39 , tr. Vos Jaczn et al ., Dordrecht: Kluwer Academic Publishers.
  • Edwards, Jonathan, 1754 [1957]. Freedom of Will , Paul Ramsey (ed.), New Haven: Yale University Press.
  • Ekstrom, Laura Waddell, 1993. “A Coherence Theory of Autonomy,” Philosophy and Phenomenological Research , 53: 599–616.
  • –––, 2000. Free Will: A Philosophical Study , Boulder, CO: Westview Press.
  • –––, 2019. “Toward a Plausible Event-Causal Indeterminist Account of Free Will,” Synthese , 196 (1): 127–144.
  • Ellis, George, 2016. How Can Physics Underlie the Mind? , Berlin: Springer-Verlag.
  • Elzein, Nadine, 2017. “Frankfurt-Style Counterexamples and the Importance of Alternative Possibilities,” Acta Analytica , 32: 169–91.
  • Fara, Michael, 2008. “Masked Abilities and Compatibilism,” Mind , 117: 844–65.
  • Fischer, John Martin, 1982. “Responsibility and Control,” Journal of Philosophy , 79: 24–40.
  • –––, 1987. “Responsiveness and Moral Responsibility,” in Responsibility, Character, and the Emotions: New Essays on Moral Psychology , Ferdinand Schoeman (ed.), New York: Cambridge University Press, 81–106; reprinted in Fischer 2006 63–83. (Citations refer to reprinted edition.)
  • –––, 1989. God, Freedom, and Foreknowledge , Stanford, CA: Stanford University Press.
  • –––, 1994. The Metaphysics of Free Will: An Essay on Control , Oxford: Blackwell Publishers.
  • –––, 1999. “Recent Work on Moral Responsibility,” Ethics , 110: 93–139.
  • –––, 2006. My Way: Essays on Moral Responsibility , New York: Oxford University Press.
  • –––, 2010. “The Frankfurt Cases: The Moral of the Stories,” Philosophical Review , 119: 315–36.
  • –––, 2011. “The Zygote Argument Remixed,” Analysis , 71: 267–72.
  • –––, 2012. Deep Control: Essays on Free Will and Value , New York: Oxford University Press.
  • –––, 2013. “The Deterministic Horn of the Dilemma Defense: A Reply to Widerker and Goetz,” Analysis , 73: 489–96.
  • –––, 2018. “The Freedom Required for Moral Responsibility,” in Virtue, Happiness, and Knowledge: Themes from the Work of Gail Fine and Terence Irwin , David Brink, Susan Meyer, and Christopher Shields (eds.), Oxford: Oxford University Press, 216–233.
  • Fischer, John Martin and Mark Ravizza, 1998. Responsibility and Control: A Theory of Moral Responsibility , Cambridge: Cambridge University Press.
  • Fischer, John Martin and Neal Tognazzini, 2011. “The Physiognomy of Responsibility,” Philosophy and Phenomenological Research , 82: 381–417.
  • Frankfurt, Harry, 1969. “Alternate Possibilities and Moral Responsibility,” Journal of Philosophy , 66: 829–39.
  • –––, 1971. “Freedom of the Will and the Concept of a Person,” Journal of Philosophy , 68: 5–20.
  • –––, 1993. “On the Necessity of Ideals,” in The Moral Self , G.C. Noam and T. Wren (eds.), Cambridge, MA: MIT Press, 16–27.
  • –––, 1994. “Autonomy, Necessity, and Love,” in Vernunftbegriffe in der Moderne: Stuttgarter Hegel-Kongress 1993 , H.F. Fulda and R.P. Horstmann (eds.), Stuttgart: Klett-Cotta, 433–47.
  • Franklin, Christopher Evan, 2011a. “Farewell to the Luck (and Mind ) Argument,” Philosophical Studies , 156: 199–230.
  • –––, 2011b. “Masks, Abilities, and Opportunities: Why the New Dispositionalism Cannot Succeed,” The Modern Schoolman , 88: 89–103.
  • –––, 2011c. “Neo-Frankfurtians and Buffer Cases: The New Challenge to the Principle of Alternative Possibilities,” Philosophical Studies , 152: 189–207.
  • –––, 2014. “Event-Causal Libertarianism, Functional Reduction, and the Disappearing Agent Argument,” Philosophical Studies , 170: 413–32.
  • –––, 2015. “Everyone Thinks that an Ability to Do Otherwise Is Necessary for Free Will and Moral Responsibility,” Philosophical Studies , 172: 2091–107.
  • –––, 2016. “If Anyone Should Be an Agent-Causalist, then Everyone Should Be an Agent-Causalist,” Mind , 125: 1101–31.
  • –––, 2018. A Minimal Libertarianism: Free Will and the Promise of Reduction , New York: Oxford University Press.
  • Freddoso, Alfred, 1988. “Medieval Aristotelianism and the Case against Secondary Causation in Nature,” in Divine and Human Action: Essays in the Metaphysics of Theism , Thomas V. Morris (ed.), Ithaca, NY: Cornell University Press, 74–118.
  • Frede, Michael, 2011. A Free Will: Origins of the Notion in Ancient Thought , Berkeley: University of California Press.
  • Ginet, Carl, 1966. “Might We Have No Choice?” in Freedom and Determinism , Keith Lehrer (ed.), New York: Random House, 87–104.
  • –––, 1990. On Action , Cambridge: Cambridge University Press.
  • –––, 1996. “In Defense of the Principle of Alternative Possibilities: Why I Don’t Find Frankfurt’s Argument Convincing,” Philosophical Perspectives , 10: 403–17.
  • –––, 2002. “Review of Living without Free Will ,” Journal of Ethics , 6: 305–309.
  • –––, 2008. “In Defense of a Non-Causal Account of Reasons Explanations,” The Journal of Ethics , 12: 229–37.
  • Goetz, Stewart C., 2005. “Frankfurt-Style Counterexamples and Begging the Question,” Midwest Studies in Philosophy , 29: 83–105.
  • –––, 2009. Freedom, Teleology, and Evil , London: T&T Clark.
  • Grant, W. Matthews, 2016. “Divine Universal Causality and Libertarian Freedom,” in Free Will and Theism: Connections, Contingencies, and Concerns , Kevin Timpe and Daniel Speak (eds.), New York: Oxford University Press, 214–33.
  • –––, 2019. Free Will and God’s Universal Causality: The Dual Sources Account , London: Bloomsbury Academic.
  • Griffith, Meghan, 2010. “Why Agent-Caused Actions Are Not Lucky,” American Philosophical Quarterly , 47: 43–56.
  • Guillon, Jean-Baptiste, 2014. “Van Inwagen on Introspected Freedom,” Philosophical Studies , 168: 645–63.
  • Haji, Ishtiyaque, 1998. Moral Appraisability: Puzzles, Proposals, and Perplexities , New York: Oxford University Press.
  • –––, 2001. “Control Conundrums: Modest Libertarianism, Responsibility, and Explanation,” Pacific Philosophical Quarterly , 82: 178–200.
  • Haji, Ishtiyaque and Michael McKenna, 2004. “Dialectical Delicacies in the Debate about Freedom and Alternative Possibilities,” Journal of Philosophy , 101: 299–314.
  • Hecht, Jonathan, 2014. “Freedom of the Will in Plato and Augustine,” British Journal for the History of Philosophy , 22: 196–216.
  • Hobart, R. E., 1934. “Free Will as Involving Determination and Inconceivable Without It,” Mind , 43: 1–27.
  • Hobbes, Thomas, 1654 [1999]. Of Liberty and Necessity , in Hobbes and Bramhall on Liberty and Necessity , Vere Chappell (ed.), Cambridge: Cambridge University Press, 15–42.
  • –––, 1656 [1999]. The Questions concerning Liberty, Necessity, and Chance , in Hobbes and Bramhall on Liberty and Necessity , Vere Chappell (ed.), Cambridge: Cambridge University Press, 69–90.
  • Hoffman, Tobias and Cyrille Michon, 2017. “Aquinas on Free Will and Intellectual Determinism,” Philosopher’s Imprint , 17 (10), 1–36.
  • Hofmann, Frank, 2022. “Explaining Free Will by Rational Abilities,” Ethical Theory and Moral Practice , 25 (2): 283–93.
  • Holton, Richard, 2009. Willing, Wanting, Waiting , New York: Oxford University Press.
  • Horgan, Terrence, 1979. “‘Could’, Possible Worlds, and Moral Responsibility,” Southern Journal of Philosophy , 17: 345–58.
  • –––, 2015. “Injecting the Phenomenology of Agency into the Free Will Debate,” Oxford Studies in Agency and Responsibility , 3: 34–61.
  • Howard-Snyder, Daniel and Paul Moser (eds.), 2002. Divine Hiddenness: New Essays , Cambridge: Cambridge University Press.
  • Hume, David, 1740 [1978]. A Treatise of Human Nature , L.A. Selby-Bigge and P.H. Nidditch (eds.), 2 nd edition, Oxford: Oxford University Press.
  • –––, 1748 [1975]. Enquiries concerning Human Understanding and concerning the Principles of Morals , P.H. Nidditch (ed.), third edition, Oxford: Oxford University Press.
  • Hunt, David P., 2005. “Moral Responsibility and Buffered Alternatives,” Midwest Studies in Philosophy , 29: 126–45.
  • Irwin, Terence, 1992. “Who Discovered the Will?” Philosophical Perspectives , 6: 453–73.
  • Israel, Jonathan, 2001. Radical Enlightenment: Philosophy and the Making of Modernity 1650–1750 , Oxford: Oxford University Press.
  • Jacobs, Jonathan D. and Timothy O’Connor, 2013. “Agent Causation in a Neo-Aristotelian Metaphysics,” in Mental Causation and Ontology , Sophie C. Gibb, E.J. Lowe, and Rögnvaldur Ingthorsson (eds.), Oxford: Oxford University Press, 173–92.
  • Jaworska, Agnieszka, 2007. “Caring and Internality,” Philosophy and Phenomenological Research , 74: 529–68.
  • Judisch, Neal, 2016. “Divine conservation and creaturely freedom,” in Free Will and Theism: Connections, Contingencies, and Concerns , Kevin Timpe and Daniel Speak (eds.), New York: Oxford University Press, 234–58.
  • Kane, Robert, 1996. The Significance of Free Will , New York: Oxford University Press.
  • –––, 1999. “Responsibility, Luck, and Chance,” Journal of Philosophy , 96: 217–40.
  • –––, 2011. “Rethinking Free Will: New Perspectives on an Ancient Problem,” in The Oxford Handbook of Free Will , Robert Kane (ed.), 2 nd edition, New York: Oxford University Press, 381–404.
  • –––, 2016. “On the Role of Indeterminism in Libertarian Free Will,” Philosophical Explorations , 19: 2–16.
  • Kant, Immanuel, 1781 [1999]. Critique of Pure Reason , trs. Paul Guyer and Allen W. Wood. Cambridge: Cambridge University Press.
  • –––, 1785 [1998]. Groundwork of the Metaphysics of Morals , tr. Mary Gregor. Cambridge: Cambridge University Press.
  • –––, 1788 [2015]. Critique of Practical Reason , tr. Mary Gregor. Cambridge: Cambridge University Press.
  • Kearns, Stephen, 2012. “Aborting the Zygote Argument,” Philosophical Studies , 160: 379–89.
  • –––, 2015. “Free Will Agnosticism,” Noûs , 47: 235–52.
  • Kittle, Simon, 2019. “Does Everyone Think the Ability to do Otherwise is Necessary for Free Will and Moral Responsibility?” Philosophia , 47 (4): 1177–83.
  • Koch, Christof, 2009. “Free Will, Physics, Biology, and the Brain,” in Downward Causation and the Neurobiology of Free Will , George F.R. Ellis, Nancey Murphy, and Timothy O’Connor (eds.), New York: Springer, 31–52.
  • Koons, Robert, 2002. “Dual Agency: A Thomistic Account of Providence and Human Freedom,” Philosophia Christi , 4: 397–410.
  • Kraay, Klaas J., 2010. “The Problem of No Best World,” in A Companion to Philosophy of Religion , Charles Taliaferro and Paul Draper (eds.), 2 nd edition, Oxford: Blackwell, 491–99.
  • Kretzmann, Norman, 1997. The Metaphysics of Theism: Aquinas’s Natural Theology in Summa Contra Gentiles I. Oxford: Clarendon Press.
  • Kvanvig, Jon and Hugh McCann, 1991. “The Occasionalist Proselytizer: A Modified Catechism,” Philosophical Perspectives , 5: 587–615.
  • Lehrer, Keith, 1976. “Can in Theory and Practice: A Possible Worlds Analysis,” in Action Theory , Myles Brand and Douglas Walton (eds.), Dordrecht: D. Reidel Publishing Company, 242–71.
  • –––, 1980. “Preferences, Conditionals, and Freedom,” in Time and Cause: Essays Presented to Richard Taylor , Peter van Inwagen (ed.), Dordrecht: D. Reidel Publishing Company, 187–200.
  • –––, 1986. “Cans without Ifs,” Analysis , 29: 29–32.
  • Leibniz, G.W., 1686 [1991]. Discourse on Metaphysics and Other Essays , trs. Daniel Garber and Roger Ariew, 9 th edition. Indianapolis: Hackett Publishing.
  • –––, 1710 [1985]. Theodicy , LaSalle, IL: Open Court.
  • Levy, Neil, 2011. Hard Luck: How Luck Undermines Free Will and Moral Responsibility , New York: Oxford University Press.
  • Lewis, David, 1976. “The Paradoxes of Time Travel,” American Philosophical Quarterly , 13: 145–52.
  • –––, 1979. “Counterfactual Dependence and Time’s Arrow,” Noûs , 13: 455–76.
  • –––, 1981. “Are We Free to Break the Laws?” Theoria , 47: 113–21.
  • –––, 1997. “Finkish Dispositions,” Philosophical Quarterly , 47: 143–58.
  • Libet, Benjamin, 2002. “Do We Have Free Will?” in Oxford Handbook of Free Will , Robert Kane (ed.), 1st edition, New York: Oxford University Press, 551–64.
  • Lippert-Rasmussen, Kasper, 2003. “Identification and Responsibility,” Ethical Theory and Moral Practice , 6: 349–76.
  • Locke, Don, 1973. “Natural Powers and Human Abilities,” Proceedings of the Aristotelian Society , 74: 171–87.
  • Locke, John, 1690 [1975]. An Essay Concerning the Human Understanding , Peter H. Nidditch (ed.), Oxford: Oxford University Press.
  • Lowe, E. J., 2008. Personal Agency: The Metaphysics of Mind and Action , Oxford: Oxford University Press.
  • MacDonald, Scott, 1998. “Aquinas’s Libertarian Account of Free Will,” Revue Internationale de Philosophie , 2: 309–28.
  • –––, 1999. “Primal Sin,” in Gareth B. Matthews (ed.), The Augustinian Tradition , Berkeley, CA: University of California Press, 110–139.
  • Malebranche, Nicolas, 1684 [1993]. Treatise on Ethics , C. Walton (trans.), Dordrecht: Kluwer.
  • Maoz, Uri and Walter Sinnott-Armstrong (eds.), 2022. Free Will: Philosophers and Neuroscientists in Conversation , New York: Oxford University Press.
  • Marchal, Kai and Christian Helmut Wenzel, 2017. “Chinese Perspectives on Free Will,” in The Routledge Companion to Free Will , Kevin Timpe, Meghan Griffith, and Neil Levy (eds.), New York: Routledge, 374–88.
  • Markosian, Ned, 1999. “A Compatibilist Version of the Theory of Agent Causation,” Pacific Philosophical Quarterly , 80: 257–77.
  • –––, 2012. “Agent Causation as the Solution to all the Compatibilist’s Problems,” Philosophical Studies , 157: 383–98.
  • McCann, Hugh, 1998. The Works of Agency: On Human Action, Will, and Freedom , Ithaca: Cornell University Press.
  • McGeer, Victoria, 2014. “P. F. Strawson’s Consequentialism,” in Oxford Studies in Agency and Responsibility , 2: 64–92.
  • McKenna, Michael, 2003. “Robustness, Control, and the Demand for Morally Significant Alternatives: Frankfurt Examples with Oodles and Oodles of Alternatives,” in Moral Responsibility and Alternative Possibilities: Essays on the Importance of Alternative Possibilities , David Widerker and Michael McKenna (eds.), Burlington, VT: Ashgate Publishing, 201–18.
  • –––, 2008. “A Hard-Line Reply to Pereboom’s Four-Case Manipulation Argument,” Philosophy and Phenomenological Research , 77: 142–59.
  • –––, 2012. Conversation & Responsibility , New York: Oxford University Press.
  • –––, 2013. “Reasons-Responsiveness, Agents, and Mechanisms,” in Oxford Studies in Agency and Responsibility , 1: 151–83.
  • –––, 2014. “Resisting the Manipulation Argument: A Hard‐Liner Takes It on the Chin,” Philosophy and Phenomenological Research , 89: 467–84.
  • Mele, Alfred R., 1992. Springs of Action , New York: Oxford University Press.
  • –––, 1995. Autonomous Agents , New York: Oxford University Press.
  • –––, 2000. “Goal-directed Action: Teleological Explanations, Causal Theories, and Deviance,” Philosophical Perspectives , 14: 279–300.
  • –––, 2003. “Agents’ Abilities,” Noûs , 37: 447–70.
  • –––, 2006. Free Will and Luck , Oxford: Oxford University Press.
  • –––, 2009. Effective Intentions: The Power of Conscious Will , Oxford: Oxford University Press.
  • –––, 2017. Aspects of Agency: Decisions, Abilities, Explanations, and Free Will , New York: Oxford University Press.
