Free Will, Determinism, and Moral Responsibility

--> Arthurs, Frank (2014) Free Will, Determinism, and Moral Responsibility. PhD thesis, University of Sheffield.

The first half of this thesis is a survey of the PSR, followed by consideration of arguments for and against the principle. This survey spans from the Ancient Greeks to the present day, and gives the reader a sense of the ways in which the PSR has been used both implicitly and explicitly throughout the history of philosophy. I argue that, while none of the arguments either for or against the PSR provide conclusive evidence of its truth or falsity, we should adopt a presumption in its favour. The best hope the PSR sceptic has of demonstrating the PSR’s falsity would be to find empirical evidence of something non-deterministic, since the PSR entails determinism. The theory of libertarianism is considered as just such a counterexample; but I argue the evidence for libertarianism is flimsy, and so the presumption in favour of the PSR remains. The second half starts from the premise that the PSR—and hence also determinism—is true, and goes on to examine what implications this has for our moral responsibility practices. We examine incompatibilist arguments by van Inwagen and Galen Strawson, both of which appeal to the origination condition. I contend that these arguments are compelling precisely because the origination condition to which they appeal is compelling. This leaves us with a dilemma: it seems like we can either accept these incompatibilist arguments, which would require us to abandon our moral responsibility practices; or we could save our moral responsibility practices by adopting some form of compatibilism, but at the cost of denying the intuitively appealing origination condition. In fact, to avoid the costs of each horn of this dilemma, we can seek to create a ‘mixed view’ instead. We consider Vargas’s revisionism, Double’s free will subjectivism, and Smilansky’s illusionism and fundamental dualism, which help to shape the mixed view I argue for here: a consequentialist compatibilist theory of moral responsibility. This theory allows us to acknowledge the impossibility of true desert without dispensing with our responsibility practices.

--> Philosophy PhD on Ethics and Metaphysics -->

Filename: PhD - Free Will, Determinism, and Moral Responsibility. Complete pdf File..pdf

Description: Philosophy PhD on Ethics and Metaphysics

Creative Commons Licence

Embargo Date:

[img]

You do not need to contact us to get a copy of this thesis. Please use the 'Download' link(s) above to get a copy. You can contact us about this thesis . If you need to make a general enquiry, please see the Contact us page.

-

Free Will (Philosophy)

  • Reference work entry
  • First Online: 01 January 2021
  • pp 3237–3240
  • Cite this reference work entry

Book cover

  • Matthew A Sarraf 3  

11 Accesses

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

Access this chapter

Institutional subscriptions

Caruso, G. D. (2013). Free will and consciousness: A determinist account of the illusion of free will . Landham: Lexington Books.

Google Scholar  

Dennett, D. C. (2015). Elbow room: The varieties of free will worth wanting . Cambridge, MA: MIT Press.

Book   Google Scholar  

Horst, S. W. (2011). Laws, mind, and free will . Cambridge, MA: MIT Press.

McKenna, M., & Pereboom, D. (2016). Free will: A contemporary introduction . New York: Routledge.

Mele, A. R. (2006). Free will and luck . Oxford: Oxford University Press.

[Moscow Center for Consciousness Studies]. (2014). Greenland 2014 Dennett on Pereboom . [Video File]. Retrieved from https://www.youtube.com/watch?v=cN7cEDPgMek

Pereboom, D. (2016). Free will, agency, and meaning in life . New York: Oxford University Press.

Sartorio, C. (2016). Causation and free will . Oxford: Oxford University Press.

Strawson, G. (1986). Freedom and belief . Oxford: Oxford University Press.

Download references

Author information

Authors and affiliations.

University of Rochester, Rochester, NY, USA

Matthew A Sarraf

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Matthew A Sarraf .

Editor information

Editors and affiliations.

Department of Psychology, Oakland University, Rochester, MI, USA

Todd K Shackelford

Viviana A Weekes-Shackelford

Section Editor information

Pennsylvania State University, University Park, PA, USA

Karin Machluf

Rights and permissions

Reprints and permissions

Copyright information

© 2021 Springer Nature Switzerland AG

About this entry

Cite this entry.

Sarraf, M.A. (2021). Free Will (Philosophy). In: Shackelford, T.K., Weekes-Shackelford, V.A. (eds) Encyclopedia of Evolutionary Psychological Science. Springer, Cham. https://doi.org/10.1007/978-3-319-19650-3_2167

Download citation

DOI : https://doi.org/10.1007/978-3-319-19650-3_2167

Published : 22 April 2021

Publisher Name : Springer, Cham

Print ISBN : 978-3-319-19649-7

Online ISBN : 978-3-319-19650-3

eBook Packages : Behavioral Science and Psychology Reference Module Humanities and Social Sciences Reference Module Business, Economics and Social Sciences

Share this entry

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

Logo for University of Central Florida Pressbooks

Chapter 5: The problem of free will and determinism

The problem of free will and determinism

Matthew Van Cleave

“You say: I am not free. But I have raised and lowered my arm. Everyone understands that this illogical answer is an irrefutable proof of freedom.”

-Leo Tolstoy

“Man can do what he wills but he cannot will what he wills.”

-Arthur Shopenhauer

“None are more hopelessly enslaved than those who falsely believe they are free.”

-Johann Wolfgang von Goethe

The term “freedom” is used in many contexts, from legal, to moral, to psychological, to social, to political, to theological. The founders of the United States often extolled the virtues of “liberty” and “freedom,” as well as cautioned us about how difficult they were to maintain. But what do these terms mean, exactly? What does it mean to claim that humans are (or are not) free? Almost anyone living in a liberal democracy today would affirm that freedom is a good thing, but they almost certainly do not all agree on what freedom is. With a concept as slippery as that of free will, it is not surprising that there is often disagreement. Thus, it will be important to be very clear on what precisely we are talking about when we are either affirming or denying that humans have free will. There is an important general point here that extends beyond the issue of free will: when debating whether or not x exists, we must first be clear on defining x, otherwise we will end up simply talking past each other . The philosophical problem of free will and determinism is the problem of whether or not free will exists in light of determinism. Thus, it is crucial to be clear in defining what we mean by “free will” and “determinism.” As we will see, these turn out to be difficult and contested philosophical questions. In this chapter we will consider these different positions and some of the arguments for, as well as objections to, them.

Let’s begin with an example. Consider the 1998 movie, The Truman Show . In that movie the main character, Truman Burbank (played by Jim Carrey), is the star of a reality television show. However, he doesn’t know that he is. He believes he is just an ordinary person living in an ordinary neighborhood, but in fact this neighborhood is an elaborate set of a television show in which all of his friends and acquaintances are just actors. His every moment is being filmed and broadcast to a whole world of fans that he doesn’t know exists and almost every detail of his life has been carefully orchestrated and controlled by the producers of the show. For example, Truman’s little town is surrounded by a lake, but since he has been conditioned to believe (falsely) that he had a traumatic boating accident in which his father died, he never has the desire to leave the small little town and venture out into the larger world (at least at first). So consider the life of Truman as described above. Is he free or not? On the one hand, he gets to do pretty much everything he wants to do and he is pretty happy. Truman doesn’t look like he’s being coerced in any explicit way and if you asked him if he was, he would almost certain reply that he wasn’t being coerced and that he was in charge of his life. That is, he would say that he was free (at least to the same extent that the rest of us would). These points all seem to suggest that he is free. For example, when Truman decides that he would rather not take a boat ride out to explore the wider world (which initially is his decision), he is doing what he wants to do. His action isn’t coerced and does not feel coerced to him. In contrast, if someone holds a gun to my head and tells me “your wallet or your life!” then my action of giving him my wallet is definitely coerced and feels so.

On the other hand, it seems clear the Truman’s life is being manipulated and controlled in a way that undermines his agency and thus his freedom. It seems clear that Truman is not the master of his fate in the way that he thinks he is. As Goethe says in the epigraph at the beginning of this chapter, there’s a sense in which people like Truman are those who are most helplessly enslaved, since Truman is subject to a massive illusion that he has no reason to suspect. In contrast, someone who knows she is a slave (such as slaves in the antebellum South in the United States) at least retains the autonomy of knowing that she is being controlled. Truman seems to be in the situation of being enslaved and not knowing it and it seems harder for such a person to escape that reality because they do not have any desire to (since they don’t know they are being manipulated and controlled).

As the Truman Show example illustrates, it seems there can be reasonable disagreement about whether or not Truman is free. On the one hand, there’s a sense in which he is free because he does what he wants and doesn’t feel manipulated. On the other hand, there’s a sense in which he isn’t free because what he wants to do is being manipulated by forces outside of his control (namely, the producers of the show). An even better example of this kind of thing comes from Aldous Huxley’s classic dystopia, Brave New World . In the society that Huxley envisions, everyone does what they want and no one is ever unhappy. So far this sounds utopic rather than dystopic. What makes it dystopic is the fact that this state of affairs is achieved by genetic and behavioral conditioning in a way that seems to remove any choice. The citizens of the Brave New World do what they want, yes, but they seems to have absolutely no control over what they want in the first place. Rather, their desires are essentially implanted in them by a process of conditioning long before they are old enough to understand what is going on. The citizens of Brave New World do what they want, but they have no control over what the want in the first place. In that sense, they are like robots: they only have the desires that are chosen for them by the architects of the society.

So are people free as long as they are doing what they want to—that is, choosing the act according to their strongest desires? If so, then notice that the citizens of Brave New World would count as free, as would Truman from The Truman Show , since these are both cases of individuals who are acting on their strongest desires. The problem is that those desires are not desires those individuals have chosen. It feels like the individuals in those scenarios are being manipulated in a way that we believe we aren’t. Perhaps true freedom requires more than just that one does what one most wants to do. Perhaps true freedom requires a genuine choice. But what is a genuine choice beyond doing what one most wants to do?

Philosophers are generally of two main camps concerning the question of what free will is. Compatibilists believe that free will requires only that we are doing what we want to do in a way that isn’t coerced—in short, free actions are voluntary actions. Incompatibilists , motivated by examples like the above where our desires are themselves manipulated, believe that free will requires a genuine choice and they claim that a choice is genuine if and only if, were we given the choice to make again, we could have chosen otherwise . I can perhaps best crystalize the difference between these two positions by moving to a theological example. Suppose that there is a god who created the universe, including humans, and who controls everything that goes on in the universe, including what humans do. But suppose that god does this not my directly coercing us to do things that we don’t want to do but, rather, by implanting the desire in us to do what god wants us to do. Thus human beings, by doing what the want to do, would actually be doing what god wanted them to do. According to the compatibilist, humans in this scenario would be free since they would be doing what they want to do. According to the incompatibilist, however, humans in this scenario would not be free because given the desire that god had implanted in them, they would always end up doing the same thing if given the decision to make (assuming that desires deterministically cause behaviors). If you don’t like the theological example, consider a sci-fi example which has the same exact structure. Suppose there is an eccentric neuroscientist who has figured out how to wire your brain with a mechanism by which he can implant desire into you.

Suppose that the neuroscientist implants in you the desire to start collecting stamps and you do so. However, you know none of this (the surgery to implant the device was done while you were sleeping and you are none the wiser).

From your perspective, one day you find yourself with the desire to start collecting stamps. It feels to you as though this was something you chose and were not coerced to do. However, the reality is that given this desire that the neuroscientist implanted in you, you could not have chosen not to have started collecting stamps (that is, you were necessitated to start collecting stamps, given the desire). Again, in this scenario the compatibilist would say that your choice to start collecting stamps was free (since it was something you wanted to do and did not feel coerced to you), but the incompatibilist would say that your choice was not free since given the implantation of the desire, you could not have chosen otherwise.

We have not quite yet gotten to the nub of the philosophical problem of free will and determinism because we have not yet talked about determinism and the problem it is supposed to present for free will. What is determinism?

Determinism is the doctrine that every cause is itself the effect of a prior cause. More precisely, if an event (E) is determined, then there are prior conditions (C) which are sufficient for the occurrence of E. That means that if C occurs, then E has to occur. Determinism is simply the claim that every event in the universe is determined. Determinism is assumed in the natural sciences such as physics, chemistry and biology (with the exception of quantum physics for reasons I won’t explain here). Science always assumes that any particular event has some law-like explanation—that is, underlying any particular cause is some set of law- like regularities. We might not know what the laws are, but the whole assumption of the natural sciences is that there are such laws, even if we don’t currently know what they are. It is this assumption that leads scientists to search for causes and patterns in the world, as opposed to just saying that everything is random. Where determinism starts to become contentious is when we move into the human sciences, such as psychology, sociology, and economics. To illustrate why this is contentious, consider the famous example of Laplace’s demon that comes from Pierre-Simon Laplace in 1814:

We may regard the present state of the universe as the effect of its past and the cause of its future. An intellect which at a certain moment would know all forces that set nature in motion, and all positions of all items of which nature is composed, if this intellect were also vast enough to submit these data to analysis, it would embrace in a single formula the movements of the greatest bodies of the universe and those of the tiniest atom; for such an intellect nothing would be uncertain and the future just like the past would be present before its eyes.

Laplace’s point is that if determinism were true, then everything that every happened in the universe, including every human action ever undertaken, had to have happened. Of course humans, being limited in knowledge, could never predict everything that would happen from here out, but some being that was unlimited in intelligence could do exactly that. Pause for a moment to consider what this means. If determinism is true, then Laplace’s demon would have been able to predict from the point of the big bang, that you would be reading these words on this page at this exact point of time. Or that you had what you had for breakfast this morning. Or any other fact in the universe. This seems hard to believe, since it seems like some things that happen in the universe didn’t have to happen. Certain human actions seem to be the paradigm case of such events. If I ate an omelet for breakfast this morning, that may be a fact but it seems strange to think that this fact was necessitated as soon as the big bang occurred. Human actions seem to have a kind of independence from web of deterministic web of causes and effects in a way that, say, billiard balls don’t.

Given that the cue ball his the 8 ball with a specific velocity, at a certain angle, and taking into effect the coefficient of friction of the felt on the pool table, the exact location of the 8 ball is, so to speak, already determined before it ends up there. But human behavior doesn’t seem to be like the behavior of the 8 ball in this way, which is why some people think that the human sciences are importantly different than the natural sciences. Whether or not the human sciences are also deterministic is an issue that helps distinguish the different philosophical positions one can take on free will, as we will see presently. But the important point to see right now is that determinism is a doctrine that applies to all causes, including human actions. Thus, if some particular brain state is what ultimately caused my action and that brain state itself was caused by a prior brain state, and so on, then my action had to occur given those earlier prior events. And that entails that I couldn’t have chosen to act otherwise, given that those earlier events took place . That means that the incompatibilist position on free will cannot be correct if determinism is true. Recall that incompatibilism requires that a choice is free only if one could have chosen differently, given all the same initial conditions. But if determinism is true, then human actions are no different than the 8 ball: given what has come before, the current event had to happen. Thus, if this morning I cooked an omelet, then my “choice” to make that omelet could not have been otherwise. Given the complex web of prior influences on my behavior, my making that omelet was determined. It had to occur.

Of course, it feels to us, when contemplating our own futures, that there are many different possible ways our lives might go—many possible choices to be made. But if determinism is true, then this is an illusion. In reality, there is only one way that things could go, it’s just that we can’t see what that is because of our limited knowledge. Consider the figure below. Each junction in the figure below represents a decision I make and let’s suppose that some (much larger) decision tree like this could represent all of the possible ways my life could go. At any point in time, when contemplating what to do, it seems that I can conceive of my life going many different possible ways. Suppose that A represents one series of choices and B another. Suppose, further, that A represents what I actually do (looking backwards over my life from the future). Although from this point in time it seems that I could also have made the series of choices represented in B, if determinism is true then this is false. That is, if A is what ends up happening, then A is the only thing that ever could have happened . If it hasn’t yet hit you how determinism conflicts with our sense of our own possibilities in life, think about that for a second.

image

As the foregoing I hope makes clear, the incompatibilist definition of free will is incompatibile with determinism (that’s why it’s called “incompatibilist”). But that leaves open the question of which one is true. To say that free will and determinism are logically incompatible is just to say that they cannot both be true, meaning that one or the other must be false. But which one? Some will claim that it is determinism which is false. This position is called libertarianism (not to be confused with political libertarianism, which is a totally different idea). Others claim that determinism is true and that, therefore, there is no free will. This position is called hard determinism. A third type of position, compatibilism , rejects the incompatibilist definition of freedom and claims that free will and determinism are compatible (hence the name). The table below compares these different positions. But which one is correct? In the remainder of the chapter we will consider some arguments for and against these three positions on free will and determinism.

image

Libertarianism

Both libertarianism and hard determinism accept the following proposition: If determinism is true, then there is no free will. What distinguishes libertarianism from hard determinism is the libertarian’s claim that there is free will. But why should we think this? This question is especially pressing when we recognize that we assume a deterministic view in many other domains in life. When you have a toothache, we know that something must have caused that toothache and whatever cause that was, something else must have caused that cause. It would be a strange dentist who told you that your toothache didn’t have a cause but just randomly occurred. When the weather doesn’t go as the meteorologist predicts, we assume there must be a cause for why the weather did what it did. We might not ever know the cause in all its specific details, but assume there must be one. In cases like meteorology, when our scientific predictions are wrong, we don’t always go back and try to figure out what the actual causes were—why our predictions were wrong. But in other cases we do. Consider the explosion of the Space Shuttle Challenger in 1986. Years later it was finally determined what led to that explosion (“O-ring” seals that were not designed for the colder condition of the launch). There’s a detailed deterministic physical explanation that one could give of how the failure of those O-rings led to the explosion of the Challenger. In all of these cases, determinism is the fundamental assumption and it seems almost nothing could overturn it.

But the libertarian thinks that the domain of human action is different than every other domain. Humans are somehow able to rise above all of the influences on them and make decisions that are not themselves determined by anything that precedes them. The philosopher Roderick Chisholm accurately captured the libertarian position when he claimed that “we have a prerogative which some would attribute only to God: each of us, when we really act, is a prime mover unmoved. In doing what we do, we cause certain events to happen, and nothing and no one, except we ourselves, causes us to cause those events to happen” (Chisholm, 1964). But why should we think that we have such a godlike ability? We will consider two arguments the libertarian makes in support of her position: the argument from intuitions and the argument from moral responsibility.

The argument from intuitions is based on the very strong intuition that there are some things that we have control over and that nothing causes us to do those things except for our own willing them. The strongest case for this concerns very simple actions, such as moving one’s finger. Suppose that I hold up my index finger and say that I am going to move it to the right or the to the left, but that I have not yet decided which way to move it. At the moment before I move my finger one way or the other, it truly seems to me that my future is open.

Nothing in my past and nothing in my present seems to be determining me to move my finger to the right or to the left. Rather, it seems to me that I have total control over what happens next. Whichever way I move my finger, it seems that I could have moved it the other way. So if, as a matter of fact, I move my finger to the right, it seems unquestionably true that I could have moved it to the left (and vice versa, mutatis mutandis ). Thus, in cases of simple actions like moving my finger to the right or left, it seems that the strong incompatibilist definition of freedom is met: we have a very strong intuition that no matter what I actually did, I could have chosen otherwise, were I to be given that exact choice again . The libertarian does not claim that all human actions are like this. Indeed, many of our actions (perhaps even ones that we think are free) are determined by prior causes. The libertarian’s claim is just that at least some of our actions do meet the incompatibilist’s definition of free and, thus, that determinism is not universally true.