  • Mele, Alfred R. and David Robb, 1998. “Rescuing Frankfurt-style Scenarios,” Philosophical Review , 107: 97–112.
  • –––, 2003. “Bbs, Magnets and Seesaws: The Metaphysics of Frankfurt-style Cases,” in Moral Responsibility and Alternative Possibilities: Essays on the Importance of Alternative Possibilities , David Widerker and Michael McKenna (eds.), Burlington, VT: Ashgate Publishing, 127–38.
  • Mitchell-Yellin, Benjamin, 2015. “The Platonic Model: Statement, Clarification and Defense,” Philosophical Explorations , 18: 378–92.
  • Moore, G. E., 1912. Ethics , Oxford: Clarendon Press.
  • Moya, Carlos, 2011. “On the Very Idea of a Robust Alternative,” Critica , 43: 3–26.
  • Mudrik, Liad, Inbal Gur Arie, Yoni Amir, Yarden Shir, Pamela Hieronymi, Uri Maoz, Timothy O’Connor, Aaron Schurger, Manuel Vargas, Tillmann Vierkant, Walter Sinnott-Armstrong, and Adina Roskies, 2022. “Free Will Without Consciousness?” Trends in Cognitive Sciences , 26 (7): 555–566.
  • Murray, Michael, 1993. “Coercion and the Hiddenness of God,” American Philosophical Quarterly , 30: 27–38.
  • –––, 2002. “Deus Absconditus,” in Divine Hiddenness: New Essays , Daniel Howard-Snyder and Paul Moser (eds.), Cambridge: Cambridge University Press, 62–82.
  • Nahmias, Eddy, 2014. “Is Free Will an Illusion? Confronting Challenges from the Modern Mind Sciences,” in Moral Psychology (Volume 4: Free Will and Moral Responsibility), Walter Sinnott-Armstrong (ed.), Cambridge, MA: MIT Press, 1–25.
  • Nelkin, Dana K., 2011. Making Sense of Freedom and Responsibility , New York: Oxford University Press.
  • Nichols, Shaun, 2015. Bound: Essays on Free Will and Moral Responsibility , New York: Oxford University Press.
  • Nietzsche, Frederick, 1886 [1966]. Beyond Good and Evil , W. Kaufmann (trans.), New York: Vintage.
  • Nowell-Smith, P. H., 1948. “Free Will and Moral Responsibility,” Mind , 57: 45–61.
  • –––, 1954. “Determinists and Libertarians,” Mind , 63: 317–37.
  • Nozick, Robert, 1981. Philosophical Explanations , Cambridge, MA: Harvard University Press.
  • O’Connor, Timothy, 2000. Persons and Causes: The Metaphysics of Free Will , New York: Oxford University Press.
  • –––, 2007. “Is It All Just a Matter of Luck?” Philosophical Explorations , 10: 157–61.
  • –––, 2009a. “Agent-Causal Power,” in Dispositions and Causes , Toby Handfield (ed.), Oxford: Clarendon Press, 189–214.
  • –––, 2009b. “Degrees of Freedom,” Philosophical Explorations , 12: 119–25.
  • –––, 2011. “Agent-Causal Theories of Freedom,” in Oxford Handbook on Free Will , Robert Kane (ed.), 2nd edition, New York: Oxford University Press, 309–28.
  • –––, 2016. “Probability and Freedom,” Res Philosophica , 93: 289–93.
  • –––, 2019. “How Do We Know That We Are Free?” European Journal of Analytic Philosophy , 15 (2): 79–98.
  • –––, 2021. “Free Will in a Network of Interacting Causes,” in W. Simpson, R. Koons & J. Orr (eds.), Neo-Aristotelian Metaphysics and the Theology of Nature , London: Routledge.
  • –––, 2022. “Why The One Did Not Remain Within Itself,” in Lara Buchak and Dean Zimmerman (eds.), Oxford Studies in the Philosophy of Religion (Volume 10), 233–46.
  • Palmer, David, 2005. “New Distinctions, Same Troubles: A Reply to Haji and McKenna,” Journal of Philosophy , 102: 474–82.
  • –––, 2011. “Pereboom on Frankfurt Cases,” Philosophical Studies , 153: 261–72.
  • –––, 2013. “The Timing Objection to the Frankfurt Cases,” Erkenntnis , 78: 1011–23.
  • –––, 2014. “Deterministic Frankfurt Cases,” Synthese , 191: 3847–64.
  • –––, 2021. “Free Will and Control: A Noncausal Approach,” Synthese , 198 (10): 10043–62.
  • Pawl, Timothy and Kevin Timpe. “Incompatibilism, Sin, and Free Will in Heaven,” Faith and Philosophy , 26: 398–419.
  • Pendergraft, Garrett, 2010. “The Explanatory Power of Local Miracle Compatibilism,” Philosophical Studies , 156: 249–66.
  • Pereboom, Derk, 2001. Living Without Free Will , Cambridge: Cambridge University Press.
  • –––, 2014. Free Will, Agency, and Meaning in Life , Oxford: Oxford University Press.
  • Pink, Thomas 2017. Self-determination: The Ethics of Action , volume 1, Oxford: Oxford University Press.
  • Plato (CW/1997). Complete Works , John Cooper (ed.), Indianapolis: Hackett Publishing.
  • Pruss, Alexander, 2016. “Divine Creative Freedom,” in Oxford Studies in the Philosophy of Religion , 7: 213–38.
  • Ragland, Scott, 2006. ‘Was Descartes a Libertarian?’ in Oxford Studies in Early Modern Philosophy , 3: 57–90.
  • Reid, Thomas, 1788 [1969]. Essays on the Active Powers of the Human Mind , Baruch Brody (ed.), Cambridge, MA: MIT Press.
  • Robinson, Michael, 2014. “The Limits of Limited-Blockage Frankfurt-style Cases,” Philosophical Studies , 169: 429–46.
  • Rogers, Katherin, 2004. “Augustine’s Compatibilism,” Religious Studies , 40 (4): 415–435.
  • Roskies, Adina, 2014. “Can Neuroscience Resolve Issues about Free Will?” in Moral Psychology (Volume 4: Free Will and Moral Responsibility), Walter Sinnott-Armstrong (ed.), Cambridge, MA: MIT Press, 103–26.
  • Rowe, William, 1995. “Two Concepts of Freedom,” in Agents, Causes, and Events: Essays on Indeterminism and Free Will , Timothy O’Connor (ed.), New York: Oxford University Press, 151–71.
  • –––, 2004. Can God Be Free? , Oxford: Oxford University Press.
  • Runyan, Jason, 2018. “Agent-causal Libertarianism, Statistical Neural Laws and Wild Coincidences,” Synthese , 195 (10): 4563–4580.
  • Russell, Paul, 2010. The Riddle of Hume’s Treatise: Skepticism, Naturalism, and Irreligion , New York: Oxford University Press.
  • –––, 2017. The Limits of Free Will , New York: Oxford University Press.
  • Sartorio, Carolina, 2016. Causation and Free Will , New York: Oxford University Press.
  • Scanlon, T. M., 2008. Moral Dimensions: Permissibility, Meaning, and Blame , Cambridge, MA: Harvard University Press.
  • Schlick, Moritz, 1939. “When Is a Man Responsible?” in Problems of Ethics , tr. by David Rynin. Englewood Cliffs, NJ: Prentice-Hall Publishing.
  • Schlosser, Markus, 2014. “The Luck Argument against Event-causal Libertarianism: It Is Here to Stay,” Philosophical Studies , 167: 375–85.
  • Schopenhauer, Arthur, 1841 [1999]. Prize Essay on the Freedom of the Will , Cambridge: Cambridge University Press.
  • Schueler, G. F., 1995. Desire: Its Role in Practical Reason and the Explanation of Action , Cambridge, MA: MIT Press.
  • –––, 2003. Reasons and Purposes: Human Rationality and the Teleological Explanation of Action , New York: Oxford University Press.
  • Schurger, Aaron, Pengbo Hu, Joanna Pak, and Adina Roskies, 2021. “What Is the Readiness Potential?” Trends in Cognitive Sciences , 25 (7): 558–570.
  • Sehon, Scott R., 2005. Teleological Realism: Mind, Agency, and Explanation , Cambridge, MA: MIT Press.
  • Shabo, Seth, 2011. “Why Free Will Remains a Mystery,” Pacific Philosophical Quarterly , 92: 105–25.
  • –––, 2013. “Free Will and Mystery: Looking Past the Mind Argument,” Philosophical Studies , 162: 291–307.
  • –––, 2020. “The Two-Stage Luck Objection,” Noûs , 54 (1): 3–23.
  • Sher, George, 2006. In Praise of Blame , New York: Oxford University Press.
  • Shoemaker, David, 2003. “Caring, Identification, and Agency,” Ethics , 114: 88–118.
  • –––, 2011. “Attributability, Answerability, and Accountability: Toward a Wider Theory of Moral Responsibility,” Ethics , 121: 602–32.
  • –––, 2015. “Ecumenical Attributability,” in The Nature of Moral Responsibility: New Essays , Randolph Clarke, Michael McKenna, and Angela Smith (eds.), New York: Oxford University Press, 115–40.
  • Slote, Michael, 1982. “Selective Necessity and the Free-Will Problem,” Journal of Philosophy , 79: 5–24.
  • Smart, J. J. C., 1961. “Free-will, Praise and Blame,” Mind , 70: 291–306.
  • Smilansky, Saul, 2000. Free Will and Illusion , Oxford: Oxford University Press.
  • Smith, Angela M., 2012. “Attributability, Answerability, and Accountability: In Defense of a Unified Account,” Ethics , 122: 575–89.
  • Smith, Michael, 2003. “Rational Capacities,” in Weakness of Will and Varieties of Practical Irrationality , Sarah Stroud and Christine Tappolet (eds.), New York: Oxford University Press, 17–38.
  • Speak, Daniel, 2007. “The Impertinence of Frankfurt-Style Argument,” Philosophical Quarterly , 57: 76–95.
  • –––, 2011. “The Consequence Argument Revisited,” in The Oxford Handbook of Free Will , Robert Kane (ed.), 2nd edition, New York: Oxford University Press, 115–30.
  • Spinoza, Baruch, 1677 [1992]. The Ethics and Selected Letters , Seymour Feldman (ed.), Samuel Shirley (trans.), Indianapolis: Hackett Publishing.
  • Sripada, Chandra, 2012. “What Makes a Manipulated Agent Unfree?” Philosophy and Phenomenological Research , 85: 563–93.
  • –––, 2016. “Self-Expression: A Deep Self Theory of Moral Responsibility,” Philosophical Studies , 173: 1203–32.
  • Steward, Helen, 2012. A Metaphysics for Freedom , Oxford: Oxford University Press.
  • Stout, Rowland, 2010. “Deviant Causal Chains,” in A Companion to the Philosophy of Action , Timothy O’Connor and Constantine Sandis (eds.), Oxford: Blackwell, 159–65.
  • Strawson, Galen, 1986. Freedom and Belief , Oxford: Clarendon Press.
  • –––, 1994. “The Impossibility of Moral Responsibility,” Philosophical Studies , 75: 5–24.
  • –––, 2000. “The Unhelpfulness of Indeterminism,” Philosophy and Phenomenological Research , 60: 149–56.
  • Strawson, P. F., 1962. “Freedom and Resentment,” Proceedings of the British Academy , 48: 187–211.
  • Stump, Eleonore, 1988. “Sanctification, Hardening of the Heart, and Frankfurt’s Concept of Free Will,” Journal of Philosophy , 85: 395–420.
  • –––, 1996. “Persons: Identification and Freedom,” Philosophical Topics , 24: 183–214.
  • –––, 1999. “Moral Responsibility and Alternative Possibilities: The Flicker of Freedom,” Journal of Ethics , 3: 299–324.
  • –––, 2003. Aquinas , London: Routledge.
  • –––, 2006. “Augustine on Free Will,” in The Cambridge Companion to Augustine , Eleonore Stump and Norman Kretzmann (eds.), Cambridge: Cambridge University Press, 124–47.
  • –––, 2010. Wandering in Darkness: Narrative and the Problem of Suffering , New York: Oxford University Press.
  • Swinburne, Richard, 2013. Mind, Brain, and Free Will , Oxford: Oxford University Press.
  • Tamburro, Richard, 2017. “The Possibility and Scope of Significant Heavenly Freedom,” in Paradise Understood: New Philosophical Essays About Heaven , Ryan Byerly and Eric Silverman (eds.), Oxford: Oxford University Press, 308–28.
  • Taylor, Richard, 1966. Action and Purpose , Englewood Cliffs, NJ: Prentice Hall Publishing.
  • –––, 1974. Metaphysics , Englewood Cliffs, NJ: Prentice-Hall Publishing.
  • Timpe, Kevin, 2006. “A Critique of Frankfurt-libertarianism,” Philosophia , 34: 189–202.
  • Todd, Patrick, 2010. “A New Approach to Manipulation Arguments,” Philosophical Studies , 152: 127–33.
  • –––, 2013. “Defending (a Modified Version of) the Zygote Argument,” Philosophical Studies , 164: 189–203.
  • van Inwagen, Peter 1975. “The Incompatibility of Free Will and Determinism,” Philosophical Studies , 27: 185–99.
  • –––, 1983. An Essay on Free Will , Oxford: Oxford University Press.
  • –––, 2000. “Free Will Remains a Mystery,” Philosophical Perspectives , 14: 1–19.
  • –––, 2004. “Freedom to Break the Laws,” Midwest Studies , 28: 334–50.
  • –––, 2008. “How to Think about the Problem of Free Will,” Journal of Ethics , 12: 327–41.
  • Vargas, Manuel, 2004. “Libertarianism and Skepticism about Free Will: Some Arguments against Both,” Philosophical Topics , 32: 403–26.
  • –––, 2007. “Revisionism,” in Four Views on Free Will , John Martin Fischer, Robert Kane, Derk Pereboom, and Manuel Vargas (eds.), Malden, MA: Blackwell Publishing, 126–64.
  • –––, 2013. Building Better Beings: A Theory of Moral Responsibility , New York: Oxford University Press.
  • Velleman, J. David, 1992. “What Happens When Someone Acts?” Mind , 101: 461–81.
  • –––, 2009. How We Get Along , New York: Cambridge University Press.
  • Vicens, Leigh, 2016. “Objective Probabilities of Free Choice,” Res Philosophica , 93: 125–35.
  • Vihvelin, Kadri, 2004. “Free Will Demystified: A Dispositional Account,” Philosophical Topics , 32: 427–50.
  • –––, 2013. Causes, Laws, and Free Will: Why Determinism Doesn’t Matter , New York: Oxford University Press.
  • Vilhauer, Ben, 2012. “Taking Free Will Skepticism Seriously,” Philosophical Quarterly , 62: 833–52.
  • Wainwright, William, 1996. “Jonathan Edwards, William Rowe, and the Necessity of Creation,” in Faith, Freedom, and Rationality , Jeff Jordan and Daniel Howard-Snyder (eds.), Lanham: Rowman and Littlefield, 119–33.
  • Wallace, R. Jay, 1994. Responsibility and the Moral Sentiments , Cambridge, MA: Harvard University Press.
  • –––, 1999. “Addiction as Defect of the Will: Some Philosophical Reflections,” Law and Philosophy , 18: 621–54.
  • Waller, Bruce, 2011. Against Moral Responsibility , Cambridge, MA: MIT Press.
  • Watson, Gary, 1975. “Free Agency,” Journal of Philosophy , 72: 205–20.
  • –––, 1986. “Review of An Essay on Free Will , by Peter van Inwagen,” Philosophy and Phenomenological Research , 46: 507–22.
  • –––, 1987. “Free Action and Free Will,” Mind , 96: 154–72.
  • –––, 1996. “Two Faces of Responsibility,” Philosophical Topics , 24: 227–48.
  • Wegner, Daniel, 2002. The Illusion of Conscious Will , Cambridge, MA: MIT Press.
  • Whittle, Ann, 2010. “Dispositional Abilities,” Philosopher’s Imprint , 10 (12), Whittle 2010 available online .
  • –––, 2022. Freedom and Responsibility in Context , Oxford: Oxford University Press.
  • Widerker, David, 1995. “Libertarianism and Frankfurt’s Attack on the Principle of Alternative Possibilities,” Philosophical Review , 104: 247–61.
  • –––, 2006. “Libertarianism and the Philosophical Significance of Frankfurt Scenarios,” Journal of Philosophy , 103: 163–87.
  • Widerker, David and Michael McKenna, 2003. Moral Responsibility and Alternative Possibilities: Essays on the Importance of Alternative Possibilities , Burlington, VT: Ashgate Publishing.
  • Widerker, David and Goetz, Stewart, 2013. “Fischer against the Dilemma Defense: The Defense Prevails,” Analysis , 73: 283–95.
  • Wiggins, David 1973. “Towards a Reasonable Libertarianism,” in Essays on Freedom and Action , Ted Honderich (ed.), London: Routledge & Kegan Paul, 31–62.
  • Wolf, Susan, 1990. Freedom within Reason , New York: Oxford University Press.
  • Zagzebski, Linda, 2000. “Does Libertarian Freedom Require Alternative Possibilities?” Philosophical Perspectives , 14: 231–48.
  • Zimmerman, Dean, 2018. “Ever Better Situations and the Failure of Expression Principles,” Faith and Philosophy , 35 (4): 408–16.
How to cite this entry . Preview the PDF version of this entry at the Friends of the SEP Society . Look up topics and thinkers related to this entry at the Internet Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers , with links to its database.
  • The Determinism and Freedom Philosophy Website , edited by Ted Honderich (University College London)
  • Bibliography on Free Will , at philpapers.org.