The argument from moral responsibility is a good example of what philosophers call a transcendental argument . Transcendental arguments attempt to establish the truth of something by showing that that thing is necessary in order for something else, which we strongly believe to be true, to be true. So consider the idea that normally developed adult human beings are morally responsible for their actions. For example, if Bob embezzles money from his charity in order to help pay for a new sports car, we would rightly hold Bob accountable for this action. That is, we would punish Bob and would see punishment as appropriate. But a necessary condition of holding Bob responsible is that Bob’s action was one that he chose, one that he was in control of, one that he could have chosen not to do. Philosophers call this principle ought implies can : if we say that someone ought (or ought not) do something, this implies that they can do it (that is, they are capable of doing it). The ought implies can principle is consistent with our legal practices. For example, in cases where we believe that a person was not capable of doing the right thing, we no longer hold them morally or criminally liable. A good example of this within our legal system is the insanity defense: if someone was determined to be incapable of appreciating the difference between right and wrong, we do not find them guilty of a crime. But notice what determinism would do to the ought implies can principle. If everything we ever do had to happen (think Laplace’s demon), that means that Bob had to embezzle those funds and buy that sports car. The universe demanded it. That means he couldn’t not have done those things. But if that is so, then, by the ought implies can principle, we cannot say that he ought not to have done those things. That is, we cannot hold Bob morally responsible for those things. But this seems absurd, the libertarian will say.

Surely Bob was responsible for those things and we are right to hold him responsible. But the only we way can reasonably do this is if we assume that his actions were chosen—that he could have chosen to do otherwise than he in fact chose. Thus, determinism is incompatible with the idea that human beings are morally responsible agents. The practice of holding each other to be morally responsible agents doesn’t make sense unless humans have incompatibilist free will—unless they could have chosen to do otherwise than they in fact did. That is the libertarian’s transcendental argument from moral responsibility.

Hard determinism

Hard determinism denies that there is free will. The hard determinist is a “tough-minded” individual who bravely accepts the implication of a scientific view of the world. Since we don’t in general accept that there are causes that are not themselves the result of prior causes, we should apply this to human actions too. And this means that humans, contrary to what they might believe (or wish to believe) about themselves, do not have free will. As noted above, hard determininsm follows from accepting the incompatibilist definition of free will as well as the claim that determinism is universally true. One of the strongest arguments in favor of hard determinism is based on the weakness of the libertarian position. In particular, the hard determinist argues that accepting the existence of free will leaves us with an inexplicable mystery: how can a physical system initiate causes that are not themselves caused?

If the libertarian is right, then when an action is free it is such that given exactly the same events leading up to one’s action, one still could have acted otherwise than they did. But this seems to require that the action/choice was not determined by any prior event or set of events. Consider my decision to make a cheese omelet for breakfast this morning. The libertarian will say that my decision to make the cheese omelet was not free unless I could have chosen to do otherwise (given all the same initial conditions). But that means that nothing was determining my decision. But what kind of thing is a decision such that it causes my actions but is not itself caused by anything? We do not know of any other kind of thing like this in the universe. Rather, we think that any event or thing must have been caused by some (typically complex) set of conditions or events. Things don’t just pop into existence without being caused . That is as fundamental a principle as any we can think of. Philosophers have for centuries upheld the principle that “nothing comes from nothing.” They even have a fancy Latin phrase for it: ex nihilo nihil fir [1] . The problem is that my decision to make a cheese omelet seems to be just that: something that causes but is not itself caused. Indeed, as noted earlier, the libertarian Roderick Chisholm embraces this consequence of the libertarian position very clearly when he claimed that when we exercise our free will,

“we have a prerogative which some would attribute only to God: each of us, when we really act, is a prime mover unmoved. In doing what we do, we cause certain events to happen, and nothing and no one, except we ourselves, causes us to cause those events to happen” (Chisholm, 1964).

How could something like this exist? At this point the libertarian might respond something like this:

I am not claiming that something comes from nothing; I am just claiming that our decisions are not themselves determined by any prior thing. Rather, we ourselves, as agents, cause our decisions and nothing else causes us to cause those decisions (at least in cases where we have acted freely).

However, it seems that the libertarian in this case has simply pushed the mystery back one step: we cause our decisions, granted, but what causes us to make those decisions? The libertarian’s answer here is that nothing causes us. But now we have the same problem again: the agent is responsible for causing the decision but nothing causes the agent to make that decision. Thus we seem to have something coming from nothing. Let’s call this argument the argument from mysterious causes. Here’s the argument in standard form:

  • The existence of free will implies that when an agent freely decides to do something, the agent’s choice is not caused (determined) by anything.
  • To say that something has no cause is to violate the ex nihilo nihil fit principle.
  • But nothing can violate the ex nihilo nihil fit principle.
  • Therefore, there is no free will (from 1-3)

The hard determinist will make a strong case for premise 3 in the above argument by invoking basic scientific principles such as the law of conservation of energy, which says that the amount of energy within a closed system stays that same. That is, energy cannot be created or destroyed. Consider a billiard ball. If it is to move then it must get the required energy to do so from someplace else (typically another billiard ball knocking into it, the cue stick hitting it or someone tilting the pool table). To allow that something could occur without any cause—in this case, the agent’s decision—would be a violation of the conservation of energy principle, which is as basic a scientific principle as we know. When forced to choose between uphold such a basic scientific principle as this and believing in free will, the hard determinist opts for the former. The hard determinist will put the ball in the libertarian’s court to explain how something could come from nothing.

I will close this section by indicating how these problems in philosophy often ramify into other areas of philosophy. In the first place, there is a fairly common way that libertarians respond to the charge that their view violates basic principles such as ex nihilo nihil fit and, more specifically, the physical law of conservation of energy. Libertarians could claim that the mind is not physical—a position known in the philosophy of mind as “substance dualism” (see philosophy of mind chapter in this textbook for more on substance dualism). If the mind isn’t physical, then neither are our mental events, such our decisions.

Rather, all of these things are nonphysical entities. If decisions are nonphysical entities, there is at least no violation of the physical laws such as the law of conservation of energy. [2] Of course, if the libertarian were to take this route of defending her position, she would then need to defend this further assumption (no small task). In any case, my main point here is to see the way that responses to the problem of free well and determinism may connect with other issue within philosophy. In this case, the libertarian’s defense of free will may turn out to depend on the defensibility of other assumptions they make about the nature of the mind. But the libertarian is not the only one who will need to ultimately connect her account of free will up with other issues in philosophy. Since hard determinists deny that free will exists, it seems that they will owe us some account of moral responsibility. If, moral responsibility requires that humans have free will (see previous section), then in denying free will we seem to also be denying that humans have moral responsibility. Thus, hard determinists will face the objection that in rejecting free will they also destroy moral responsibility.

But since it seems we must hold individuals morally to account for certain actions (such as the embezzler from the previous section), the hard determinist

needs some account of how it makes sense to do this given that human being don’t have free will. My point here is not to broach the issue of how the hard determinist might answer this, but simply to show how hard determinist’s position on the problem of free will and determinism must ultimately connect with other issues in philosophy, such issues in metaethics [3] . This is a common thing that happens in philosophy. We may try to consider an issue or problem in isolation, but sooner or later that problem will connect up with other issues in philosophy.

Compatibilism

The best argument for compatibilism builds on a consideration of the difficulties with the incompatibilist definition of free will (which both the libertarian and the hard determinist accept). As defined above, compatibilists agree with the hard determinists that determinism is true, but reject the incompatibilist definition of free will that hard determinists accept. This allows compatbilists to claim that free will is compatible with determinism. Both libertarians and hard compatibilists tend to feel that this is somehow cheating, but the compatibilist attempts to convince us arguing that the strong incompatibilist definition of freedom is problematic and that only the weaker compatibilist definition of freedom—free actions are voluntary actions—will work. We will consider two objections that the compatibilist raises for the incompatibilist definition of freedom: the epistemic objection and the arbitrariness objection . Then we will consider the compatibilist’s own definition of free will and show how that definition fits better with some of our common sense intuitions about the nature of free actions.

The epistemic objection is that there is no way for us to ever know whether any one of our actions was free or not. Recall that the incompatibilist definition of freedom says that a decision is free if and only if I could have chosen otherwise than I in fact chose, given exactly all the same conditions. This means that if we were, so to speak, rewind the tape of time and be given that decision to make over again, we could have chosen differently. So suppose the question is whether my decision to make a cheese omelet for breakfast was free. To answer this question, we would have to know when I could have chosen differently. But how am I supposed to know that ? It seems that I would have to answer a question about a strange counterfactual : if given that decision to make over again, would I choose the same way every time or not? How on earth am I supposed to know how to answer that question? I could say that it seems to me that I could make a different decision regarding making the cheese omelet (for example, I could have decided to eat cereal instead), but why should I think that that is the right answer? After all, how things seem regularly turn out to be not the case—especially in science. The problem is that I don’t seem to have any good way of answering this counterfactual question of what I would choose if given the same decision to make over again. Thus the epistemic objection [4] is that since I have no way of knowing whether I would/wouldn’t make the same decision again, I can never know whether any of my actions are free.

The arbitrariness objection is that it turns our free actions into arbitrary actions. And arbitrary actions are not free actions. To see why, consider that if the incompatibilist definition is true, then nothing determines our free choices, not even our own desires . For if our desires were determining our choices then if we were to rewind the tape of time and, so to speak, reset everything— including our desires —the same way, then given those same desires we would choose the same way every time. And that would mean our choice was not free, according to the incompatbilist. It is imperative to remember that incompatibilism says that if an action is not free if it is determined ( including if it is determined by our own desires ). But now the question is: if my desires are not causing my decision, what is? When I make a decision, where does that decision come from, if not from my desires and beliefs? Presumably it cannot come from nothing ( ex nihilo nihil fit ). The problem is that if the incompatibilist rejects that anything is causing my decisions, then how can my decisions be anything but arbitrary?

Presumably an arbitrary decision—a decision not driven by any reason at all—is not an exercise of my freedom. Freedom seems to require that we are exercising some kind of control over my actions and decisions. If my action or decision is arbitrary that means that no reason or explanation of the action/decision can be given. Here’s the arbitrariness objection cast as a reductio ad absurdum argument:

  • A free choice is one that isn’t determined by anything, including our desires. [incompatibilist definition of freedom]
  • If our own desires are not determining our choices, then those choices are arbitrary.
  • If a choice is arbitrary then it is not something over which we have control
  • If a choice isn’t something over which we have control, then it isn’t a free choice
  • Therefore, a free choice is not a free choice (from 1-4)

A reductio ad absurdum argument is one that starts with a certain assumption and then derives a contradiction from that assumption, thereby showing that assumption must be false. In this case, the incompatibilist’s definition of a free choice leads to the contradiction that something that is a free choice isn’t a free choice.

What has gone wrong here? The compatibilist will claim that what has gone wrong is incompatibilist’s idea that a free action must be one that isn’t caused/determined by anything. The compatibilist claims that free actions can still be determined, as long as what is determining them is our own internal reasons, over which we have some control, rather than external things over which we have no control. Free choices, according to the compatibilist, are just choices that are caused by our own best reasons . The fact that, given the exact same choices, I couldn’t have chosen otherwise doesn’t undermine what freedom is (as the incompatibilist claims) but defines what it is. Consider an example. Suppose that my goal is to spend the shortest amount of time on my commute home from work so that I can be on time for a dinner date. Also suppose, for simplicity, that there are only three possible routes that I can take: route 1 is the shortest, whereas route 2 is longer but scenic and 3 is more direct but has potentially more traffic, especially during rush hour. I am leaving early from work today so that I can make my dinner date but before I leave, I check the traffic and learn that there has been a wreck on route 1. Thus, I must choose between routes 2 and 3. I reason that since I am leaving earlier, route 3 will be the quickest since there won’t be much traffic while I’m on my early commute home. So I take route 3 and arrive home in a timely manner: mission accomplished. The compatibilist would say that this is a paradigm case of a free action (or a series of free actions). The decisions I made about how to get home were drive both by my desire to get home quickly and also by the information I was operating with. Assuming that that information was good and I reasoned well with it and was thereby able to accomplish my goal (that is, get home in a timely manner), then my action is free. My action is free not because my choices were undetermined, but rather because my choices were determined (caused) by my own best reasons—that is, by my desires and informed beliefs. The incompatibilist, in contrast, would say that an action is free only if I could have chosen otherwise, given all the same conditions again. But think of what that would mean in the case above! Why on earth would I choose routes 1 or 2 in the above scenario, given that my most pressing goal is to be able to get to my dinner date on time? Why would anyone knowingly choose to do something that thwarts their primary goals? It doesn’t seem that, given the set of beliefs and desires that I actually had at the time, I could have chosen otherwise in that situation. Of course, if you change the information I had (my beliefs) or you change what I wanted to accomplish (my desires), then of course I could have acted otherwise than I did. If I didn’t have anything pressing to do when I left work and wanted a scenic and leisurely drive home in my new convertible, then I probably would have taken route 2! But that isn’t what the incompatibilist requires for free will. As we’ve seen, they require the much stronger condition that one’s action be such that it could have been different even if they faced exactly the same condition over again. But in this scenario that would be an irrational thing to do. Of course, if one’s goal were to be irrational and to thwart one’s own desires, I suppose they could do that. But that would still seem to be acting in accordance with one’s desires.

Many times free will is treated as an all or nothing thing, either humans have it or they don’t. This seems to be exactly how the libertarian and hard determinist see the matter. And that makes sense given that they are both incompatibilists and view free will and determinism like oil and water—they don’t mix. But it is interesting to note that it is common for us to talk about decisions, action, or even whole lives (or periods of a life) as being more or less free . Consider the Goethe quotation at the beginning of this chapter: “none are more enslaved than those who falsely believe they are free.” Here Goethe is conceiving of freedom as coming in degrees and claiming that those who think they are free but aren’t as less free than those who aren’t free and know it. But this way of speaking implies that free will and determinism are actually on a continuum rather than a black and white either or. The compatibilist can build this fact about our ordinary ways of speaking about freedom into an argument for their position. Call this the argument from ordinary language . The argument from ordinary language is that only compatibilism is able to accommodate our common way of speaking about freedom coming in degrees—that is, as actions or decisions being more or less free. The libertarian can’t account for this since the libertarian sees freedom as an all or nothing matter: if you couldn’t have done otherwise then your action was not truly free; if you could have done otherwise, then it was. In contrast, the compatibilist is able to explain the difference between more/less free action on the continuum. For the compatibilist, the freest actions are those in which one reasons well with the best information, thus acting for one’s own best reasons, thus furthering one’s interests. The least free actions are those in which one lacks information and reason poorly, thus not acting for one’s own best reasons, thus not furthering one’s interests. Since reasoning well, being informed, and being reflective are all things that come in degrees (since one can possess these traits to a greater or lesser extent) and since these attribute define what free will is for the compatibilist, it follows that free will comes in degrees. And that means that the compatibilist is able to make sense or a very common way that we talk about freedom (as coming in degrees) and thus make sense of ourselves, whereas the libertarian isn’t.

There’s one further advantage that compatibilists can claim over libertarians. Libertarians defend the claim that there are at least some cases where one exercises one’s free will and that this entails that determinism is false. However, this leaves totally open the extent of human free will. Even if it were true that there are at least some cases where humans exercise free will, there might not be very many instances and/or those decisions in which we exercise free will might be fairly trivial (for example, moving one’s finger to the left or right). But if it were to turn out that free will was relatively rare, then even if the libertarian were correct that there are at least some instances where we exercise free will, it would be cold comfort to those who believe in free will. Imagine: if there were only a handful of cases in your life where your decision was an exercise of your free will, then it doesn’t seem like you have lived a life which was very free. In other words, in such a case, for all practical purposes, determinism would be true.

Thus, it seems like the question of how widespread free will is is an important one. However, the libertarian seems unable to answer it for reasons that we’ve already seen. Answering the question requires knowing whether or not one could have acted otherwise than one in fact did. But in order to know this, we’d have to know how to answer a strange counterfactual—whether I could have acted differently given all the same conditions. As noted earlier (“the epistemic objection”), this raises a tough epistemological question for the libertarian: how could he ever know how to answer this question? And so how could he ever know whether a particular action was free or not? In contrast, the compatibilist can easily answer the question of how widespread free will is: how “free” one’s life is depends on the extent to which one’s actions are driven by their own best reasons. And this, in turn, depends on factors such as how well-informed, reflective, and reasonable a person is. This might not always be easy to determine, but it seems more tractable than trying to figure out the truth conditions of the libertarian’s counterfactual.

In short, it seems that compatibilism has significant advantages over both libertarianism and hard determinism. As compared to the libertarian, compatibilism gives a better answer to how free will can come in degrees as well as how widespread free will is. It also doesn’t face the arbitrariness objection or the epistemic objection. As compared to the hard determinist, the compatibilist is able to give a more satisfying answer to the moral responsibility issue. Unlike the hard determinist, who sees all action as equally determined (and so not under our control), the compatibilist thinks there is an important distinction within the class of human actions: those that are under our control versus those that aren’t. As we’ve seen above, the compatibilist doesn’t see this distinction as black and white, but, rather, as existing on a continuum. However, a vague boundary is still a boundary. That is, for the compatibilist there are still paradigm cases in which a person has acted freely and thus should be held morally responsible for that action (for example, the person who embezzles money from a charity and then covers it up) and clear cases in which a person hasn’t acted freely (for example, the person who was told to do something by their boss but didn’t know that it was actually something illegal). The compatibilist’s point is that this distinction between free and unfree actions matters, both morally and legally, and that we would be unwise to simply jettison this distinction, as the hard determinist does. We do need some distinction within the class of human actions between those for which we hold people responsible and those for which we don’t. The compatibilist’s claim is that they are able to do with while the hard determinist isn’t. And they’re able to do it without inheriting any of the problems of the libertarian position.

Study questions

  • True or false: Compatibilists and libertarians agree on what free will is (on the concept of free will).
  • True or false: Hard determinists and libertarians agree that an action is free only when I could have chosen otherwise than I in fact chose.
  • True or false: the libertarian gives a transcendental argument for why we must have free will.
  • True or false: both compatibilists and hard determinists believe that all human actions are determined.
  • True or false: compatibilists see free will as an all or nothing matter: either an action is free or it isn’t; there’s no middle ground.
  • True or false: compatibilists think that in the case of a truly free action, I could have chosen otherwise than I in fact did choose.
  • True or false: One objection to libertarianism is that on that view it is difficult to know when a particular action was free.
  • True or false: determinism is a fundamental assumption of the natural sciences (physics, chemistry, biology, and so on).
  • True or false: the best that support the libertarian’s position are cases of very simple or arbitrary actions, such as choosing to move my finger to the left or to the right.
  • True or false: libertarians thinks that as long as my choices are caused by my desires, I have chosen freely.