action | agency | blame | causation: the metaphysics of | compatibilism | determinism: causal | fatalism | freedom: divine | free will: divine foreknowledge and | incompatibilism: (nondeterministic) theories of free will | incompatibilism: arguments for | moral responsibility | quantum mechanics | skepticism: about moral responsibility

Copyright © 2022 by Timothy O’Connor < toconnor @ indiana . edu > Christopher Franklin < cefranklin @ gcc . edu >

  • Accessibility

Support SEP

Mirror sites.

View this site from another server:

  • Info about mirror sites

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

Library of Congress Catalog Data: ISSN 1095-5054

  • Advanced Search
  • All new items
  • Journal articles
  • Manuscripts
  • All Categories
  • Metaphysics and Epistemology
  • Epistemology
  • Metaphilosophy
  • Metaphysics
  • Philosophy of Action
  • Philosophy of Language
  • Philosophy of Mind
  • Philosophy of Religion
  • Value Theory
  • Applied Ethics
  • Meta-Ethics
  • Normative Ethics
  • Philosophy of Gender, Race, and Sexuality
  • Philosophy of Law
  • Social and Political Philosophy
  • Value Theory, Miscellaneous
  • Science, Logic, and Mathematics
  • Logic and Philosophy of Logic
  • Philosophy of Biology
  • Philosophy of Cognitive Science
  • Philosophy of Computing and Information
  • Philosophy of Mathematics
  • Philosophy of Physical Science
  • Philosophy of Social Science
  • Philosophy of Probability
  • General Philosophy of Science
  • Philosophy of Science, Misc
  • History of Western Philosophy
  • Ancient Greek and Roman Philosophy
  • Medieval and Renaissance Philosophy
  • 17th/18th Century Philosophy
  • 19th Century Philosophy
  • 20th Century Philosophy
  • History of Western Philosophy, Misc
  • Philosophical Traditions
  • African/Africana Philosophy
  • Asian Philosophy
  • Continental Philosophy
  • European Philosophy
  • Philosophy of the Americas
  • Philosophical Traditions, Miscellaneous
  • Philosophy, Misc
  • Philosophy, Introductions and Anthologies
  • Philosophy, General Works
  • Teaching Philosophy
  • Philosophy, Miscellaneous
  • Other Academic Areas
  • Natural Sciences
  • Social Sciences
  • Cognitive Sciences
  • Formal Sciences
  • Arts and Humanities
  • Professional Areas
  • Other Academic Areas, Misc
  • Submit a book or article
  • Upload a bibliography
  • Personal page tracking
  • Archives we track
  • Information for publishers
  • Introduction
  • Submitting to PhilPapers
  • Frequently Asked Questions
  • Subscriptions
  • Editor's Guide
  • The Categorization Project
  • For Publishers
  • For Archive Admins
  • PhilPapers Surveys
  • Bargain Finder
  • About PhilPapers
  • Create an account

An Essay on Free Will

Author's profile.

an essay on free will

Reprint years

Call number, buy this book.

an essay on free will

PhilArchive

External links.

an essay on free will

  • From the Publisher via CrossRef (no proxy)
  • pq.oxfordjournals.org (no proxy)
  • jstor.org (no proxy)
  • Available at Amazon.com

Through your library

  • Sign in / register and customize your OpenURL resolver
  • Configure custom resolver

Similar books and articles

Citations of this work, references found in this work.

No references found.

Phiosophy Documentation Center

Logo for

Want to create or adapt books like this? Learn more about how Pressbooks supports open publishing practices.

30 Existence of Free Will

Determinism and freedom.

Determinism and free will are often thought to be in deep conflict. Whether or not this is true has a lot to do with what is meant by determinism and an account of what free will requires.

First of all, determinism is not the view that free actions are impossible. Rather, determinism is the view that at any one time, only one future is physically possible. To be a little more specific, determinism is the view that a complete description of the past along with a complete account of the relevant laws of nature logically entails all future events. 1

Indeterminism is simply the denial of determinism. If determinism is incompatible with free will, it will be because free actions are only possible in worlds in which more than one future is physically possible at any one moment in time. While it might be true that free will requires indeterminism, it’s not true merely by definition. A further argument is needed and this suggests that it is at least possible that people could sometimes exercise the control necessary for morally responsible action, even if we live in a deterministic world.

It is worth saying something about fatalism before we move on. It is really easy to mistake determinism for fatalism, and fatalism does seem to be in straightforward conflict with free will. Fatalism is the view that we are powerless to do anything other than what we actually do. If fatalism is true, then nothing that we try or think or intend or believe or decide has any causal effect or relevance as to what we actually end up doing.

But note that determinism need not entail fatalism. Determinism is a claim about what is logically entailed by the rules/laws governing a world and the past of said world. It is not the claim that we lack the power to do other than what we actually were already going to do. Nor is it the view that we fail to be an important part of the causal story for why we do what we do. And this distinction may allow some room for freedom, even in deterministic worlds.

An example will be helpful here. We know that the boiling point for water is 100°C. Suppose we know in both a deterministic world and a fatalistic world that my pot of water will be boiling at 11:22am today. Determinism makes the claim that if I take a pot of water and I put it on my stove, and heat it to 100°C, it will boil. This is because the laws of nature (in this case, water that is heated to 100°C will boil) and the events of the past (I put a pot of water on a hot stove) bring about the boiling water. But fatalism makes a different claim. If my pot of water is fated to boil at 11:22am today, then no matter what I or anyone does, my pot of water will boil at exactly 11:22am today. I could try to empty the pot of water out at 11:21. I could try to take the pot as far away from a heating source as possible. Nonetheless, my pot of water will be boiling at 11:22 precisely because it was fated that this would happen. Under fatalism, the future is fixed or preordained, but this need not be the case in a deterministic world. Under determinism, the future is a certain way because of the past and the rules governing said world. If we know that a pot of water will boil at 11:22am in a deterministic world, it’s because we know that the various causal conditions will hold in our world such that at 11:22 my pot of water will have been put on a heat source and brought to 100°C. Our deliberations, our choices, and our free actions may very well be part of the process that brings a pot of water to the boiling point in a deterministic world, whereas these are clearly irrelevant in fatalistic ones.

Three Views of Freedom

Most accounts of freedom fall into one of three camps. Some people take freedom to require merely the ability to “do what you want to do.” For example, if you wanted to walk across the room, right now, and you also had the ability, right now, to walk across the room, you would be free as you could do exactly what you want to do. We will call this easy freedom.

Others view freedom on the infamous “Garden of Forking Paths” model. For these people, free action requires more than merely the ability to do what you want to do. It also requires that you have the ability to do otherwise than what you actually did. So, If Anya is free when she decides to take a sip from her coffee, on this view, it must be the case that Anya could have refrained from sipping her coffee. The key to freedom, then, is alternative possibilities and we will call this the alternative possibilities view of free action.

Finally, some people envision freedom as requiring, not alternative possibilities but the right kind of relationship between the antecedent sources of our actions and the actions that we actually perform. Sometimes this view is explained by saying that the free agent is the source, perhaps even the ultimate source of her action. We will call this kind of view a source view of freedom.

Now, the key question we want to focus on is whether or not any of these three models of freedom are compatible with determinism. It could turn out that all three kinds of freedom are ruled out by determinism, so that the only way freedom is possible is if determinism is false. If you believe that determinism rules out free action, you endorse a view called incompatibilism. But it could turn out that one or all three of these models of freedom are compatible with determinism. If you believe that free action is compatible with determinism, you are a compatibilist.

Let us consider compatibilist views of freedom and two of the most formidable challenges that compatibilists face: the consequence argument and the ultimacy argument.

Begin with easy freedom. Is easy freedom compatible with determinism? A group of philosophers called classic compatibilists certainly thought so. 2  They argued that free will requires merely the ability for an agent to act without external hindrance. Suppose, right now, you want to put your textbook down and grab a cup of coffee. Even if determinism is true, you probably, right now, can do exactly that. You can put your textbook down, walk to the nearest Starbucks, and buy an overpriced cup of coffee. Nothing is stopping you from doing what you want to do. Determinism does not seem to be posing any threat to your ability to do what you want to do right now. If you want to stop reading and grab a coffee, you can. But, by contrast, if someone had chained you to the chair you are sitting in, things would be a bit different. Even if you wanted to grab a cup of coffee, you would not be able to. You would lack the ability to do so. You would not be free to do what you want to do. This has nothing to do with determinism, of course. It is not the fact that you might be living in a deterministic world that is threatening your free will. It is that an external hindrance (the chains holding you to your chair) is stopping from you doing what you want to do. So, if what we mean by freedom is easy freedom, it looks like freedom really is compatible with determinism.

Easy freedom has run into some rather compelling opposition, and most philosophers today agree that a plausible account of easy freedom is not likely. But, by far, the most compelling challenge the view faces can be seen in the consequence argument. 3  The consequence argument is as follows:

  • If determinism is true, then all human actions are consequences of past events and the laws of nature.
  • No human can do other than they actually do except by changing the laws of nature or changing the past.
  • No human can change the laws of nature or the past.
  • If determinism is true, no human has free will.

This is a powerful argument. It is very difficult to see where this argument goes wrong, if it goes wrong. The first premise is merely a restatement of determinism. The second premise ties the ability to do otherwise to the ability to change the past or the laws of nature, and the third premise points out the very reasonable assumption that humans are unable to modify the laws of nature or the past.

This argument effectively devastates easy freedom by proposing that we never act without external hindrances precisely because our actions are caused by past events and the laws of nature in such a way that we not able to contribute anything to the causal production of our actions. This argument also seems to pose a deeper problem for freedom in deterministic worlds. If this argument works, it establishes that, given determinism, we are powerless to do otherwise, and to the extent that freedom requires the ability to do otherwise, this argument seems to rule out free action. Note that if this argument works, it poses a challenge for both the easy and alternative possibilities view of free will.

How might someone respond to this argument? First, suppose you adopt an alternative possibilities view of freedom and believe that the ability to do otherwise is what is needed for genuine free will. What you would need to show is that alternative possibilities, properly understood, are not incompatible with determinism. Perhaps you might argue that if we understand the ability to do otherwise properly we will see that we actually do have the ability to change the laws of nature or the past.

That might sound counterintuitive. How could it possibly be the case that a mere mortal could change the laws of nature or the past? Think back to Quinn’s decision to spend the night before her exam out with friends instead of studying. When she shows up to her exam exhausted, and she starts blaming herself, she might say, “Why did I go out? That was dumb! I could have stayed home and studied.” And she is sort of right that she could have stayed home. She had the general ability to stay home and study. It is just that if she had stayed home and studied the past would be slightly different or the laws of nature would be slightly different. What this points to is that there might be a way of cashing out the ability to do otherwise that is compatible with determinism and does allow for an agent to kind of change the past or even the laws of nature. 4

But suppose we grant that the consequence argument demonstrates that determinism really does rule out alternative possibilities. Does that mean we must abandon the alternative possibilities view of freedom? Well, not necessarily. You could instead argue that free will is possible, provided determinism is false. 5  That is a big if, of course, but maybe determinism will turn out to be false.

What if determinism turns out to be true? Should we give up, then, and concede that there is no free will? Well, that might be too quick. A second response to the consequence argument is available. All you need to do is deny that freedom requires the ability to do otherwise.

In 1969, Harry Frankfurt proposed an influential thought experiment that demonstrated that free will might not require alternative possibilities at all (Frankfurt [1969] 1988). If he’s right about this, then the consequence argument, while compelling, does not demonstrate that no one lacks free will in deterministic worlds, because free will does not require the ability to do otherwise. It merely requires that agents be the source of their actions in the right kind of way. But we’re getting ahead of ourselves. Here is a simplified paraphrase of Frankfurt’s case:

Black wants Jones to perform a certain action. Black is prepared to go to considerable lengths to get his way, but he prefers to avoid unnecessary work. So he waits until Jones is about to make up his mind what to do, and he does nothing unless it is clear to him (Black is an excellent judge of such things) that Jones is going to decide not to do what Black wants him to do. If it does become clear that Jones is going to decide to do something other than what Black wanted him to do, Black will intervene, and ensure that Jones decides to do, and does do, exactly what Black wanted him to do. Whatever Jones’ initial preferences and inclinations, then, Black will have his way. As it turns out, Jones decides, on his own, to do the action that Black wanted him to perform. So, even though Black was entirely prepared to intervene, and could have intervened, to guarantee that Jones would perform the action, Black never actually has to intervene because Jones decided, for reasons of his own, to perform the exact action that Black wanted him to perform. (Frankfurt [1969] 1988, 6-7)

Now, what is going on here? Jones is overdetermined to perform a specific act. No matter what happens, no matter what Jones initially decides or wants to do, he is going to perform the action Black wants him to perform. He absolutely cannot do otherwise. But note that there seems to be a crucial difference between the case in which Jones decides on his own and for his own reasons to perform the action Black wanted him to perform and the case in which Jones would have refrained from performing the action were it not for Black intervening to force him to perform the action. In the first case, Jones is the source of his action. It the thing he decided to do and he does it for his own reasons. But in the second case, Jones is not the source of his actions. Black is. This distinction, thought Frankfurt, should be at the heart of discussions of free will and moral responsibility. The control required for moral responsibility is not the ability to do otherwise (Frankfurt [1969] 1988, 9-10).