For deeper thought

  • Consider the Shopenhauer quotation at the beginning of the chapter. Which of the three views do you think this supports and why?
  • Consider the movie The Truman Show . How would the libertarian and compatibilist disagree regarding whether or not Truman has free will?
  • Consider the Tolstoy quote at the beginning of the chapter. Which of the three views does this support and why?
  • Consider a child being raised by white supremacist parents who grows up to have white supremacist views and to act on those views. As a child, does this individual have free will? As an adult, do they have free will? Defend your answer with reference to one of the three views.
  • Consider the eccentric neuroscientist example (above). How might a compatibilist try to show that this isn’t really an objection to her view? That is, how might the compatibilist show that this is not a case in which the individual’s action is the result of a well-informed, reflective choice?
  • Actually, the phrase was originally a Latin phrase, not an English one because at the time in Medieval Europe philosophers wrote in Latin. ↵
  • On the other hand, if these nonphysical decisions are supposed to have physical effects in the world (such as causing our behaviors) then although there is no problem with the agent’s decision itself being uncaused, there would still be a problem with how that decision can be translated into the physical world without violating the law of conservation of energy. ↵
  • One well-known and influential attempt to reconcile moral responsibility with determinism is P.F. Strawson’s “Freedom and Resentment” (1962). ↵
  • The term “epistemic” just denotes something relating to knowledge. It comes from the Greek work episteme, which means knowledge or belief. ↵

Introduction to Philosophy Copyright © by Matthew Van Cleave is licensed under a Creative Commons Attribution 4.0 International License , except where otherwise noted.

Share This Book

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

The Oxford Handbook of Free Will (2nd edn)

  • < Previous chapter
  • Next chapter >

The Oxford Handbook of Free Will (2nd edn)

4 Chaos, Indeterminism, and Free Will

Robert C. Bishop is associate professor of Physics and Philosophy and the John and Madeleine McIntyre Professor for the Philosophy and History of Science at Wheaton College. He has published numerous articles on reduction and emergence, nonlinear dynamics, complexity, determinism and free will. His most recent book is The Philosophy of the Social Science.

  • Published: 18 September 2012
  • Cite Icon Cite
  • Permissions Icon Permissions

This article begins with a discussion of modern efforts to clarify and define the meaning of physical determinism. It distinguishes four features of the Laplacean vision of physical determinism—differential dynamics, unique evolution, value determinateness, and absolute prediction—and the relevance of each to free-will debates. It then turns to the role of indeterminism in quantum mechanics and discusses current philosophical debates about the nature of indeterministic or probabilistic causation. It also considers debates about the possible relevance of chaos theory and nonlinear dynamics in physical systems to free will as well as the possible relevance of recent research on far-from equilibrium physical systems pioneered by Nobel laureate Ilya Prigogine. The article concludes with some general remarks about the causal completeness of physical explanations and the possibility of emergent phenomena in physical systems.

Free-will debates can be boiled down to the question of how to make sense of human free choice and action in a realm of ordered causes. In the modern era, this question has been cast as the question of free will and determinism. We face an immediate problem when framing the question this way because being clear about the nature of determinism is notoriously difficult. At an abstract level of analysis, there are at least ninety varieties of determinism (Sobel 1998 , 77–166). Furthermore, there are several ways to construe determinism (e.g., physical, psychological, theological, logical, metaphysical).

Because the question of how physical laws are related to the exercise of free will appears explicitly in free-will discussions (e.g., see the essays in this volume on the Consequence Argument; Kane 1996 ), my primary focus will be on physical determinism, the thesis that all physical events are determined to occur according to physical laws. This choice of focus only removes part of the complication, however, because there are several conceptions of physical determinism as well. Physical determinism could be viewed as a metaphysical thesis in which all of physical reality is ontologically determined. (This appears to be how most free-will discussants construe physical determinism.) Or it could be viewed as an observed phenomenon of our experience. Alternatively, physical determinism might be a concept only relevant to the mathematical models of physics and other physical sciences, although its relevance to the world of everyday choice and action is questionable. (See the next essay of this volume by Bishop and Atmanspacher.) Indeed, since the seventeenth century the tendency has been to use the supposedly deterministic theories of physics as the models for construing all other forms of determinism, metaphysical or otherwise. However, if thoughts, feelings, and desires are not physical events, it is unlikely that physical theories are appropriate models for thinking about such nonphysical events.

Physical Determinism

There have been numerous attempts to clarify and explain physical determinism. 1 Syntactic approaches favored by the logical positivists have largely been abandoned, having proven inadequate for describing scientific practices (Bishop and Kronz 1999 , 129–30). Even innovative approaches trying to combine formal systems and semantics (e.g., Richard Montague 1974 , especially 303–60) are not free of serious limitations. The types of second-order formal languages Montague studies can only represent a finite number of physical magnitudes and, therefore, are inadequate for cases where an uncountable number of distinct physical magnitudes interact to produce determinism. Instead, the laws and models of physics are more appropriately described by mathematical equations. The important questions for physical determinism in such laws and models are the existence and uniqueness properties of the solutions as well as dynamics, not quantification or logical entailment.

The mathematical and dynamical character of physical determinism is captured well by what has been called the “Laplacean vision” of determinism:

Differential dynamics . There exists a nonprobabilistic algorithm relating a state of a model at any given time to a state at any other time.

Unique evolution . A model is such that a given state is always followed by the same history of state transitions.

Value determinateness . Any state of a model can be described with arbitrarily small (nonzero) error.

  Absolute prediction . Any state of a model can be generated from the algorithm with arbitrarily small (nonzero) error from any other state of the model. 2

The equations of physical theories, along with their initial and boundary conditions, provide the motivation for DD expressing the Laplacean belief that there are no indeterministic events in classical physics. UE is closely associated with DD and captures the Laplacean belief that models of classical physics will repeat their behaviors exactly if the same initial and boundary conditions are specified. The third element of this vision, VD, is motivated by the Laplacean belief that there is nothing in principle in classical physics that prevents mathematical descriptions of arbitrary accuracy. The final element, AP, is a prima facie reasonable expectation that given DD, UE, and VD, it should be possible in principle to predict the exact state a model would take on at any time. This implication, however, does not follow (Bishop 2003a ). An example of the Laplacean vision would be the equations of motion describing the flight of a cannon ball. Given identical physical conditions (e.g., the initial position and velocity of the cannon ball on exiting the cannon, wind, air pressure, temperature) the cannon ball will follow the identical trajectory in the state space characterized by the equations of motion. 3

Unique Evolution

Formally, UE may be stated as follows: Let M stand for the collection of all models sharing the same set L of physical laws and suppose that P is the set of relevant physical properties for specifying the time evolution of a model recognized by L . Then

A model m ∈ M exhibits UE if and only if every model m ’ ∈ M isomorphic to m with respect to P undergoes the same evolution as m (Bishop and Kronz 1999 , 131). 4

UE can be given two readings. It can be construed as a statement of “causal determinism” where “every event has a cause that is an event that takes place at some antecedent time or times” (Sobel 1998 , 84). Or it may also be read as a statement of “block-universe determinism” characterized by William James ( 1956 , 150) as “[t]he whole is in each and every part, and welds it with the rest into an absolute unity, an iron block, in which there can be no equivocation or shadow of turning.”

These two forms of determinism can be distinguished (Sobel 1998 , 102–5). The former derives from the causal principle that every event has an antecedent cause, a flow from cause to effect, if you will, that may be continuous or have gaps. The latter reflects the intuition that a difference anywhere in the universe requires a difference everywhere. For example in this context Sobel ( 1998 , 89) distinguishes “fast-starting” series of causally linked states. These are series where every state has a temporally antecedent determining cause, but the series itself has no antecedent deterministic cause (its beginning is undetermined by prior events and may have a probabilistic cause) and no state in the series occurs before a specified time. The causal principle that every event has an antecedent cause would fail for a fast-starting series as a whole, though it would apply within such a series. This would be an example where causal determinism failed, but where block-universe determinism would still hold. 5 On the other hand, if one explicates the causal principle in terms of laws L and properties P , then the idea that a difference anywhere in a model m ’ isomorphic to m requires a difference everywhere in m ’ with respect to m can be explicated in terms of differences in either L or P .

Fixed Laws, the Past, and the Future

I will make no attempt to explicate the notion of physical laws other than to note the fact that in classical physics laws are often taken to be expressed by the equations fulfilling DD, UE, and VD. There is currently no consensus among philosophers of science on a problem-free conception of natural laws (e.g., Bunge 1998 ; Cartwright 1983 , 1989 ; Giere 1999 ; Suppe 1989 ; Van Fraassen 1989 , 1991 ) 6 but, fortunately, this minimal construal of physical laws is sufficient for many free-will discussions such as the various versions of the Consequence Argument. By restricting the notion of laws of nature occurring in the Consequence Argument to the set L ( P should also be included, strictly speaking), one can discuss the fixedness of the past in terms of physical principles or models. One might be tempted to construe the evolution of the universe as evidence of change in L or P over time or the emergence of new laws or properties in the history of this evolution. In the practice of physics, however, this evolution is seen as the unfolding of the consequences of L and P , and the history of the discovery of new laws or properties is simply the temporal discovery of fixed elements of the sets L and P that have always existed since some beginning (e.g., the Big Bang).

There are cases, such as the standard models of the evolution of the universe, in which particular epochs are characterized by a smaller set of laws and properties, whereas others exhibit the emergence of new laws and properties (Kolb, Turner, Lindley, Olive, and Seckel 1986 ; Kolb and Turner 1990 ). In such cases, these “new” laws and properties are viewed as consequences of the “old” laws and properties. Often the former can be mathematically derived from the latter given appropriate limiting conditions (though this derivation is not a logical entailment; see Bishop and Atmanspacher 2006 ; Primas 1998 ). In this sense, the set of laws and properties are still considered as fixed. One can object to construing L and P as fixed for these cases, but over the time span relevant to human history, we have no evidence suggesting that L and P are anything but fixed, so the history of human choices and actions is typically thought of as taking place within a domain over which L and P are unchanging. 7

Fixed L and P , combined with DD, UE, and VD, leads to a strong sense in which past and future events are fixed such that given the same L and P for two identical models m and m ’, the past histories of both m and m ’ will be identical as well as their future histories. If this conception is applied to the history of all events in the world, then the world's past and future history of events—including, for example, all human motives, reasons, decisions, actions—are fixed and the Consequence Argument, under the Laplacean vision, bites very deeply. For instance, given the past of our world, Sally chooses to go for a walk at noon today rather than for a bike ride. In contrast, if the past were slightly different than what it was, she would choose to go for a bike ride at noon today rather than a walk. In many free-will discussions an appeal to a slightly different past is invoked to cash out how Sally's desires for a walk rather than a bike ride might lead to her ability to do otherwise. However, if UE applies to all events in our world, then there was always some unique unalterable history of choices and actions leading from the past to Sally's choice at noon (and, likewise, there is a unique unalterable history of choices and actions for Sally's complete future). This is the price we pay if determinism in free-will debates is modeled after physical determinism and that determinism applies to all events: game, set, and match to the determinist position on free will.

Indeterminism

The past (as well as the future) history of events may not be fixed even though L and P are fixed, however, if indeterminism is present. 8 Indeterminism can enter into the models of physics given fixed L and P by making various modifications to the Laplacean vision. There are two ways of achieving indeterminism. The first is to modify DD by making the equations of the model irreducibly indeterministic or by introducing irreducibly indeterministic initial or boundary conditions. The second way to indeterminism is to drop UE (e.g., the famous Epicurean “swerve” of the atom where indeterminism emerges from the lack of unique evolution. 9 Clearly, the first strategy implies the second as irreducibly indeterministic equations, initial or boundary conditions guarantee that it is a contingent matter whether identical models will follow the same history of state transitions. The second strategy does not imply the first, because the loss of UE alone does not imply that the evolution equations be explicitly probabilistic (Bishop 2005 ; Van Fraassen 1991 , 51). For our models obeying L and P to have any intelligible content other than that they are indeterministic, however, an explicitly probabilistic prescription is required. 10

Quantum Mechanics, Probabilistic Causation, and the Nature of Indeterminism

This lack of explanation (or sufficient reason) can lead one to think that indeterministic models are noncausal even if probabilities are brought explicitly into the mathematical models. 11 This thought forms the basis for one of the most common objections to incompatibilist theories of free will, which make use of indeterminism in some crucial fashion: Namely, if indeterminism plays an important role in determining the outcome of Sally's decision to walk at noon today, then there would be no sufficient reason for her having chosen to walk rather than ride her bike. Everything was a matter of chance and, as such, explains neither Sally's decision nor her power to decide (see Kane 1996 , 105–23 for a discussion of and possible response to this type of argument).

Quantum mechanics is taken to be the paradigm indeterministic theory in physics. Its explicit probability assignments have been confirmed in a wide range of laboratory experiments and have made possible a number of now common devices found in everyday life (e.g., transistors, lasers, NMR scanners). Let me illustrate the difference between the quantum mechanical world and our everyday world in the following way. In our everyday world in the USA, stoplights operate with a predictable pattern of green, then yellow, then red. To get a sense for how different the quantum realm is, suppose quantum stoplights have two possible patterns: green, yellow, red, or green, red, yellow.

The key feature of quantum stoplights is that they have a 50 percent probability for exhibiting the green-yellow-red pattern and 50 percent for exhibiting the green-red-yellow pattern. Let me emphasize that the probability refers to the green-yellow-red pattern or the green-red-yellow pattern, not to the appearance of any individual color. 12 If you were approaching an intersection having a quantum stoplight that was currently yellow, you would not know if the light was going to turn red or green next, because you would not know what pattern the light was exhibiting based on the color you were seeing. You could observe the quantum stoplight over a long period of time to determine probabilities for the two patterns. But, according to conventional interpretations of quantum mechanics, you have no way of knowing in advance which pattern the light will exhibit as you come to the intersection because the patterns at the level of single observations are indeterministic.

If the quantum stoplights are observed to be in a particular pattern, an explanatory account can be given for why that pattern was exhibited in terms of the equations and probabilities governing the model. Take radioactive decay as a concrete example. Our models for radioactive decay do not allow us to predict the precise time a radioactive nucleus will decay, but only the probabilities for a decay event in a specified time range. We can then take measurements to test the predictions. Based on the physics of radioactive decay we can explain why the atom decayed in the time range it did with the predicted probability. 13

Radioactive decay, the tunneling of particles through energy barriers, and the absorption and emission of light are typical quantum phenomena that fit this kind of probabilistic form of explanation. Philosophers of science have developed a rich literature on the notion of indeterministic or probabilistic causation. 14 The general idea of probabilistic causation is for an event (or set of events) C to be a cause of an event E, C 's occurring must raise the probability of E 's occurring and this probability is conditional on an appropriately chosen set of background factors B . The probability must be conditionalized on background factors so that the increase in probability picks out a genuine case of causation rather than a spurious case of accidental correlation. There are two particularly difficult problems for theories of probabilistic causation: (1) The causal relevance of (some of the events) C may appear or disappear as the partition of the factors B is refined (e.g., in terms of the level of detail). Related to this problem is (2) the fact of C 's occurring raises the probability of E 's occurring on B (and that causally irrelevant events do not) should be robust with respect to B in the sense that additions to B (e.g., new evidence) do not introduce such changes to the probabilities that these facts no longer hold. Although these problems are difficult to address in an abstract general theory (indeed, they tell against the prospects for such a general theory), focusing on a specific context like quantum mechanics makes the difficulties more tractable (though not necessarily less messy!). We can then reference the appropriate laws in the set L and properties in the set P as well as the particular experimental setup for the background B and explicate causal relevance and robustness in terms of B .

Although it is possible to give a philosophical account of probabilistic causation in quantum mechanics, the nature of the probability involved remains an open question. Let us return to our quantum stoplights for a moment. There is a 50 percent probability for each pattern to be exhibited by such lights upon any given approach to the intersection. The key question is whether to understand the nature of this probability as epistemic or ontic. Along epistemic lines, one possibility is that there is some additional factor (i.e., a hidden mechanism) such that once we discover and understand this factor, we would be able to predict the observed behavior of the quantum stoplight with certainty (physicists call this approach a “hidden variable theory”; see, e.g., Bell 1987 , 1–13, 29–39; Bohm 1952a , 1952b ; Bohm and Hiley 1993 ; Bub 1997 , 40–114; Holland 1993 ; see also the preceding essay in this volume by Hodgson). Or perhaps there is an interaction with the broader environment (e.g., neighboring buildings, trees) that we have not taken into account in our observations that explains how these probabilities arise (physicists call this approach decoherence or consistent histories 15 ). Under either of these approaches, we would interpret the observed indeterminism in the behavior of the stoplights as an expression of our ignorance about the actual workings. Under an ignorance interpretation, indeterminism would not be a fundamental feature of quantum stoplights, but merely epistemic in nature due to our lack of knowledge about the system. Quantum stoplights would turn out to be deterministic after all.

The alternative possibility is that the observed indeterminism is ontic in the sense that there are no factors that fully determine what pattern the stoplights are going to exhibit at any given moment as in the so-called von Neumann projection postulate 16 or quantum stochastics. 17 Under an ontic interpretation, indeterminism would be a fundamental feature of quantum stoplights.

Free Will and Physics

By now perhaps it is clear how physics in the form of the laws L and properties P under the Laplacean vision of determinism has potential relevance to free-will discussions. The fundamental principles of physics certainly play a role in the brain (a bio physical organ) that in turn is involved in consciousness and free will. 18 If classical physics in this Laplacean picture is the whole story of the matter, then whatever free will amounts to, it must be influenced by L and P .

Quantum mechanics is typically thought of as more fundamental than classical mechanics and is one of the most empirically successful of our physical theories, so many authors have looked to it to support theoretical accounts of human freedom. 19 Still, it has been less clear to many that quantum mechanics is relevant to free will. For example, an early objection to quantum effects influencing human volitions was offered by philosopher J. J. C. Smart ( 1963 , 123–24). Smart accepted the truth of indeterminism at the quantum level, but argued that the brain remains deterministic in its operations because microscopic events are insignificant by comparison. After all, a single neuron is known to be excited by on the order of a thousand molecules, each molecule consisting of ten to twenty atoms. Quantum effects though substantial when focusing on single atoms are presumed negligible when focusing on systems involving large numbers of atoms. Therefore, it looks like quantum effects would be too insignificant in comparison to the effects of thousands of molecules to play any possible role in deliberation.