If alternative possibilities are not what free will requires, what kind of control is needed for free action? Here we have the third view of freedom we started with: free will as the ability to be the source of your actions in the right kind of way. Source compatibilists argue that this ability is not threatened by determinism, and building off of Frankfurt’s insight, have gone on to develop nuanced, often radically divergent source accounts of freedom. 6  Should we conclude, then, that provided freedom does not require alternative possibilities that it is compatible with determinism? 7 Again, that would be too quick. Source compatibilists have reason to be particularly worried about an argument developed by Galen Strawson called the ultimacy argument (Strawson [1994] 2003, 212-228).

Rather than trying to establish that determinism rules out alternative possibilities, Strawson tried to show that determinism rules out the possibility of being the ultimate source of your actions. While this is a problem for anyone who tries to establish that free will is compatible with determinism, it is particularly worrying for source compatibilists as they’ve tied freedom to an agent’s ability to be source of its actions. Here is the argument:

  • A person acts of her own free will only if she is the act’s ultimate source.
  • If determinism is true, no one is the ultimate source of her actions.
  • Therefore, if determinism is true, no one acts of her own free will. (McKenna and Pereboom 2016, 148) 8

This argument requires some unpacking. First of all, Strawson argues that for any given situation, we do what we do because of the way we are ([1994] 2003, 219). When Quinn decides to go out with her friends rather than study, she does so because of the way she is. She prioritizes a night with her friends over studying, at least on that fateful night before her exam. If Quinn had stayed in and studied, it would be because she was slightly different, at least that night. She would be such that she prioritized studying for her exam over a night out. But this applies to any decision we make in our lives. We decide to do what we do because of how we already are.

But if what we do is because of the way we are, then in order to be responsible for our actions, we need to be the source of how we are, at least in the relevant mental respects (Strawson [1994] 2003, 219). There is the first premise. But here comes the rub: the way we are is a product of factors beyond our control such as the past and the laws of nature ([1994] 2003, 219; 222-223). The fact that Quinn is such that she prioritizes a night with friends over studying is due to her past and the relevant laws of nature. It is not up to her that she is the way she is. It is ultimately factors extending well beyond her, possibly all the way back to the initial conditions of the universe that account for why she is the way she is that night. And to the extent that this is compelling, the ultimate source of Quinn’s decision to go out is not her. Rather, it is some condition of the universe external to her. And therefore, Quinn is not free.

Once again, this is a difficult argument to respond to. You might note that “ultimate source” is ambiguous and needing further clarification. Some compatibilists have pointed this out and argued that once we start developing careful accounts of what it means to be the source of our actions, we will see that the relevant notion of source-hood is compatible with determinism.

For example, while it may be true that no one is the ultimate cause of their actions in deterministic worlds precisely because the ultimate source of all actions will extend back to the initial conditions of the universe, we can still be a mediated source of our actions in the sense required for moral responsibility. Provided the actual source of our action involves a sophisticated enough set of capacities for it to make sense to view us as the source of our actions, we could still be the source of our actions, in the relevant sense (McKenna and Pereboom 2016, 154). After all, even if determinism is true, we still act for reasons. We still contemplate what to do and weigh reasons for and against various actions, and we still are concerned with whether or not the actions we are considering reflect our desires, our goals, our projects, and our plans. And you might think that if our actions stem from a history that includes us bringing all the features of our agency to bear upon the decision that is the proximal cause of our action, that this causal history is one in which we are the source of our actions in the way that is really relevant to identifying whether or not we are acting freely.

Others have noted that even if it is true that Quinn is not directly free in regard to the beliefs and desires that suggest she should go out with her friends rather than study (they are the product of factors beyond her control such as her upbringing, her environment, her genetics, or maybe even random luck), this need not imply that she lacks control as to whether or not she chooses to act upon them. 9  Perhaps it is the case that even though how we are may be due to factors beyond our control, nonetheless, we are still the source of what we do because it is still, even under determinism, up to us as to whether we choose to exercise control over our conduct.

Free Will and the Sciences

Many challenges to free will come, not from philosophy, but from the sciences. There are two main scientific arguments against free will, one coming from neuroscience and one coming from the social sciences. The concern coming from research in the neurosciences is that some empirical results suggest that all our choices are the result of unconscious brain processes, and to the extent choices must be consciously made to be free choices, it seems that we never make a conscious free choice.

The classic studies motivating a picture of human action in which unconscious brain processes are doing the bulk of the causal work for action were conducted by Benjamin Libet. Libet’s experiments involved subjects being asked to flex their wrists whenever they felt the urge to do so. Subjects were asked to note the location of a clock hand on a modified clock when they became aware of the urge to act. While doing this their brain activity was being scanned using EEG technology. What Libet noted is that around 550 milliseconds before a subject acted, a readiness potential (increased brain activity) would be measured by the EEG technology. But subjects were reporting awareness of an urge to flex their wrist around 200 milliseconds before they acted (Libet 1985).

This painted a strange picture of human action. If conscious intentions were the cause of our actions, you may expect to see a causal story in which the conscious awareness of an urge to flex your wrist shows up first, then a ramping up of brain activity, and finally an action. But Libet’s studies showed a causal story in which an action starts with unconscious brain activity, the subject later becomes consciously aware that they are about to act, and then the action happens. The conscious awareness of action seemed to be a byproduct of the actual unconscious process that was causing the action. It was not the cause of the action itself. And this result suggests that unconscious brain processes, not conscious ones, are the real causes of our actions. To the extent that free action requires our conscious decisions to be the initiating causes of our actions, it looks like we may never act freely.

While this research is intriguing, it probably does not establish that we are not free. Alfred Mele is a philosopher who has been heavily critical of these studies. He raises three main objections to the conclusions drawn from these arguments.

First, Mele points out that self-reports are notoriously unreliable (2009, 60-64). Conscious perception takes time, and we are talking about milliseconds. The actual location of the clock hand is probably much closer to 550 milliseconds when the agent “intends” or has the “urge” to act than it is to 200 milliseconds. So, there’s some concerns about experimental design here.

Second, an assumption behind these experiments is that what is going on at 550 milliseconds is that a decision is being made to flex the wrist (Mele 2014, 11). We might challenge this assumption. Libet ran some variants of his experiment in which he asked subjects to prepare to flex their wrist but to stop themselves from doing so. So, basically, subjects simply sat there in the chair and did nothing. Libet interpreted the results of these experiments as showing that we might not have a free will, but we certainly have a “free won’t” because we seem capable of consciously vetoing or stopping an action, even if that action might be initiated by unconscious processes (2014, 12-13). Mele points out that what might be going on in these scenarios is that the real intention to act or not act is what happens consciously at 200 milliseconds, and if so, there is little reason to think these experiments are demonstrating that we lack free will (2014, 13).

Finally, Mele notes that while it may be the case that some of our decisions and actions look like the wrist-flicking actions Libet was studying, it is doubtful that all or even most of our decisions are like this (2014, 15). When we think about free will, we rarely think of actions like wrist-flicking. Free actions are typically much more complex and they are often the kind of thing where the decision to do something extends across time. For example, your decision about what to major in at college or even where to study was probably made over a period of months, even years. And that decision probably involved periods of both conscious and unconscious cognition. Why think that a free choice cannot involve some components that are unconscious?

A separate line of attack on free will comes from the situationist literature in the social sciences (particularly social psychology). There is a growing body of research suggesting that situational and environmental factors profoundly influence human behavior, perhaps in ways that undermine free will (Mele 2014, 72).

Many of the experiments in the situationist literature are among the most vivid and disturbing in all of social psychology. Stanley Milgram, for example, conducted a series of experiments on obedience in which ordinary people were asked to administer potentially lethal voltages of electricity to an innocent subject in order to advance scientific research, and the vast majority of people did so! 10  And in Milgram’s experiments, what affected whether or not subjects were willing to administer the shocks were minor, seemingly insignificant environmental factors such as whether the person running the experiment looked professional or not (Milgram 1963).

What experiments like Milgram’s obedience experiments might show is that it is our situations, our environments that are the real causes of our actions, not our conscious, reflective choices. And this may pose a threat to free will. Should we take this kind of research as threatening freedom?

Many philosophers would resist concluding that free will does not exist on the basis of these kinds of experiments. Typically, not everyone who takes part in situationist studies is unable to resist the situational influences they are subject to. And it appears to be the case that when we are aware of situational influences, we are more likely to resist them. Perhaps the right way to think about this research is that there all sorts of situations that can influence us in ways that we may not consciously endorse, but that nonetheless, we are still capable of avoiding these effects when we are actively trying to do so. For example, the brain sciences have made many of us vividly aware of a whole host of cognitive biases and situational influences that humans are typically subject to and yet, when we are aware of these influences, we are less susceptible to them. The more modest conclusion to draw here is not that we lack free will, but that exercising control over our actions is much more difficult than many of us believe it to be. We are certainly influenced by the world we are a part of, but to be influenced by the world is different from being determined by it, and this may allow us to, at least sometimes, exercise some control over the actions we perform.

No one knows yet whether or not humans sometimes exercise the control over their actions required for moral responsibility. And so I leave it to you, dear reader: Are you free?

Chapter Notes

  • I have hidden some complexity here. I have defined determinism in terms of logical entailment. Sometimes people talk about determinism as a causal relationship. For our purposes, this distinction is not relevant, and if it is easier for you to make sense of determinism by thinking of the past and the laws of nature causing all future events, that is perfectly acceptable to do.
  • Two of the more well-known classic compatibilists include Thomas Hobbes and David Hume. See: Hobbes, Thomas, (1651) 1994, Leviathan , ed. Edwin Curley, Canada: Hackett Publishing Company; and Hume, David, (1739) 1978, A Treatise of Human Nature , Oxford: Oxford University Press.
  • For an earlier version of this argument see: Ginet, Carl, 1966, “Might We Have No Choice?” in Freedom and Determinism, ed. Keith Lehrer, 87-104, Random House.
  • For two notable attempts to respond to the consequence argument by claiming that humans can change the past or the laws of nature see: Fischer, John Martin, 1994, The Metaphysics of Free Will , Oxford: Blackwell Publishers; and Lewis, David, 1981, “Are We Free to Break the Laws?” Theoria 47: 113-21.
  • Many philosophers try to develop views of freedom on the assumption that determinism is incompatible with free action. The view that freedom is possible, provided determinism is false is called Libertarianism. For more on Libertarian views of freedom, see: Clarke, Randolph and Justin Capes, 2017, “Incompatibilist (Nondeterministic) Theories of Free Will,” Stanford Encyclopedia of Philosophy , https://plato.stanford.edu/entries/incompatibilism-theories/ .
  • For elaboration on recent compatibilist views of freedom, see McKenna, Michael and D. Justin Coates, 2015, “Compatibilism,” Stanford Encyclopedia of Philosophy , https://plato.stanford.edu/entries/compatibilism/ .
  • You might be unimpressed by the way source compatibilists understand the ability to be the source of your actions. For example, you might that what it means to be the source of your actions is to be the ultimate cause of your actions. Or maybe you think that to genuinely be the source of your actions you need to be the agent-cause of your actions. Those are both reasonable positions to adopt. Typically, people who understand free will as requiring either of these abilities believe that free will is incompatible with determinism. That said, there are many Libertarian views of free will that try to develop a plausible account of agent causation. These views are called Agent-Causal Libertarianism. See: Clarke, Randolph and Justin Capes, 2017, “Incompatibilist (Nondeterministic) Theories of Free Will,” Stanford Encyclopedia of Philosophy , https://plato.stanford.edu/entries/incompatibilism-theories/ .
  • As with most philosophical arguments, the ultimacy argument has been formulated in a number of different ways. In Galen Strawson’s original paper he gives three different versions of the argument, one of which has eight premises and one that has ten premises. A full treatment of either of those versions of this argument would require more time and space than we have available here. I have chosen to use the McKenna/Pereboom formulation of the argument due its simplicity and their clear presentation of the central issues raised by the argument.
  • For two attempts to respond to the ultimacy argument in this way, see: Mele, Alfred, 1995, Autonomous Agents , New York: Oxford University Press; and McKenna, Michael, 2008, “Ultimacy & Sweet Jane” in Nick Trakakis and Daniel Cohen, eds, Essays on Free Will and Moral Responsibility , Newcastle: Cambridge Scholars Publishing: 186-208.
  • Fortunately, no real shocks were administered. The subjects merely believed they were doing so.

Frankfurt, Harry. (1969) 1988. “Alternative Possibilities and moral responsibility.” In The Importance of What We Care About: Philosophical Essays , 10th ed. New York: Cambridge University Press.

Libet, Benjamin. 1985. “Unconscious Cerebral Initiative and the Role of Conscious Will in Voluntary Action.” Behavioral and Brain Sciences 8: 529-566.

McKenna, Michael and Derk Pereboom. 2016. Free Will: A Contemporary Introduction. New York: Routledge.

Mele, Alfred. 2014. Free: Why Science Hasn’t Disproved Free Will . Oxford: Oxford University Press.

Mele, Alfred. 2009. Effective Intentions: The Power of Conscious Will. Oxford: Oxford University Press.

Milgram, Stanley. 1963. “Behavioral Study of Obedience.” The Journal of Abnormal and Social Psychology 67: 371-378.

Strawson, Galen. (1994) 2003. “The Impossibility of Moral Responsibility.” In Free Will, 2nd ed. Edited by Gary Watson, 212-228. Oxford: Oxford University Press.

van Inwagen, Peter. 1983. An Essay on Free Will. Oxford: Clarendon Press.

Further Reading

Deery, Oisin and Paul Russell, eds. 2013. The Philosophy of Free Will: Essential Readings from the Contemporary Debates . New York: Oxford University Press.

Mele, Alfred. 2006. Free Will and Luck. New York: Oxford University Press.

Attribution

This section is composed of text taken from Chapter 8 Freedom of the Will created by Daniel Haas in Introduction to Philosophy: Philosophy of Mind , edited by Heather Salazar and Christina Hendricks, and produced with support from the Rebus Community. The original is freely available under the terms of the CC BY 4.0 license at https://press.rebus.community/intro-to-phil-of-mind/ . The material is presented in its original form, with the exception of the removal of introductory material Introduction: Are We Free?

A Brief Introduction to Philosophy Copyright © 2021 by Southern Alberta Institution of Technology (SAIT) is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Share This Book

Free Will in an Indeterministic World?

Van Inwagen’s Discussion of the Mind-Argument and a ‘Scientistic’ Critique of Free Will Revisited

  • First Online: 13 February 2018

Cite this chapter

Book cover

  • Alfonso Savarino 3 &
  • Annegret Sock 3  

Part of the book series: Münster Lectures in Philosophy ((MUELP,volume 4))

349 Accesses

In his An Essay on Free Will Peter van Inwagen pursues two main strategies in order to argue for the conceivability of free will in an indeterministic world: discussing the Mind-argument and rejecting a so called ‘scientistic’ critique of free will. By analyzing the decision-making process, we pointedly formulate the problems which are fundamental to both of these strategies against the existence of free will, clarify how they are related, and try to identify the requirements of possible solutions to these problems. We conclude that in either case a person must be credited with a certain kind of causal ability if her act is supposed to be free. This type of causality is problematic for various reasons as it differs from our usual notions of causality and as it seems to be incompatible with our current physical theories. Although we agree in many ways with van Inwagen’s arguments, we are less skeptical about possible answers to the Mind-argument. But when we look at the free will debate from a rather physical point of view, the problems are, in our opinion, more severe than van Inwagen might be willing to admit.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
  • Durable hardcover edition

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Cf. van Inwagen 1983 , 191: “ Science shows determinism to be true. This simply does not seem to be the case. If anything, quite the opposite is true, since the standard interpretation of quantum mechanics is indeterministic. This bald statement requires a variety of qualifications and comments.”

Cf. van Inwagen 1983 , 209: “It is this and only this, I think, that provides us with a reason for believing in free will.”

In this paper we will use the same notion for free will that Peter van Inwagen adheres to: “When I say of a man that he ‘has free will’ I mean that very often, if not always, when he has to choose between two or more mutually incompatible courses of action – that is, courses of action that it is impossible for him to carry out more than one of – each of these courses of action is such that he can, or is able to, or has it within his power to carry it out.” (van Inwagen 1983 , 8)

Van Inwagen uses the term “scientistic” to describe an argument against free will. However, to us this term seems unjustified since the argument arises from our scientific and not scientistic understanding of the physical world, as will be discussed below.

Cf. Mind-argument strand (i) in van Inwagen 1983 , 128–129: “The doubtful premise in this argument is the assertion that if our acts are undetermined they are mere ‘random’ or ‘chance’ events. What does this mean? The words ‘random’ and ‘chance’ most naturally apply to patterns or sequences of events.”

See Mind-argument strand (ii) in van Inwagen 1983 , 129: “But, as we shall discover when we discuss the second strand, an undetermined act need not be an uncaused act.”

See the second model of action in van Inwagen 1983 , 137–150.

One could try to argue that the person might still be able to influence the states that are present at t 1 , for example, by influencing the attitudes one holds. But this would only shift the problem to a prior situation and it could be asked how the person can influence her attitudes.

van Inwagen 1983 , 126: “Incompatibilists maintain that free will requires indeterminism. But it should be clear even to them that not just any sort of indeterminism will do […] If the question whether there are any undetermined events is relevant to the question whether we have free will, this can only be because the question whether there are undetermined events that shape or influence our acts is relevant to the question whether we have free will. Therefore, the incompatibilist must believe not only that free will entails that there are undetermined events; he must believe that free will entails that there are undetermined events that shape our behaviour. And not just any such event will do for his purposes.”

Cf. van Inwagen 2015 , 277–281. There may be other theories of human action that might fulfill these requirements, see Clarke and Capes 2015 .

Here we cannot provide a full characterization of causation. However, we use the notion that some x causes y or that x has a causal influence on y if x is in some way or other a reason why y obtains. This means that anything that contributes to the determination of some event is at least a contributing cause to that event. So if, for example, in the case outlined in M3 the person has a choice about whether the input and the system cause the menu B to be chosen, the person is a cause of that decision at t 2 . It should be noted that this seems to be another kind of causality than between the input and the system on one side and the output on the other side since the person would influence the causal relationship between them.

Cf. van Inwagen 2000 , 16–18, and van Inwagen 2013 , 216: “[…]if it is undetermined whether that agent will lie or tell the truth, then it is undetermined whether that agent will agent-cause ‘lie’ or agent-cause ‘tell the truth’.”

Note that this does not have to mean that the ratio will just settle at 33,3% for every menu: For example, she might choose menu A 10 times, then menu B in the following 20 replays and then menu C 40 times and so on. Or the ratio will be 33,3% after, say, a thousand replays, but will change to a ratio of 50:30:20 in the next thousand replays and so on. Also, her behavior does not need to be just random in the way of a brute fact as outlined above. In fact, our analysis shows that such an undetermined choice does not have to be a random event, a brute fact or chance, but can be the decision of the person.

Cf. van Inwagen 1983 , 223: “[…] and indeed, there is something objectionable about incompatibilism. The incompatibilist must either reject the free-will thesis or else accept the thesis that an agent often has a choice about whether a certain one of his inner states is followed by a certain overt act (the state being the cause of the act), even though neither the inner state nor anything else determines that act to occur.” See also footnote 22 on the same page: “This dilemma, strictly speaking, faces only the incompatibilist who accepts our second model of action. But some such dilemma will face any incompatibilist who holds any very explicit theory of action.”

It may be that those fundamentally different causal relationships will still be compatible with our current understanding if, for example, these additional properties affect the quantum indeterminacy and are therefore embedded in our current theories. But any theory of free will that would claim this behavior bears the problem that all free act must somehow still uphold the probabilities set by quantum physics. So in our example where Lucy is in exactly the same situation 100 times, she may choose menu B 95 times, but she would be forced to choose menu C 5 times if those are the probabilities set by quantum physics. In addition, this influence on quantum events might still be called a new kind of causal relationship.

Being a physicalist, van Inwagen does, of course, not distinguish these levels in the way we do.

Cf. van Inwagen 1983 , 215: “I am convinced that in a large number of cases the answer is that the people who regard my central thesis as simply incredible are victims of scientism. Scientism, as I use the word, is a sort of exaggerated respect for science – that is, for the physical and biological sciences – and a corresponding disparagement of all other areas of human intellectual endeavour.”

One speculative theory that could achieve this is the theory of top-down causation that claims that the whole of a system causally affects its parts while being somewhat independent from them.

Clarke, Randolph, and Justin Capes. 2015. Incompatibilist (nondeterministic) theories of free will. In The Stanford encyclopedia of philosophy (Fall 2015 Edition), ed. Edward N. Zalta. http://plato.stanford.edu/archives/fall2015/entries/incompatibilism-theories/ .

van Inwagen, Peter. 1983. An essay on free will . Oxford: University Press.

Google Scholar  

———. 2000. Free will remains a mystery. Philosophical Perspectives 14: 1–19.

———. 2008. How to think about the problem of free will? The Journal of Ethics 12: 327–341.

Article   Google Scholar  

———. 2013. A dialogue on free will. Methode 3: 212–221.

———. 2015. Metaphysics . Boulder: Westview.

Download references

Author information

Authors and affiliations.

Faculty for Catholic Theology, Ruhr University, Bochum, Germany

Alfonso Savarino & Annegret Sock

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Alfonso Savarino .

Editor information

Editors and affiliations.

Ludger Jansen

Department of Philosophy, WWU Münster, Münster, Germany

Paul M. Näger

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG

About this chapter

Savarino, A., Sock, A. (2018). Free Will in an Indeterministic World?. In: Jansen, L., Näger, P. (eds) Peter van Inwagen. Münster Lectures in Philosophy, vol 4. Springer, Cham. https://doi.org/10.1007/978-3-319-70052-6_8

Download citation

DOI : https://doi.org/10.1007/978-3-319-70052-6_8

Published : 13 February 2018

Publisher Name : Springer, Cham

Print ISBN : 978-3-319-70051-9

Online ISBN : 978-3-319-70052-6

eBook Packages : Religion and Philosophy Philosophy and Religion (R0)

Share this chapter

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research

Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

Enter the email address you signed up with and we'll email you a reset link.

  • We're Hiring!
  • Help Center

paper cover thumbnail

An Essay on Free will and Determinism

Profile image of Munawar Sarker

Related Papers

Paul Carron

What are the freedom-relevant conditions necessary for someone to be a morally responsible person? I examine several key authors beginning with Harry Frankfurt that have contributed to this debate in recent years, and then look back to the writings or Søren Kierkegaard to provide a solution to the debate. In this project I investigate the claims of semi-compatibilism and argue that while its proponents have identified a fundamental question concerning free will and moral responsibility—namely, that the agential properties necessary for moral responsibility ascriptions are found in scenarios where the agent acts on her own as opposed to her action resulting from freedom undermining external causes such as manipulation, phobias, etc.—they have failed to show that the freedom-relevant agential properties identified in those actual-sequence scenarios are compatible with causal determinism. My argument is that only a voluntarist-libertarian theory can adequately account for the kinds of cases that the semicompatibilist identify. I argue that there are three freedom-relevant conditions necessary for someone to be a morally responsible person: a hierarchical understanding of human desires [specifically and mental states generally], an incompatibilist (non-deterministic) understanding of human action, and a historical understanding of character development. The ability to reflect critically about one’s own desires and emotions, and thus to have a kind of self-knowledge and understanding with regard to the springs of one’s own actions, is required to make it possible for the agent to be the “source” of her own actions and character. The non-deterministic understanding of human action is needed for a similar reason: if determinism is true, then every action a person performs can be ultimately traced to and exhaustively explained in terms of factors outside the agent’s control, thus making the agent’s responsibility for his actions an illusion. And finally,human nature must be such that, over time, one’s choices leave a dispositional residue self-understanding and motivation in the person’s self, out of which, in mature understanding and motivation, the person acts as a fully responsible agent.

an essay on free will

Matthew Talbert

Most people would agree that a small child, or a cognitively impaired adult, is less responsible for their actions, good or bad, than an unimpaired adult. But how do we explain this difference, and how far can anyone be praised or blamed for what they have done? This introductory text explores some of the key questions shaping current philosophical debates about moral responsibility, including: • What is free will and is it required for moral responsibility? • Can a bad upbringing undermine blameworthiness? • Can we be blamed for having bad characters? • Is it fair to blame people for doing what they believe is right? • Are psychopaths open to blame? • Are there grounds for skepticism about moral responsibility?

Christian Rostboll

Christian Rostbøll

Behavioral Sciences & the Law

Patrick Grim

Rick Repetti

I argue that central Buddhist tenets and meditation methodology support a view of free will similar to Harry Frankfurt’s optimistic view and contrary to Galen Strawson’s pessimistic view. For Frankfurt, free will involves a relationship between actions, voli- tions, and “metavolitions” (volitions about volitions): simplifying greatly, volitional actions are free if the agent approves of them. For Buddhists, mental freedom involves a relationship between mental states and “metamental” states (mental attitudes toward mental states): simplifying greatly, one has mental freedom if one is able to control one’s mental states, and to the extent one has mental freedom when choosing, one has free will. Philosophical challenges to free will typically question whether it is compatible with “determinism,” the thesis of lawful universal causation. Both Frankfurt’s metavolitional approval and the Buddhist’s me- tamental control are consistent with determinism. Strawson has argued, however, that free will is impossible, determinism not- withstanding, because one’s choice is always influenced by one’smental state. I argue, however, that Buddhist meditation culti- vates control over mental states that undermine freedom, whether they are deterministic or not, making both mental free- dom and free will possible. The model I develop is only a sketch of a minimally risky theory of free will, but one that highlights the similarities and differences between Buddhist thought on this subject and relevantly-related Western thought and has explana- tory promise.

matt osolinski

Matteo Grasso

The problem of free will is deeply linked with the causal relevance of mental events. The causal exclusion argument claims that, in order to be causally relevant, mental events must be identical to physical events. However, Gibb has recently criticized it, suggesting that mental events are causally relevant as double preventers. For Gibb, mental events enable physical effects to take place by preventing other mental events from preventing a behaviour to take place. The role of mental double preventers is hence similar to what Libet names free won’t, namely the ability to veto an action initiated unconsciously by the brain. In this paper I will propose an argument against Gibb’s account, the causal irrelevance argument, showing that Gibb’s proposal does not overcome the objection of systematic overdetermination of causal relevance, because mental double preventers systematically overdetermine physical double preventers, and therefore mental events are causally irrelevant.

Journal of Philosophy of Education

Johan Dahlbeck

In this Spinozist defence of the educational promotion of students’ autonomy I argue for a deterministic position where freedom of will is deemed unrealistic in the metaphysical sense, but important in the sense that it is an undeniable psychological fact. The paper is structured in three parts. The first part investigates the concept of autonomy from different philosophical points of view, looking especially at how education and autonomy intersect. The second part focuses on explicating the philosophical position of causal determinism and it seeks to open up a way to conceive of education for autonomy without relying on the notion of free will in a metaphysical sense. The concluding part attempts to outline a Spinozistic understanding of education for autonomy where autonomy is grounded in the student’s acceptance and understanding of the necessary constraints of natural causation rather than processes of self-causation.

RELATED PAPERS

Ethical Theory and Moral Practice

kasper lippert-rasmussen

Philosophical Books

Elinor Mason

Justin Capes

RELATED TOPICS

  •   We're Hiring!
  •   Help Center
  • Find new research papers in:
  • Health Sciences
  • Earth Sciences
  • Cognitive Science
  • Mathematics
  • Computer Science
  • Academia ©2024

University of Notre Dame

Notre Dame Philosophical Reviews

  • Home ›
  • Reviews ›

Merit, Meaning, and Human Bondage: An Essay on Free Will

Placeholder book cover

Nomy Arpaly, Merit, Meaning, and Human Bondage: An Essay on Free Will , Princeton University Press, 2006, 158pp., $29.95 (hbk), ISBN 0691124337.

Reviewed by Angela M. Smith, University of Washington

Nomy Arpaly’s new book represents a nice companion piece to her earlier work, Unprincipled Virtue (Oxford, 2003). In her first book, Arpaly developed a “quality-of-will” based account of moral worth, and argued that agent-autonomy — in the sense of deliberative self-control - is not a necessary condition of moral praise or blameworthiness. Her aim in this first book was to challenge conceptions of moral psychology that give pride of place to the first-person deliberative perspective, and to argue that we need a richer conception of moral agency to accommodate the lived complexity of moral life. In this new book, Arpaly turns her attention to more metaphysical concerns, asking whether the truth of determinism would preclude moral responsibility, reasons-responsiveness, or the ability to lead meaningful lives. Though she answers “no” to these questions — we can still live responsible, rational, meaningful lives in a deterministic setting — she concedes that her own view is still “bittersweet” (p. 5), for there remains a type of freedom that we cannot have if determinism is true. This is the freedom of “absolute self-authorship,” and Arpaly regards the desire for such freedom as entirely understandable, even if it should turn out that such a notion is, ultimately, “incoherent” (p. 127).

In the first chapter, Arpaly provides an overview of the theory of praiseworthiness and blameworthiness she defended in Unprincipled Virtue . On this theory, people are praiseworthy for acts of good will and blameworthy for acts of ill will or lack of good will. Good will, ill will, and lack of good will, in turn, are defined in terms of responsiveness to moral reasons. A person who in fact responds to good moral reasons — whether she herself is aware of this or not — acts out of good will, while a person who in fact fails to respond to moral reasons, or acts for reasons that make his action wrong, displays either moral indifference or ill will. Her view also allows for degrees of praiseworthiness and blameworthiness, depending on the circumstances and how hard or easy it is for the agent in those circumstances to do the right thing. After spelling out her theory of praiseworthiness and blameworthiness, and showing some of its advantages over incompatibilist views that make the exercise of “contra-causal freedom” a precondition of moral assessment, Arpaly briefly discusses the nature of blame and explains why she does not rely on the Strawsonian idea of “reactive attitudes” to ground her account.

In Chapter 2, Arpaly addresses an important objection to her quality-of-will account of moral worth: Can there be such a thing as “reason responsiveness” in a deterministic world? The worry here is that if causal determinism is true, our “responses” to reasons are really no different than a robot moving in response to the inputs of its programmer, or a glass breaking in response to a high, clear note. In response, Arpaly argues persuasively that accepting this view would require us to reject standard materialistic accounts of mental causation in which mental content plays an important role. With the use of many creative and illustrative examples (a hallmark of her work to date), Arpaly distinguishes between what she calls “robot causation,” “content-efficacious causation,” and “reasons responsiveness,” and argues that there is no reason to think that the truth of determinism would leave us stuck at the level of either robot causation or mere content-efficacious causation. Arpaly then goes on to suggest, again quite plausibly, that in everyday life our explanations of people’s behavior usually involve “a mixture of the reason-responsive, the merely-content-responsive, and the robot-like — and woven fine, too” (p. 72).

In Chapter 3, Arpaly argues that the fundamental intuition driving incompatibilism about moral responsibility is the familiar principle that “ought implies can,” and the thought that, in a deterministic universe, there is no sense to be made of the idea that a person can do anything she does not do. Arpaly argues against this principle directly, using the case of responsiveness to epistemic reasons as an example. Most philosophers (and ordinary folk) accept the view that we do not have voluntary control over our beliefs. That is to say, we cannot simply “choose” or “decide” what to believe in the way that we ordinarily can choose or decide to raise our arm. Yet, Arpaly points out, most of us also believe that there are genuine epistemic norms, norms that we sometimes violate: we sometimes believe things we ought not to believe, and draw inferences we should not draw. To say that we “ought not” to believe X or that we “should not” have drawn a certain inference, however, does not imply that we could simply have chosen to believe differently. We can choose to engage in certain activities that will make it more likely that we will come to hold a certain belief, but that is different from saying that we have control over our beliefs themselves. When it comes to beliefs, therefore, it appears “ought” does not imply “can”: “epistemic norms do not come with corresponding abilities to act so as to meet the norms” (p. 101). Arpaly then suggests that a similar claim can be made about moral oughts: we cannot normally choose what we care about, but we “get moral credits for being moved by duty” (p. 108). In this respect, Arpaly claims, “morality and normative epistemology are a lot more similar than it seems” (ibid.).