Routes for Amplification: Chaos and Nonlinear Dynamics

Arguments such as Smart's do not take into consideration the possibility for amplifying quantum effects through the interplay between chaos (Baker and Gollub 1990 ; Hilborn 2001 ) in the domain of the classical world (e.g., brains, weather systems) on the one hand and quantum effects (e.g., tunneling through energy barriers, radioactive decay) on the other. Philosophers Jesse Hobbs ( 1991 ) and Stephen Kellert ( 1993 , 69–75) have argued that chaos in classical systems can amplify quantum fluctuations due to sensitivity to the smallest changes in initial conditions. Briefly, the reasoning behind such sensitive dependence arguments runs as follows. Given two chaotic models m and m ’ of classical mechanics in nearly identical initial states (e.g., specification of the initial positions and velocities of the model that differ ever so slightly), they will evolve in radically different ways in a relatively short time period as the slight differences in initial conditions are amplified. There is no known lower limit to this sensitivity, thus affording the possibility of chaotic macroscopic systems being sensitive to quantum fluctuations because quantum mechanics sets a lower bound on how precise the initial conditions actually can be. Hence, UE must fail for chaotic models in classical mechanics.

In a rather crude simplification to illustrate the ideas, suppose the patterns of neural firings in the brain correspond to decision states. Chaos could amplify quantum events causing a single neuron to fire that would not have fired otherwise. If the brain (a macroscopic object) is also in a chaotic dynamical state, making it sensitive to small disturbances, the effect of this additional neural firing, small as it is, would then be further amplified to the point where the brain states would evolve differently than if the neuron had not fired. In turn, these altered neural firings and brain states would carry forward such quantum effects affecting the outcomes of human choices. 20

There are several objections to this line of argument. First, the presence of chaos in the brain and its operations is a hotly debated empirical matter. 21 On the one hand, various signs of chaos are detected in neural and cognitive dynamics experiments such as sensitive dependence and so-called pink noise. On the other hand, appropriate technical definitions for chaos and its signatures have their difficulties (Bishop 2008b , sec. 1.2 and references therein) and it is rare that all the signatures associated with chaos are present in any given experimental observation of brain dynamics. Moreover, it is very difficult to distinguish empirically between chaos and random noise, particularly in such messy environments as brains. However, the exquisite sensitivity needed for both the sensitive dependence arguments and the neural amplification of quantum effects is a general feature of nonlinear dynamics and is present whenever nonlinear effects are likely to make significant contributions to the dynamics of the system (Bishop 2008a , 2011 ). The presence of chaos per se in neural dynamics is not essential for those dynamics to be sensitively dependent to various kinds of small changes in its inputs or environment. 22

A second objection to these kinds of sensitivity arguments is that they depend crucially on how quantum mechanics itself and measurements are interpreted (Bishop 2008a ). As noted earlier, some versions of quantum mechanics are ontically deterministic whereas others are ontically indeterministic, implying the nature of any quantum effects amplified by chaos is open to interpretation. 23 Though neural dynamics likely exhibits the nonlinearity that allows quantum events to be amplified to the macroscopic level of neural assemblies, at least occasionally, it is an open question as to whether these quantum effects are deterministic or indeterministic until the interpretation of quantum mechanics and the measurement problem are clarified. 24

A third objection is that although abstract sensitive dependence arguments do correctly lead to the conclusion that the smallest of effects can be amplified, applying such arguments to concrete physical systems shows that the amplification process may be severely constrained (Bishop 2008a ). For instance, investigating the role of quantum effects in the process of friction in sliding surfaces indicates quantum effects can be amplified by nonlinearities to produce a difference in macroscopic behavior, provided that the effects are large enough to break molecular bonds and are amplified quickly enough. In the case of the brain, we currently do not know what constraints on amplification exist. 25

Routes for Amplification: Nonequilibrium Statistical Mechanics

An alternative possibility for amplifying indeterminism is suggested by the research on far-from-equilibrium systems by Ilya Prigogine and his Brussels-Austin Group. Their work offers reasons to search for a different type of indeterminism in both the microphysical and macrophysical domains (Bishop 2004 ; Prigogine 1997 ).

If a system of particles is distributed uniformly in position in a region of space, the system is said to be in thermodynamic equilibrium (e.g., cream uniformly distributed throughout a cup of coffee). In contrast a system is far-from-equilibrium (i.e., nonequilibrium) if the particles are arranged so that highly ordered structures appear (e.g., a cube of ice floating in tea). The following properties characterize nonequilibrium statistical systems:

Large number of particles

High degree of structure and order

Collective behavior

Irreversibility

Emergent properties 26

The brain possesses all these properties, so work of the Brussels-Austin group can be applied to analyzing the brain as a nonequilibrium system (though Prigogine never pursued such applications).

Let me quickly sketch a simplified version of the approach to point out why the developments of the Brussels-Austin Group offer a possible alternative for investigating the connections between physics and free will. Conventional physics describes physical systems using particle trajectories as a fundamental explanatory element of its models (as noted in the earlier articulation of the Laplacean view). This means that the behavior of a model is derivable from the trajectories of the particles composing the model. The equations governing the motion of these particles are reversible with respect to time (they can be run backwards and forwards like a film). When there are too many particles involved to make these types of calculations feasible (as in gases or liquids), coarse-grained averaging procedures are used to develop a statistical picture of how the system behaves rather than focusing on the behavior of individual particles.

In contrast, the Brussels-Austin approach views these systems in terms of models whose fundamental explanatory elements are distributions; in other words, the arrangements of the particles are the fundamental explanatory elements and not the individual particles and trajectories. The equations governing the behavior of these distributions are generally irreversible with respect to time. In addition, focusing exclusively on distribution functions opens the possibility that macroscopic nonequilibrium models are irreducibly indeterministic, an indeterminism that has nothing to do with ignorance about the system or the vagaries of quantum mechanics. If true, this would mean that probabilities are as much an ontologically fundamental element of the macroscopic world as they are of the microscopic, while being free of the interpretive difficulties found in conventional quantum mechanics.

One important insight of the Brussels-Austin Group's shift away from trajectories to distributions as fundamental elements is that explanation also shifts from a local context (i.e., a set of particle trajectories) to a global context (i.e, the distribution of the entire set of particles). Systems acting as a whole may produce collective effects that are not reducible to a summation of the trajectories and subelements that make up the system (Bishop 2004 , 2008c , 2011 ; Bishop and Atmanspacher 2006 ). The brain exhibits this type of collective behavior in many circumstances (Engel, Roelfsema, König, and Singer 1997 ; Cosmelli, Lachaux, and Thompson 2007 ) and the work of Prigogine and his colleagues gives us another tool for trying to understand that behavior. Though it is still speculative and contains open technical questions (Bishop 2004 ), this approach offers a potential alternative for exploring the relationship between physics and free will as well as pointing to a new possible source for indeterminism to be explored in free-will theories. 27

Determinism, Causal Completeness of Physics, and Free Will

So far I have assumed a rather straightforward relationship between physics and free will when in actually this relationship is much more subtle. I will briefly mention two subtleties that have important implications for thinking about this relationship.

Subtleties Regarding Physical Determinism

The presumed deterministic character of classical physics has served as an influential model for conceiving of metaphysical, theological, and other forms of determinism since the early eighteenth century. Physical theories were viewed through the lens of the mechanical world picture that became the dominant picture of the world in the seventeenth century as promulgated by such luminaries as Boyle, Descartes, and Leibniz. Physical theories were also influenced by that century's fascination with the many new machines that had been developed. 28 Combined with the intense desire to mathematize all aspects of nature to the extent possible, the mechanical world picture led to the Laplacean vision of classical physics sketched in an earlier section. 29 Though QM has put a significant dent in the mechanical world picture, this vision of determinism implicitly or explicitly shapes the background for contemporary free-will debates. 30

However, our best physical theories are terribly ambiguous regarding the status of physical determinism. There are several examples of violations of UE in classical physics and the fate of UE in the special and general theories of relativity is currently unclear (Bishop 2006a ; Earman 1986 ). As noted earlier, there are both deterministic and indeterministic versions of quantum mechanics that are empirically equivalent to the best of our current knowledge. Hence, someone looking to our best physical theories as models for determinism in free-will debates is actually invoking a deterministic interpretation of those theories rather than some unambiguous feature of those theories.

Not only are our current best physical theories remarkably unclear about the truth of determinism in the physical sciences, there is a further significant issue regarding their application to metaphysical questions about our world in the context of any nonlinear systems. In physical theories, we typically invoke the “faithful model assumption”: The state of a system is faithfully characterized by the values of the crucial variables and a physical state of that system faithfully corresponds to a point in the model's mathematical state space through these values. In many contexts, this assumption is fairly unproblematic. However, if the system in question is nonlinear—in other words, has the property that a small change in the state or conditions of the system is not guaranteed to result in a small change in the system's behavior—this assumption faces serious difficulties (see Bishop 2008a ; 2011 for discussion). 31 Because our world is populated with such systems, one's purchase on applying our best physical laws and theories to our world to answer larger metaphysical questions about determinism is problematic.

Subtleties Regarding Causal Completeness

Another relevant subtlety is the causal completeness of physics, the metaphysical thesis that every physical event in the world has a physical cause (Kim 2005 ; Papineau 2002 ). If this thesis is true, then the conditions necessary for human action are not satisfied and neither compatibilist nor incompatibilist notions of free will are possible for our world. 32 However, the difficulties raised by the faithful model assumption, just mentioned, imply similar difficulties for inferring, from our best physical theories, that our world is causally closed at the physical level. Furthermore, even if our physical theories do not suffer any serious problems regarding the faithful model assumption, the causal completeness of the physical is, at most, a typicality condition (Bishop 2006b ). This means that in the absence of nonphysical influences, physical events will proceed typically as our physical theories describe, provided that the system in question is appropriately isolated from nonphysical effects during its evolution.