After engaging in a short interlude to discuss “The Science Fiction of Mind Design,” Arpaly concludes with a chapter discussing the desire many people still have for what she calls “absolute self-authorship” (p. 127). She argues that compatibilists about responsibility, reasons-responsiveness, and meaning in life are often too dismissive of the yearning some people have not only to have their actions be determined by their deepest concerns, but also to have control over what those deepest concerns are. Some of us, she claims, "want a self apart from our mental states that can choose them at will. We want freedom" (p. 126). Though Arpaly herself finds this notion of “absolute self-authorship” incoherent, she argues that, nevertheless, it is still something we can legitimately wish for, and something we can regret not having.

Overall, this is an elegantly written, provocative, and refreshingly different approach to questions of free will and moral responsibility. As with her earlier book, Arpaly does not show much interest in discussing or responding in detail to the work of others, but prefers to approach her topic from a more general intuitive perspective. This is not meant as a criticism: I think it is worthwhile to step back from what has become a very technical (and, in my estimation, often rather sterile) literature on these issues to ask more general questions about what kind of freedom, if any, we really need in order to lead responsible and meaningful lives. Arpaly also does an excellent job of reminding us of the many contexts in which we positively embrace notions of necessity: “It had to be you,” “Here I stand, I can do no other,” “This book had to be written,” etc. These cases, which Arpaly refers to as cases of “romantic necessity,” seem to show that we do not always take determinism to be incompatible with deep forms of love, responsible action, and meaning. And her insightful discussion of the complex relations between reason responsiveness and freedom is a real contribution to the literature.

Nevertheless, I do have some quibbles. One quibble concerns the rather large number of passages and, in some cases, whole sections that have been lifted directly from Unprincipled Virtue . This is not a long book to begin with (137 pages of text); yet large chunks of it have simply been reprinted from her earlier book (in many cases without any explicit acknowledgment of this fact). 1 This is particularly unfortunate, because Arpaly is such a creative and imaginative writer that it would have been nice to see some new examples and arguments put forward in defense of her views.

More substantively, I would have liked to have seen a more detailed discussion of Strawson’s view and how her own theory departs from it, as well as a more careful sorting out of the difference between judgments of blameworthiness, on the one hand, and attitudes of blame, on the other. Regarding the latter issue, at certain points in the first chapter Arpaly is careful to distinguish these notions, suggesting that to judge a person blameworthy is to judge that she has performed a wrong action out of ill will (or lack of good will), while blaming involves both recognizing that a blameworthy action has been performed and disapproving of such actions and the person who performs them (pp. 25-28). In her discussion of constitutive luck, however, she seems to muddy this distinction:

To hold someone blameworthy is not, in itself, to hold that any course of action is appropriate in regard to him, but to hold that a certain attitude toward him is epistemically rational: there was ill will, there was a wrong act, thus blame is warranted. In this way, on my view, blame is analogous to holding someone to be a bad businessman or a lousy artist. (p. 35)

But it seems clear that we can judge someone to be a bad businessman, a lousy artist, or (to use another of her examples) “stupid,” without thinking that we are warranted in feeling disapproval toward him for these traits. If, as she suggests earlier, blame goes beyond judgments of blameworthiness insofar as it involves attitudes of disapproval, then the analogy with assessments of business acumen, artistic competence, and intelligence seems inapt. While I quite agree with Arpaly that it is a mistake to see blame as analogous to punishment, I think it is also a mistake to see blame as analogous to mere fault-finding assessment. What makes blame so tricky, in my view, is precisely that it falls in between these two categories: to blame is not to punish, nor is it merely to negatively assess; it essentially involves the making of a demand, a demand for reasonable regard. If that is correct, then it seems the question about whether blame is “warranted” cannot simply be reduced to the question of whether a wrong act has been performed from ill will. We must also ask whether the demand for reasonable regard implicit in blame is itself justified in the circumstances. Those who view determinism as a threat to responsibility, presumably, think that in a deterministic setting a demand for reasonable regard is not justified, because the agents to whom it is addressed do not have the ability to meet that demand. While Arpaly indirectly addresses this worry in Chapter 3 (when she discusses the failure of “ought implies can” in the epistemic context), I think much more needs be said about this issue as it applies in the moral context.

This brings me to my most serious concern about the book, which is that Arpaly devotes very little time to explaining why her rejection of “ought implies can” in the epistemic context should be thought to have implications for its validity in the moral context. Though she does have a three-page section in which she asks "is there something special about the moral ought?" (p. 106), she does not seem to take this question seriously, and certainly does not discuss it in any detail. I myself am very sympathetic to the claim that there are close analogies between the epistemic case and moral case, but there seems, prima facie , to be a difference between the rational criticism of beliefs and the moral criticism of actions. It may turn out, at the end of the day, that this apparent difference is illusory (moral criticism may be just one species of rational criticism). But I think Arpaly’s argument would have been strengthened considerably if she had addressed this concern more fully.

Nevertheless, there is much to appreciate in this book. Arpaly’s careful analysis of reason responsiveness (and how it differs from other forms of mental causation) is rich and insightful, and the suggestion that moral philosophers should pay more attention to the way normative demands work in the epistemic context is highly suggestive. Her writing is engaging and blessedly unencumbered by tedious technical jargon. I recommend this book to anyone who is looking for a fresh perspective on this age-old philosophical problem.

1 This is perhaps somewhat more understandable in the first chapter, where Arpaly’s aim is to give a summary of the view of moral worth she defends in her earlier book. But passages from the earlier book appear in all but the final chapter of the new book, and the “Interlude” is taken almost in its entirety from that work.

  Take 10% OFF— Expires in h m s Use code save10u during checkout.

Chat with us

  • Live Chat Talk to a specialist
  • Self-service options
  • Search FAQs Fast answers, no waiting
  • Ultius 101 New client? Click here
  • Messenger  

International support numbers

Ultius

For reference only, subject to Terms and Fair Use policies.

  • How it Works

Learn more about us

  • Future writers
  • Explore further

Ultius Blog

Sample essay on free will and moral responsibility.

Ultius

Select network

Free will is a fundamental aspect of modern philosophy. This sample philosophy paper explores how moral responsibility and free will represent an important area of moral debate between philosophers. This type of writing would of course be seen in a philosophy course, but many people might also be inclined to write an essay about their opinions on free will for personal reasons.

History of free will and moral responsibility

In our history, free will and moral responsibility have been longstanding debates amongst philosophers. Some contend that free will does not exist while others believe we have control over our actions and decisions. For the most part, determinists believe that free will does not exist because our fate is predetermined. An example of this philosophy is found in the Book of Genisis .

The biblical story states God created man for a purpose and designed them to worship him. Since God designed humans to operate in a certain fashion and he knew the outcome, it could be argued from a determinist point of view that free will didn't exist. Because our actions are determined, it seems that we are unable to bear any responsibility for our acts.

Galen Strawson has suggested that “in order to be truly deserving, we must be responsible for that which makes us deserving.”

However, Strawson also has implied that we are unable to be responsible. We are unable to be responsible because, as determinists suggest, all our decisions are premade; therefore, we do not act of our own free will. Consequently, because our actions are not the cause of our free will, we cannot be truly deserving because we lack responsibility for what we do.

Defining free will

Free will implies we are able to choose the majority of our actions ("Free will," 2013). While we would expect to choose the right course of action, we often make bad decisions. This reflects the thinking that we do not have free will because if we were genuinely and consistently capable of benevolence, we would freely decide to make the ‘right’ decisions.

In order for free will to be tangible, an individual would have to have control over his or her actions regardless of any external factors. Analyzing the human brain's development over a lifetime proves people have the potential for cognitive reasoning and to make their own decisions.

Casado has argued “the inevitability of free will is such that if one considers freedom an illusion, the internal perspective – and one’s own everyday life – would be totally contradictory” ( 2011, p. 369).

On the other hand, while we can determine whether or not we will wake up the next day, it is not an aspect of our free will because we cannot control this. Incidentally, determinism suggests everything happens exactly the way it should have happened because it is a universal law ("Determinism," 2013). In this way, our free will is merely an illusion.

Have a philosophy assignment? Think about buying an essay from Ultius .

The determinism viewpoint

For example, if we decided the previous night that we would wake up at noon, we are unable to control this even with an alarm clock. One, we may die in our sleep. Obviously, as most would agree, we did not choose this. Perhaps we were murdered in our sleep. In that case, was it our destiny to become a victim of violent crimes, or was it our destiny to be murdered as we slept? Others would mention that the murderer was the sole cause of the violence and it their free will to decide to kill.

Therefore, the same people might argue that the murderer deserved a specific punishment. The key question, then, is the free will of the murderer. If we were preordained to die in the middle of the night at the hand of the murderer, then the choice of death never actually existed. Hence, the very question of choice based on free will is an illusion.

Considering that our wills are absolutely subject to the environment in which they are articulated in, we are not obligated to take responsibility for them as the product of their environment. For example, if we were born in the United States, our actions are the result of our country’s laws. Our constitutional laws allow us the right to bear arms and have access to legal representation. In addition, our constitutional laws allow us the freedom to express our thoughts through spoken and written mediums and the freedom to believe in a higher power or not. We often believe we are free to act and do what we want because of our free will.

Harris (2012) has agreed that “free will is more than an illusion (or less), in that it cannot even be rendered coherent” conceptually.

Moral judgments, decisions, and responsibility for free will

Either our wills are determined by prior causes, and we are not responsible for them, or they are a product of chance, and we are not responsible for them” (p. 46). This being the case, can we be deserving if we can so easily deflect the root of our will and actions? Perhaps, our hypothetical murder shot us. It could be argued that gun laws in the United States provided them with the mean to commit murder.

Either the murderer got a hold of a gun by chance or he or she was able to purchase one. While the purchase is not likely, one would have to assume that someone, maybe earlier, purchased the weapon. Therefore, it was actually the buyer’s action that allowed this particular crime to take place. Essentially, both would ‘deserve’ some sort of punishment.

According to The American Heritage Dictionary (2001), the word “deserving” means "Worthy, as of reward or praise” (p. 236), so it regards to punishments, it seems deserving has a positive meaning.

Free will and changing societal views

However, the meanings will change depending on our position. For example, some would suggest that the murderer acted with his or her own free will. However, once they are caught and convicted, they are no longer free in the sense that they can go wherever they want. On the other hand, they are free to think however they want.

If they choose to reenact their crimes in their thoughts, they are free to do so. Some many say, in the case of the murderer, he or she is held responsible for his or her crime, thus he or she deserves blame. However, if the murderer had a mental illness and was unaware he or she committed a crime, should we still consider that the murderer acted with his or her free will? With that in mind, it seems that Strawson’s argument is valid because the murderer was not acting of his or her free will.

Many would consider Strawson to be a “free will pessimist” (Timpe c. Compatibilism, Incompatibilism, and Pessimism, 2006, para. 5). Strawson does not believe we have the ability to act on our own free will. However, he does not believe our actions are predetermined either.

Specifically, in his article “Luck Swallows Everything,” Strawson (1998) has claimed that “One cannot be ultimately responsible for one's character or mental nature in any way at all” (para. 33).

Determining when free will is not applicable

While some would agree young children and disabled adults would not hold any responsibility, others would claim that criminals should bear responsibility when they commit a crime. What if the actions are caused by both nature and nurturing of the parents ? Or, what if they're caused by prior events including a chain of events that goes back before we are born, libertarians do not see how we can feel responsible for them. If our actions are directly caused by chance, they are simply random and determinists do not see how we can feel responsible for them (The Information Philosopher Responsibility n.d.).

After all, one would not argue that murderers are worthy of a positive reward; however, Strawson has argued that we, whether good or evil, do not deserve any types of rewards. Instead, our actions and their consequences are based on luck or bad luck. In order to have ultimate moral responsibility for an action, the act must originate from something that is separate from us.

We consider free will the ability to act or do as we want; however, there is a difference between freedom of action and freedom of will. Freedom of action suggests we are able to physically act upon our desire. In a way, some believe that freedom of will is the choice that precedes that action. In addition to freedom of act or will, free will also suggests we have a sense of moral responsibility. This moral responsibility, however, is not entirely specified. For example, is this responsibility to ourselves or those around us? While this is a question that may never be answered, no matter how many essays are written on the subject, it is one that many consider important to ask, nonetheless.

https://www.ultius.com/ultius-blog/entry/sample-essay-on-free-will-and-moral-responsibility.html

  • Chicago Style

Ultius, Inc. "Sample Essay on Free Will and Moral Responsibility." Ultius | Custom Writing and Editing Services. Ultius Blog, 18 May. 2014. https://www.ultius.com/ultius-blog/entry/sample-essay-on-free-will-and-moral-responsibility.html

Copied to clipboard

Click here for more help with MLA citations.

Ultius, Inc. (2014, May 18). Sample Essay on Free Will and Moral Responsibility. Retrieved from Ultius | Custom Writing and Editing Services, https://www.ultius.com/ultius-blog/entry/sample-essay-on-free-will-and-moral-responsibility.html

Click here for more help with APA citations.

Ultius, Inc. "Sample Essay on Free Will and Moral Responsibility." Ultius | Custom Writing and Editing Services. May 18, 2014 https://www.ultius.com/ultius-blog/entry/sample-essay-on-free-will-and-moral-responsibility.html.

Click here for more help with CMS citations.

Click here for more help with Turabian citations.

Ultius

Ultius is the trusted provider of content solutions and matches customers with highly qualified writers for sample writing, academic editing, and business writing. 

McAfee Secured

Tested Daily

Click to Verify

About The Author

This post was written by Ultius.

Ultius - Writing & Editing Help

  • Writer Options
  • Custom Writing
  • Business Documents
  • Support Desk
  • +1-800-405-2972
  • Submit bug report
  • A+ BBB Rating!

Ultius is the trusted provider of content solutions for consumers around the world. Connect with great American writers and get 24/7 support.

Download Ultius for Android on the Google Play Store

© 2024 Ultius, Inc.

  • Refund & Cancellation Policy

Free Money For College!

Yeah. You read that right —We're giving away free scholarship money! Our next drawing will be held soon.

Our next winner will receive over $500 in funds. Funds can be used for tuition, books, housing, and/or other school expenses. Apply today for your chance to win!

* We will never share your email with third party advertisers or send you spam.

** By providing my email address, I am consenting to reasonable communications from Ultius regarding the promotion.

Past winner

Past Scholarship Winner - Shannon M.

  • Name Samantha M.
  • From Pepperdine University '22
  • Studies Psychology
  • Won $2,000.00
  • Award SEED Scholarship
  • Awarded Sep. 5, 2018

Thanks for filling that out.

Check your inbox for an email about the scholarship and how to apply.

  • Skip to main content
  • Keyboard shortcuts for audio player

NPR suspends veteran editor as it grapples with his public criticism

David Folkenflik 2018 square

David Folkenflik

an essay on free will

NPR suspended senior editor Uri Berliner for five days without pay after he wrote an essay accusing the network of losing the public's trust and appeared on a podcast to explain his argument. Uri Berliner hide caption

NPR suspended senior editor Uri Berliner for five days without pay after he wrote an essay accusing the network of losing the public's trust and appeared on a podcast to explain his argument.

NPR has formally punished Uri Berliner, the senior editor who publicly argued a week ago that the network had "lost America's trust" by approaching news stories with a rigidly progressive mindset.

Berliner's five-day suspension without pay, which began last Friday, has not been previously reported.

Yet the public radio network is grappling in other ways with the fallout from Berliner's essay for the online news site The Free Press . It angered many of his colleagues, led NPR leaders to announce monthly internal reviews of the network's coverage, and gave fresh ammunition to conservative and partisan Republican critics of NPR, including former President Donald Trump.

Conservative activist Christopher Rufo is among those now targeting NPR's new chief executive, Katherine Maher, for messages she posted to social media years before joining the network. Among others, those posts include a 2020 tweet that called Trump racist and another that appeared to minimize rioting during social justice protests that year. Maher took the job at NPR last month — her first at a news organization .

In a statement Monday about the messages she had posted, Maher praised the integrity of NPR's journalists and underscored the independence of their reporting.

"In America everyone is entitled to free speech as a private citizen," she said. "What matters is NPR's work and my commitment as its CEO: public service, editorial independence, and the mission to serve all of the American public. NPR is independent, beholden to no party, and without commercial interests."

The network noted that "the CEO is not involved in editorial decisions."

In an interview with me later on Monday, Berliner said the social media posts demonstrated Maher was all but incapable of being the person best poised to direct the organization.

"We're looking for a leader right now who's going to be unifying and bring more people into the tent and have a broader perspective on, sort of, what America is all about," Berliner said. "And this seems to be the opposite of that."

an essay on free will

Conservative critics of NPR are now targeting its new chief executive, Katherine Maher, for messages she posted to social media years before joining the public radio network last month. Stephen Voss/Stephen Voss hide caption

Conservative critics of NPR are now targeting its new chief executive, Katherine Maher, for messages she posted to social media years before joining the public radio network last month.

He said that he tried repeatedly to make his concerns over NPR's coverage known to news leaders and to Maher's predecessor as chief executive before publishing his essay.

Berliner has singled out coverage of several issues dominating the 2020s for criticism, including trans rights, the Israel-Hamas war and COVID. Berliner says he sees the same problems at other news organizations, but argues NPR, as a mission-driven institution, has a greater obligation to fairness.

"I love NPR and feel it's a national trust," Berliner says. "We have great journalists here. If they shed their opinions and did the great journalism they're capable of, this would be a much more interesting and fulfilling organization for our listeners."

A "final warning"

The circumstances surrounding the interview were singular.

Berliner provided me with a copy of the formal rebuke to review. NPR did not confirm or comment upon his suspension for this article.

In presenting Berliner's suspension Thursday afternoon, the organization told the editor he had failed to secure its approval for outside work for other news outlets, as is required of NPR journalists. It called the letter a "final warning," saying Berliner would be fired if he violated NPR's policy again. Berliner is a dues-paying member of NPR's newsroom union but says he is not appealing the punishment.

The Free Press is a site that has become a haven for journalists who believe that mainstream media outlets have become too liberal. In addition to his essay, Berliner appeared in an episode of its podcast Honestly with Bari Weiss.

A few hours after the essay appeared online, NPR chief business editor Pallavi Gogoi reminded Berliner of the requirement that he secure approval before appearing in outside press, according to a copy of the note provided by Berliner.

In its formal rebuke, NPR did not cite Berliner's appearance on Chris Cuomo's NewsNation program last Tuesday night, for which NPR gave him the green light. (NPR's chief communications officer told Berliner to focus on his own experience and not share proprietary information.) The NPR letter also did not cite his remarks to The New York Times , which ran its article mid-afternoon Thursday, shortly before the reprimand was sent. Berliner says he did not seek approval before talking with the Times .

NPR defends its journalism after senior editor says it has lost the public's trust

NPR defends its journalism after senior editor says it has lost the public's trust

Berliner says he did not get permission from NPR to speak with me for this story but that he was not worried about the consequences: "Talking to an NPR journalist and being fired for that would be extraordinary, I think."

Berliner is a member of NPR's business desk, as am I, and he has helped to edit many of my stories. He had no involvement in the preparation of this article and did not see it before it was posted publicly.

In rebuking Berliner, NPR said he had also publicly released proprietary information about audience demographics, which it considers confidential. He said those figures "were essentially marketing material. If they had been really good, they probably would have distributed them and sent them out to the world."

Feelings of anger and betrayal inside the newsroom

His essay and subsequent public remarks stirred deep anger and dismay within NPR. Colleagues contend Berliner cherry-picked examples to fit his arguments and challenge the accuracy of his accounts. They also note he did not seek comment from the journalists involved in the work he cited.

Morning Edition host Michel Martin told me some colleagues at the network share Berliner's concerns that coverage is frequently presented through an ideological or idealistic prism that can alienate listeners.

"The way to address that is through training and mentorship," says Martin, herself a veteran of nearly two decades at the network who has also reported for The Wall Street Journal and ABC News. "It's not by blowing the place up, by trashing your colleagues, in full view of people who don't really care about it anyway."

Several NPR journalists told me they are no longer willing to work with Berliner as they no longer have confidence that he will keep private their internal musings about stories as they work through coverage.

"Newsrooms run on trust," NPR political correspondent Danielle Kurtzleben tweeted last week, without mentioning Berliner by name. "If you violate everyone's trust by going to another outlet and sh--ing on your colleagues (while doing a bad job journalistically, for that matter), I don't know how you do your job now."

Berliner rejected that critique, saying nothing in his essay or subsequent remarks betrayed private observations or arguments about coverage.

Other newsrooms are also grappling with questions over news judgment and confidentiality. On Monday, New York Times Executive Editor Joseph Kahn announced to his staff that the newspaper's inquiry into who leaked internal dissent over a planned episode of its podcast The Daily to another news outlet proved inconclusive. The episode was to focus on a December report on the use of sexual assault as part of the Hamas attack on Israel in October. Audio staffers aired doubts over how well the reporting stood up to scrutiny.

"We work together with trust and collegiality everyday on everything we produce, and I have every expectation that this incident will prove to be a singular exception to an important rule," Kahn wrote to Times staffers.

At NPR, some of Berliner's colleagues have weighed in online against his claim that the network has focused on diversifying its workforce without a concomitant commitment to diversity of viewpoint. Recently retired Chief Executive John Lansing has referred to this pursuit of diversity within NPR's workforce as its " North Star ," a moral imperative and chief business strategy.

In his essay, Berliner tagged the strategy as a failure, citing the drop in NPR's broadcast audiences and its struggle to attract more Black and Latino listeners in particular.

"During most of my tenure here, an open-minded, curious culture prevailed. We were nerdy, but not knee-jerk, activist, or scolding," Berliner writes. "In recent years, however, that has changed."

Berliner writes, "For NPR, which purports to consider all things, it's devastating both for its journalism and its business model."

NPR investigative reporter Chiara Eisner wrote in a comment for this story: "Minorities do not all think the same and do not report the same. Good reporters and editors should know that by now. It's embarrassing to me as a reporter at NPR that a senior editor here missed that point in 2024."

Some colleagues drafted a letter to Maher and NPR's chief news executive, Edith Chapin, seeking greater clarity on NPR's standards for its coverage and the behavior of its journalists — clearly pointed at Berliner.

A plan for "healthy discussion"

On Friday, CEO Maher stood up for the network's mission and the journalism, taking issue with Berliner's critique, though never mentioning him by name. Among her chief issues, she said Berliner's essay offered "a criticism of our people on the basis of who we are."

Berliner took great exception to that, saying she had denigrated him. He said that he supported diversifying NPR's workforce to look more like the U.S. population at large. She did not address that in a subsequent private exchange he shared with me for this story. (An NPR spokesperson declined further comment.)

Late Monday afternoon, Chapin announced to the newsroom that Executive Editor Eva Rodriguez would lead monthly meetings to review coverage.

"Among the questions we'll ask of ourselves each month: Did we capture the diversity of this country — racial, ethnic, religious, economic, political geographic, etc — in all of its complexity and in a way that helped listeners and readers recognize themselves and their communities?" Chapin wrote in the memo. "Did we offer coverage that helped them understand — even if just a bit better — those neighbors with whom they share little in common?"

Berliner said he welcomed the announcement but would withhold judgment until those meetings played out.

In a text for this story, Chapin said such sessions had been discussed since Lansing unified the news and programming divisions under her acting leadership last year.

"Now seemed [the] time to deliver if we were going to do it," Chapin said. "Healthy discussion is something we need more of."

Disclosure: This story was reported and written by NPR Media Correspondent David Folkenflik and edited by Deputy Business Editor Emily Kopp and Managing Editor Gerry Holmes. Under NPR's protocol for reporting on itself, no NPR corporate official or news executive reviewed this story before it was posted publicly.

  • Katherine Maher
  • uri berliner

NPR suspends journalist who publicly accused network of liberal bias

Uri Berliner attends a 2017 event.

  • Show more sharing options
  • Copy Link URL Copied!

NPR has suspended a veteran editor who wrote an essay criticizing the public broadcaster for having what he described as a lack of politically diverse viewpoints.

Uri Berliner, an award-winning business journalist who has worked at the network for 25 years, will be off the job for five days without pay. Berliner acknowledged the suspension Monday in an interview with National Public Radio. He did not respond to The Times’ request for comment.

The suspension came after Berliner put a harsh spotlight on NPR with an April 9 opinion piece for the Substack newsletter the Free Press . He said the decline in NPR’s audience levels is due to a move toward liberal political advocacy and catering to “a distilled worldview of a very small segment of the U.S. population.” The overall thrust of the piece asserted that NPR has “lost America’s trust.”

An NPR representative said the network “does not comment on individual personnel matters, including discipline. We expect all of our employees to comply with NPR policies and procedures, which for our editorial staff includes the NPR Ethics Handbook.”

Berliner was told by management last week that he violated company policy by failing to secure its approval to supply work for other news outlets, according to an NPR news report by media correspondent David Folkenflik. Berliner was informed that he will be fired if he violates that policy again.

Berliner’s essay has been seized on by right-wing media outlets that frequently accuse NPR and other mainstream news sources of a liberal bias.

On Monday, conservative activists resurfaced years old social media posts by current NPR Chief Executive Katherine Maher, in which she expressed her disdain for former President Trump. In one 2020 post, she called Trump a racist.

Maher took on her NPR role in January. She previously headed the nonprofit Wikimedia Foundation, which operates Wikipedia, and has no previous experience in journalism. NPR has said Maher was not in an editorial role at the foundation when she made the social media posts, adding that she “is entitled to free speech as a private citizen.”

Berliner’s essay said the network began to lose its way after Trump’s 2016 election victory.

Robert MacNeil and Jim Lehrer were first teamed to cover the Watergate hearings for PBS in 1973.

Company Town

Robert MacNeil, the stately journalist who brought news to PBS, dies at 93

MacNeil is the founding anchor of “PBS NewsHour,” which he launched as “The Robert MacNeil Report.” He also covered the JFK assassination for NBC.

April 12, 2024

“I eagerly voted against Trump twice but felt we were obliged to cover him fairly,” Berliner wrote. “But what began as tough, straightforward coverage of a belligerent, truth-impaired president veered toward efforts to damage or topple Trump’s presidency.”

Berliner said the network overplayed the investigation of Russian collusion with the Trump campaign in the 2016 presidential election. He also said the news operation turned a blind eye to the story of the laptop abandoned by President Biden’s son Hunter in October 2020, out of concern that coverage of the matter would help reelect Trump.

Berliner was also critical of NPR’s coverage of the Israel-Hamas war and the origins of the COVID-19 virus, as well as the organization’s focus on race and identity, which he said “became paramount in nearly every aspect of the workplace.”

Edith Chapin, NPR’s chief news executive, rejected Berliner’s analysis in a memo to staff after his piece was published.

“We’re proud to stand behind the exceptional work that our desks and shows do to cover a wide range of challenging stories,” she wrote. “We believe that inclusion — among our staff, with our sourcing, and in our overall coverage — is critical to telling the nuanced stories of this country and our world.”

More to Read

FILE - Republican National Committee chair Ronna McDaniel speaks before a Republican presidential primary debate hosted by NBC News, Nov. 8, 2023, at the Adrienne Arsht Center for the Performing Arts of Miami-Dade County in Miami. Former NBC News “Meet the Press” moderator Chuck Todd criticized his network Sunday, March 24, 2024, for hiring former Republican National Committee head McDaniel as a paid contributor, saying on the air that many NBC journalists are uncomfortable with the decision. (AP Photo/Rebecca Blackwell, File)

Letters to the Editor: NBC’s firing of Ronna McDaniel was no surprise. But it was a disaster

April 2, 2024

Re-elected Republican National Committee Chair Ronna McDaniel holds a gavel while speaking at the committee's winter meeting in Dana Point, Calif., Friday, Jan. 27, 2023. (AP Photo/Jae C. Hong)

NBC News cuts ties with Ronna McDaniel after internal backlash

March 26, 2024

FILE - Republican National Committee Chair Ronna McDaniel speaks before a Republican presidential primary debate hosted by NBC News, Nov. 8, 2023, at the Adrienne Arsht Center for the Performing Arts of Miami-Dade County in Miami. Facing a cash crunch and harsh criticism from a faction of far-right conservatives, McDaniel, on Friday, Feb. 2, 2024, called for the party to unite behind the goal of defeating President Joe Biden. (AP Photo/Rebecca Blackwell, File)

NBC News in revolt over Ronna McDaniel hiring. Will the network reverse course?

March 25, 2024

Inside the business of entertainment

The Wide Shot brings you news, analysis and insights on everything from streaming wars to production — and what it all means for the future.

You may occasionally receive promotional content from the Los Angeles Times.

an essay on free will

Stephen Battaglio writes about television and the media business for the Los Angeles Times out of New York. His coverage of the television industry has appeared in TV Guide, the New York Daily News, the New York Times, Fortune, the Hollywood Reporter, Inside.com and Adweek. He is also the author of three books about television, including a biography of pioneer talk show host and producer David Susskind.

More From the Los Angeles Times

FILE - A Google sign hangs over an entrance to the company's new building, Sept. 6, 2023, in New York. On Thursday, Oct. 26, Prabhakar Raghavan, Google's senior vice president for knowledge and information products, testified that the company's success is precarious and said its leadership fears their product could slide into irrelevance with younger internet users. Raghavan testified for the tech giant as it defends itself in the biggest antitrust trial in the last 25 years. (AP Photo/Peter Morgan, File)

News publishers’ alliance calls on feds to investigate Google for limiting California links

April 16, 2024

Screens show different segments for a nightly news segment in the control room at the San Diego based One America News on Sept. 5, 2019. Host, Patick Hussion is shown on the bottom left screen.

Smartmatic settles defamation suit against right-wing network OAN

This photo provided by courtesy of Open Road Films shows, Rachel McAdams, from left, as Sacha Pfeiffer, Mark Ruffalo as Michael Rezendes, Brian d’Arcy James as Matt Carroll, Michael Keaton as Walter "Robby" Robinson and John Slattery as Ben Bradlee Jr., in a scene from the film, "Spotlight." (Kerry Hayes/Open Road Films via AP)

Participant, maker of ‘Green Book’ and ‘An Inconvenient Truth,’ is shutting down

Hannah Gutierrez-Reed, center, sits with her attorney Jason Bowles and paralegal Carmella Sisneros during her sentencing hearing in Santa Fe, New Mexico, on Monday, April 15, 2024. Gutierrez-Reed, the armorer on the set of the Western film "Rust," was sentenced to 18 months in prison for involuntary manslaughter in the death of cinematographer Halyna Hutchins, who was fatally shot by Alec Baldwin in 2021. (Eddie Moore/The Albuquerque Journal via AP, Pool)

‘Rust’ armorer Hannah Gutierrez sentenced to 18 months in prison

April 15, 2024

  • Share full article

Advertisement

Supported by

NPR Suspends Editor Whose Essay Criticized the Broadcaster

Uri Berliner, a senior business editor at NPR, said the public radio network’s liberal bias had tainted its coverage of important stories.

Uri Berliner is looking down and to his right. Behind him, there is a large plant, a mustard-yellow couch and a mirror hanging on a wall that shows the reflection of the rest of the room.

By Benjamin Mullin

NPR has suspended Uri Berliner, the senior business editor who broke ranks and published an essay arguing that the nonprofit radio network had allowed liberal bias to affect its coverage.

Mr. Berliner was suspended by the network for five days, starting Friday, for violating the network’s policy against doing work outside the organization without first getting permission.

Mr. Berliner acknowledged his suspension in an interview with NPR on Monday , providing one of the network’s reporters with a copy of the written rebuke. In presenting the warning, NPR said Mr. Berliner had failed to clear his work for outside outlets, adding that he would be fired if he violated the policy again.

Mr. Berliner’s essay was published last week in The Free Press, a popular Substack publication.

He declined to comment about the suspension. NPR said it did not comment on personnel matters.

The revelation of Mr. Berliner’s punishment is the latest aftershock to rattle NPR since he published his essay. Employees at the public radio network were taken aback by Mr. Berliner’s public condemnation of the broadcaster, and several have said they no longer trust him because of his remarks. Mr. Berliner told The New York Times last week that he did not reach out to the network before publishing his essay.

After Mr. Berliner’s essay was published, NPR’s new chief executive, Katherine Maher, came under renewed scrutiny as conservative activists resurfaced a series of years-old social media posts criticizing former President Donald J. Trump and embracing progressive causes. One of the activists, Christopher Rufo, has pressured media organizations into covering controversies involving influential figures, such as the plagiarism allegations against Claudine Gay, the former Harvard president.

NPR said on Monday that Ms. Maher’s social media posts were written long before she was named chief executive of NPR, and that she was not working in the news industry at the time. NPR also said that while she managed the business side of the nonprofit, she was not involved in its editorial process. Ms. Maher said in a statement that “in America everyone is entitled to free speech as a private citizen.”

Several NPR employees have urged the network’s leaders to more forcefully renounce Mr. Berliner’s claims in his essay. Edith Chapin, NPR’s top editor, said in a statement last week that managers “strongly disagree with Uri’s assessment of the quality of our journalism,” adding that the network was “proud to stand behind” its work.

Some employees have begun to speak out. Tony Cavin, NPR’s managing editor for standards and practices, took issue with many of Mr. Berliner’s claims in an interview with The Times on Tuesday, saying Mr. Berliner’s essay mischaracterized NPR’s coverage of crucial stories.

Mr. Cavin said NPR’s coverage of Covid-19, one of the lines of reporting that Mr. Berliner criticized, was in step with reporting from other mainstream news organizations at the time. The coverage, he said, attributed the origins of the virus to a market in Wuhan, China. He also defended NPR’s coverage of Russian interference in the 2016 U.S. election, another area Mr. Berliner focused on, noting that Robert Mueller, the special counsel investigating the issue, concluded that Russian state actors had made attempts to sway the election.

Mr. Cavin also pointed out that NPR had no way to verify early articles about Hunter Biden’s laptop after the story broke but pursued follow-up stories examining the situation. Mr. Berliner wrote that NPR had “turned a blind eye” to the story about Mr. Biden’s laptop.

“To somehow think that we were driven by politics is both wrong and unfair,” Mr. Cavin said.

Benjamin Mullin reports on the major companies behind news and entertainment. Contact Ben securely on Signal at +1 530-961-3223 or email at [email protected] . More about Benjamin Mullin

NPR suspends senior editor Uri Berliner after essay accusing outlet of liberal bias

Npr suspended senior editor uri berliner a week after he authored an online essay accusing the outlet of allowing liberal bias in its coverage..

an essay on free will

NPR has suspended a senior editor who authored an essay published last week on an online news site in which he argued that the network had "lost America's trust" because of a liberal bias in its coverage, the outlet reported.

Uri Berliner was suspended Friday for five days without pay, NPR reported Tuesday . The revelation came exactly a week after Berliner publicly claimed in an essay for The Free Press, an online news publication, that NPR had allowed a "liberal bent" to influence its coverage, causing the outlet to steadily lose credibility with audiences.

The essay reignited the criticism that many prominent conservatives have long leveled against NPR and prompted newsroom leadership to implement monthly internal reviews of the network's coverage, NPR reported. Berliner's essay also angered many of his colleagues and exposed NPR's new chief executive Katherine Maher to a string of attacks from conservatives over her past social media posts.

In a statement Monday to NPR, Maher refuted Berliner's claims by underscoring NPR's commitment to objective coverage of national issues.

"In America everyone is entitled to free speech as a private citizen," Maher said. "What matters is NPR's work and my commitment as its CEO: public service, editorial independence, and the mission to serve all of the American public. NPR is independent, beholden to no party, and without commercial interests."

Heat exposure law: Florida joins Texas in banning local heat protections for outdoor workers

Berliner rails against NPR's coverage of COVID-19, diversity efforts

Berliner, a senior business editor who has worked at NPR for 25 years, argued in the Free Press essay that “people at every level of NPR have comfortably coalesced around the progressive worldview.”

While he claimed that NPR has always had a "liberal bent" ever since he was hired at the outlet, he wrote that it has since lost its "open-minded spirit," and, hence, "an audience that reflects America."

The Peabody Award-winning journalist highlighted what he viewed as examples of the network's partisan coverage of several major news events, including the origins of COVID-19 and the war in Gaza . Berliner also lambasted NPR's diversity, equity and inclusion (DEI) policies – as reflected both within its newsroom and in its coverage – as making race and identity "paramount in nearly every aspect of the workplace.”

"All this reflected a broader movement in the culture of people clustering together based on ideology or a characteristic of birth," he wrote.

Uri Berliner's essay fuels conservative attacks on NPR

In response to the essay, many prominent conservatives and Republicans, including former President Donald Trump, launched renewed attacks at NPR for what they perceive as partisan coverage.

Conservative activist Christopher Rufo in particular targeted Maher for messages she posted to social media years before joining the network – her  first at a news organization . Among the posts singled out were  a 2020 tweet that called Trump racist .

Trump reiterated on his social media platform, Truth Social, his longstanding argument that NPR’s government funding should be rescinded.

NPR issues formal rebuke to Berliner

Berliner provided an NPR reporter with a copy of the formal rebuke for review in which the organization told the editor he had not been approved to write for other news outlets, as is required of NPR journalists.

NPR also said he publicly released confidential proprietary information about audience demographics, the outlet reported.

Leadership said the letter was a "final warning" for Berliner, who would be fired for future violations of NPR's policies, according to NPR's reporting. Berliner, who is a dues-paying member of NPR's newsroom union, told the NPR reporter that he is not appealing the punishment.

A spokeswoman for NPR said the outlet declined to comment on Berliner's essay or the news of his suspension when reached Tuesday by USA TODAY.

"NPR does not comment on individual personnel matters, including discipline," according to the statement. "We expect all of our employees to comply with NPR policies and procedures, which for our editorial staff includes the NPR Ethics Handbook ."

NPR staffer express dismay; leadership puts coverage reviews in place

According to the NPR article, Berliner's essay also invoked the ire of many of his colleagues and the reporters whose stories he would be responsible for editing.

"Newsrooms run on trust," NPR political correspondent Danielle Kurtzleben said in a post last week on social media site X, though he didn't mention Berliner by name. "If you violate everyone's trust by going to another outlet and [expletive] on your colleagues (while doing a bad job journalistically, for that matter), I don't know how you do your job now."

Amid the fallout, NPR reported that NPR's chief news executive Edith Chapin announced to the newsroom late Monday afternoon that Executive Editor Eva Rodriguez would lead monthly meetings to review coverage.

Berliner expressed no regrets about publishing the essay in an interview with NPR, adding that he tried repeatedly to make his concerns over NPR's coverage known to news leaders.

"I love NPR and feel it's a national trust," Berliner says. "We have great journalists here. If they shed their opinions and did the great journalism they're capable of, this would be a much more interesting and fulfilling organization for our listeners."

Eric Lagatta covers breaking and trending news for USA TODAY. Reach him at [email protected]

IMAGES

  1. Free Will Essay

    an essay on free will

  2. College Essay Format: Simple Steps to Be Followed

    an essay on free will

  3. 50 Free Persuasive Essay Examples (+BEST Topics) ᐅ TemplateLab

    an essay on free will

  4. 024 Free Full Essay Example ~ Thatsnotus

    an essay on free will

  5. Free Will Essay

    an essay on free will

  6. We Have No Free Will

    an essay on free will

VIDEO

  1. Class 9 English ll How To Write a Problem Solution Essay ll Lesson: 5.4.2 ll English Hut

  2. Study with me until I blink 👀

  3. How to use AI for Essay Writing #college #texteroai

  4. Free Will? A Documentary

  5. How to use Ai to write essays (no cheating) #texteroai

  6. How to Write a Professional Email : How to Ask for an Extension on a Paper #texteroai

COMMENTS

  1. Peter van Inwagen, An Essay on Free Will (Oxford 1983)

    2 An objection might be raised with the help of an example. A logically necessary condition for giving an eyewitņess account of events that occurred in 1939 is having been born in 1939 or earlier; the fact that X does not satisfy this condition would seem to entail that X lacks the ability to perform this act. Note however that 'giving an eyewitness account of events that occurred in 1939 ...

  2. van Inwagen: An Essay on Free Will

    III. DELIBERATION According to van Inwagen, a rejection of free will would have at least "two interesting consequences"; one who rejected free will, first, could not consistently deliberate and, second, would be committed to denying the existence of moral responsibility. I will contest him on both scores, in order.

  3. Revised ed. Edition

    "An Essay on Free Will" is the best book ever written on the subject of free will. Van Inwagen presents the best arguments for and against compatibilism, the thesis that free will is compatible with determinism. He concludes that this thesis is false. Subsequently, he argues that we have free will, and thus that determinism is false.

  4. An Essay on Free Will

    An Essay on Free Will. Peter Van Inwagen. Clarendon Press, 1983 - Philosophy - 248 pages. In this stimulating and thought-provoking book, the author defends the thesis that free will is incompatible with determinism. He disputes the view that determinism is necessary for moral responsbility.

  5. Peter Van Inwagen, An Essay on Free Will

    Abstract. "This is an important book, and no one interested in issues which touch on the free will will want to ignore it."--Ethics. In this stimulating and thought-provoking book, the author defends the thesis that free will is incompatible with determinism. He disputes the view that determinism is necessary for moral responsbility.

  6. An Essay on Free Will y First edition

    "An Essay on Free Will" is the best book ever written on the subject of free will. Van Inwagen presents the best arguments for and against compatibilism, the thesis that free will is compatible with determinism. He concludes that this thesis is false. Subsequently, he argues that we have free will, and thus that determinism is false.

  7. Van Inwagen on Free Will

    i Peter Van Inwagen, An Essay on Free Will (Oxford, 1984); page references hereafter are to this work. Van Inwagen first presented similar arguments in: "A Formal Approach to the Free Will Problem", Theoria 40 (1974), and "The Incompatibility of Free Will and Determinism", Philosophical Studies 27 (1975), pp. 185-99. VAN INWAGEN ON FREE WILL 253

  8. An Essay on Free Will

    Books. An Essay on Free Will. Inwagen Van, Peter Van Inwagen. Clarendon Press, 1983 - Philosophy - 248 pages. "This is an important book, and no one interested in issues which touch on the free will will want to ignore it."--Ethics. In this stimulating and thought-provoking book, the author defends the thesis that free will is incompatible with ...

  9. An Essay on Free Will by Peter van Inwagen

    Peter van Inwagen. 3.95. 76 ratings5 reviews. This book is a defence of the thesis that free will and determinism are incompatible, and an exploration of some of the consequences of this thesis. Free will is understood at the power to act otherwise than one in fact does, and determinism is understood as the thesis that the past and the laws of ...

  10. CRITICAL NOTICE that free will and determinism are incompatible, and

    Peter van Inwagen An Essay on Free Will (Oxford 1983) Peter van Inwagen wishes to argue for two currently unpopular theses: that free will and determinism are incompatible, and that human beings have free will. He is up against some formidable opponents. As he remarks on p. 190, he is not happy in enlisting in a war in which the

  11. An Essay on Free Will.

    An Essay on Free Will. Gary Watson, P. Inwagen. Published 1 April 1984. Philosophy. The Philosophical Quarterly. In this stimulating and thought-provoking book, the author defends the thesis that free will is incompatible with determinism. He disputes the view that determinism is necessary for moral responsbility.

  12. PDF van Inwagen, AN ESSAY ON FREE WILL

    An Essay on Free Will, by Peter van Inwagen. The Clarendon Press, 1983. Pp. vi, 248. $29.95. Reviewed by ROBERT AUm, The University of Nebraska, Lincoln. This book is a detailed and rigorous study of the relation between freedom and determinism. It provides a conception of what constitutes freedom, and it carefully characterizes determinism.

  13. Free Will

    "Free Will" is a philosophical term of art for a particular sort of capacity of rational agents to choose a course of action from among various alternatives. Which sort is the free will sort is what all the fuss is about. ... An Essay on Free Will. Oxford: Oxford University Press.----- (1994). "When the Will is Not Free," Philosophical ...

  14. An essay on free will by Peter Van Inwagen

    April 1, 2008. Created by an anonymous user. Imported from Scriblio MARC record . An essay on free will by Peter Van Inwagen, 1983, Clarendon Press, Oxford University Press edition, in English.

  15. Free Will

    The term "free will" has emerged over the past two millennia as the canonical designator for a significant kind of control over one's actions. Questions concerning the nature and existence of this kind of control (e.g., does it require and do we have the freedom to do otherwise or the power of self-determination?), and what its true significance is (is it necessary for moral ...

  16. John Martin Fischer's The Metaphysics of Free Will: An Essay on Control

    Second, Fischer defends the striking thesis that such control is not necessary for moral responsibility. This review essay examines Fischer's arguments for each thesis. Fischer's defense of the incompatibilist thesis is the most innovative to date, and I argue that his formulation restructures the free will debate.

  17. Peter van Inwagen, An Essay on Free Will

    An Essay on Free Will. Peter Van Inwagen - 1983 - New York: Oxford University Press. Peter van Inwagen, An Essay on Free Will, Oxford: Clarendon Press, 1983. Stanisław Wnęk - 1984 - Dialectics and Humanism 11 (4):700-701. Free will, chance, and mystery. Laura Ekstrom - 2003 - Philosophical Studies 113 (2):153-80. Van Inwagen on free will.

  18. Thinking about Free Will

    Peter van Inwagen, Thinking about Free Will, Cambridge University Press, 2017, 232pp., $99.99 (hbk), ISBN 9781316617656. No one writes more sensibly about the traditional philosophical problem of free will than does Peter van Inwagen. This book, a collection of his essays on free will, ought to join his An Essay on Free Will, the best modern ...

  19. Existence of Free Will

    Here is the argument: A person acts of her own free will only if she is the act's ultimate source. If determinism is true, no one is the ultimate source of her actions. Therefore, if determinism is true, no one acts of her own free will. (McKenna and Pereboom 2016, 148) 8. This argument requires some unpacking.

  20. Free Will in an Indeterministic World?

    In his An Essay on Free Will Peter van Inwagen pursues two main strategies in order to argue for the conceivability of free will in an indeterministic world: discussing the Mind-argument and rejecting a so called 'scientistic' critique of free will. By analyzing the decision-making process, we pointedly formulate the problems which are fundamental to both of these strategies against the ...

  21. (PDF) An Essay on Free will and Determinism

    Frankfurt defines the idea of will as the first-order desire, effectively causing one to do what one desires. Therefore, to understand the person's will, is to know the freedom of action, which he expresses as " [the] freedom to do what one wants to do". The freedom of action is therefore correlative to having a free will.

  22. Fate, Time, and Language: An Essay on Free Will

    Quite apart from the inclusion of Wallace's essay, the collection of essays in response to Taylor's article could stand alone as a useful (if short) anthology. The addition of Wallace's essay, together with the various bits of reflection on his life as a student and writer, make it both intellectually rich and psychologically illuminating.

  23. Merit, Meaning, and Human Bondage: An Essay on Free Will

    Nomy Arpaly's new book represents a nice companion piece to her earlier work, Unprincipled Virtue (Oxford, 2003). In her first book, Arpaly developed a "quality-of-will" based account of moral worth, and argued that agent-autonomy — in the sense of deliberative self-control - is not a necessary condition of moral praise or blameworthiness. Her aim in this first book was to challenge ...

  24. Sample Essay on Free Will and Moral Responsibility

    Free will is a fundamental aspect of modern philosophy. This sample philosophy paper explores how moral responsibility and free will represent an important area of moral debate between philosophers. This type of writing would of course be seen in a philosophy course, but many people might also be inclined to write an essay about their opinions on free will for personal reasons.

  25. NPR Editor Uri Berliner suspended after essay criticizing network : NPR

    The Free Press is a site that has become a haven for journalists who believe that mainstream media outlets have become too liberal. In addition to his essay, Berliner appeared in an episode of its ...

  26. NPR suspends journalist who publicly accused network of liberal bias

    April 16, 2024 1:32 PM PT. NPR has suspended a veteran editor who wrote an essay criticizing the public broadcaster for having what he described as a lack of politically diverse viewpoints. Uri ...

  27. NPR Suspends Editor Whose Essay Criticized the Broadcaster

    April 16, 2024, 12:09 p.m. ET. NPR has suspended Uri Berliner, the senior business editor who broke ranks and published an essay arguing that the nonprofit radio network had allowed liberal bias ...

  28. NPR suspends editor Uri Berliner over essay accusing outlet of bias

    0:03. 1:47. NPR has suspended a senior editor who authored an essay published last week on an online news site in which he argued that the network had "lost America's trust" because of a liberal ...