If the causal closure of the physical were true, therefore, in principle, physical theories would describe a set of necessary and sufficient conditions for all physical events, human choices and actions included. The failure of causal closure means that physical theories only describe a set of necessary conditions for all physical events. In the absence of any other events, physical events proceed typically, but in the presence of any additional conditions, the situation is more complicated. Human desires, feelings, and values leading to choices and actions would be examples of events in which physical conditions are necessary but not sufficient for their complete description. This latter situation is more appropriately characterized by “contextual emergence,” where domain A provides necessary conditions for the description or existence of events of domain B , but lacks sufficient conditions for the description or existence of events of domain B (Atmanspacher and Bishop 2007 ; Bishop 2010b ; Bishop and Atmanspacher 2006 ). In other words, the sufficient conditions necessary to complete a set of jointly necessary and sufficient conditions for the description or existence of events of domain B cannot be obtained from domain A alone. Information from domain B —a new context—is crucially needed. The failure of the causal closure of the physical implies that theories like QM are less relevant to free will and action than is typically supposed (see the next essay by Bishop and Atmanspacher).

Concluding Thoughts

One way to characterize free-will debates is as an attempt to find a home for free will in a physical world picture that seems hostile to such freedom under the threat of determinism and the mechanical world picture. There would seem to be no viable sense of free will without some form of determination or ordered realm of causes and influences in which to act and make a difference. On the other hand, the freedom in question has to be real and meaningful and cannot just amount to the effect of causes that play upon the human agent (Alison 1997 ). It is possible to argue that the full reality of free will and constraints upon it can only be fully explicated in terms of an emergent realm of human influences and counterinfluences that depend upon the physical world for its existence, but, as contextual emergence suggests, is not reducible to physical processes. Moreover, the attribution of physical determinism or indeterminism to systems turns out to also be a contextual rather than ontological affair. 33

The contemporary developments in determinism and physical theories surveyed in this essay indicate that the existence of pockets of determinism in physics do not imply that determinism holds sway over all domains of physics. Furthermore, our physical theories are unreliable guides to the ultimate metaphysical character of our world. So it seems that the existence of genuine human action is becoming more plausible for our world (Richardson and Bishop 2001 ). Contextual emergence (as exemplified in complex systems) provides hints that the physical conditions underlying human free will and action need further specification in appropriate contexts (see the section “Discussion and Perspectives” in the following essay by Bishop and Atmanspacher). Efforts in this ambitious direction are only just getting underway and will require a significant rethinking of our understandings of physical theories and causation (e.g., Bishop 2011 ). Moreover, such efforts will have to include a reexamination of our cultural ideals and value commitments, because these play a significant role in both our appraisals and applications of scientific theorizing to the domain of human activity (e.g., Bishop 2007 ; Richardson, Fowers, and Guignon 1999 ; Taylor 1989 , 2009 ). The resources for such an ambitious task are likely to be found in a broad range of developments from nonlinear dynamics and complexity, to cognitive and neural dynamics, to action theory and philosophical hermeneutics. Such a transdisciplinary effort will be needed to make sense of human free will and action in our world where determination, indeterminacy, and contingency are such important players in all domains.

See, e.g., Bishop ( 2005 ), Earman ( 1986 ), Kellert ( 1993 ), Laplace ( 1951 ), Montague ( 1974 ), Popper ( 1982 ), Russell ( 1953 ), Stone ( 1989 ), and Van Fraassen ( 1989 , 1991 ).

For extensive discusions, see Bishop ( 2002 , 112–14; 2003a ; 2005 ), Stone ( 1989 , 124–25), and Kellert ( 1993 , 49–62).

A state space is a mathematical space of points taken to faithfully represent the space of physical possibilities for the physical system being modeled by the equations of motion.

Both Hobbes and Kellert use John Earman's construal of Laplacean determinism (Earman 1986 , 13) to spell out UE. However, there are technical reasons for preferring the definition given in the text (Bishop and Kronz 1999 , 130–31).

Sobel ( 1998 , 97) suggests that such fast-starting series could be seen as free actions because they “leave open that they should be produced by ‘out-of-the-world’ agency,” but he is dubious that such agency makes sense. Perhaps this supposed need for “out-of-the-world” agency might be alleviated by shifting to a dialogical conception of agency. Here within-the-world processes of mutual influence and shaping among persons (e.g., conversations, parenting, mentoring) could provide a genuine source of influence for free actions provided that dialogical influence is treated at least on par with efficient causation (Bishop 2007 ; Gadamer 1989 ; Richardson and Bishop 2001 ). In this sense such influences are not “out-of-the-world”—they are part of the ordinary social world in which we live—but they likely are not fully determined by physical laws (see the section “Determinism, Causal Completeness of Physics, and Free Will,” below).

Among the more extreme and interesting positions on physical laws is that of Nancy Cartwright, who argues that our conception of nature as a seamless web of causal law-like connections is mistaken. Although short on details, John Dupré ( 1996 ), drawing on Cartwright's work, argues against the idea that every event is governed by some quantitatively precise law, so that the causal order turns out to be partial and incomplete. He conceives of human agents as sources of causal power and order acting in this partially complete causal web and bringing order to this web. Necessary to such an account is the realization that physical laws are quantitatively precise because they represent idealizations away from the larger context of myriad physical influences (e.g., a horribly complicated environment) as well as all such laws ignore the perspectives and agency of embodied beings (e.g., they take no account of human agency or interventions). Though Dupré ( 1996 , 400) does not propose a positive account of agency, he suggests that clarifying human agents as causal sources of freedom and action requires viewing agents as embedded in the languages and practices of society, which means broadening the context for discussions of free will and agency beyond the idealizations assumed in physical laws (e.g., Bishop 2007 ; Richardson and Bishop 2001 ).

Some multiverse theories permit scenarios in which different universes have the same laws but different conditions, or even universes that have different. Such scenarios can give new meaning to considering possible worlds where the past was different because either the laws or the conditions (or both) were slightly different than in our universe. A further complication arises from considering the possibility that new properties arise in physical systems that are not implied by L and P , but are due solely to the organization of the components of the systems in question (Beckermann, Flohr, and Kim 1992 ). A strong emergentist might argue that among these emergent properties are causal powers that, although not fully constrained by L, P and the constituents of the physical system, nevertheless are able to influence or manipulate the physical system in question. Jaegwon Kim (e.g., 1999 , 2005 ) has argued forcefully against this strong emergentist possibility. There are, however, numerous counterexamples to his arguments in physics and chemistry in phenomena such as temperature, chemical potential, chirality, and fluid convection (Bishop 2008c , 2010b , 2011 ; Bishop and Atmanspacher 2006 ; Primas 1998 ). What these counterexamples mean regarding the fixedness of L and P is an open question.

Earman ( 1986 ) has discussed several cases in the context of Newtonian mechanics in which determinism in the form of UE appears to fail. Particularly noteworthy are his discussions of so-called “space invader” cases (33–35, 45–47) and systems of colliding billiard balls (39–40). These cases have been critically discussed in Bishop 1999 , 34–39; Bishop and Kronz 1999 , 132–33. For an overview of the issues, see Bishop ( 2006a ).

Some have thought that relaxing VD would introduce indeterminism into the models of physics (e.g., Glymour 1971 , 744–45); however, determinism can be revised to allow for set- and interval-valued properties evolving along uniquely determined paths (Earman 1986 , 217–18; Fine 1971 ; Teller 1979 ).

I have stated this requirement for intelligibility only in the context of physical models and theories. This is not meant to suggest that human rationality is limited to algorithmic processing (either deterministic or probabilistic). See Hodgson (in the preceding essay of this volume) and Penrose (1991, 1994 ) for discussion.

This is mainly because the principle of sufficient reason does not hold for such models. An insightful treatment of this and related principles and axioms may be found in Kane ( 1986 ).

The patterns of my quantum stoplights are analogous to a two-state system. The green-yellow-red pattern would be an up state and the green-red-yellow pattern would be a down state. Approaching the intersection would be analogous to a measurement on the system.

Sometimes it is claimed that events such as a radioactive atom's decay are random in the sense of being lawless. This would be a misunderstanding of QM, however, as the probabilities that such decay events fulfill are governed by laws (sometimes called statistical laws).

See, e.g., Baker ( 1999 ); Collins ( 2000 ); DeVito ( 1996 ); Dowe and Noordhof ( 2003 ); Eells ( 1991 ); Good ( 1961 , 1962 ); Hall ( 2000 ); Harper and Skyrms ( 1988 ); Hitchock ( 1993 ); Humphreys ( 1980 , 25–57; 1989 ); Koons ( 2000 ); Lewis ( 1986a , 175–85; 2000 ); Mellor ( 1986 , 166–86; 1999 ); Menzies ( 1996 ); Noordhof ( 1998 , 1999 ); Paul ( 2000 ); Pearl ( 1997 , 2000 ); Reichenbach ( 1956 ); Salmon ( 1984 ; 1993 , 137–53; 1998 ); Schaffer ( 2000 ); Skyrms ( 1999 ); Skyrms and Harper ( 1988 ); Suppes ( 1970 ).

See, e.g., Giulini, Joos, Kiefer, Kupsch, Stamatescu, and Zeh ( 1996 ), Omnès ( 1994 ), Zurek ( 1981 , 1982 , 1991 ).

See, e.g., Rae ( 1986 ) and Von Neumann ( 1955 ).

See, e.g., Diósi ( 1988 , 1989 ); Diósi, Gisin, and Strunz ( 1998 ); Ghirardi, Rimini, and Weber ( 1986 ); Gisin ( 1984 , 1989 ); Gisin and Percival ( 1992 ); Plenio and Knight ( 1998 ); Von Neumann ( 1955 ).

Although not necessarily in a reductive sense (see Bishop 2006b ).

See, e.g., Beck and Eccles ( 1992 ); Compton ( 1935 ); Eccles ( 1970 ); Kane ( 1996 ); Penrose (1991, 1994 , 1997 ); and Stapp ( 1993 , 2005 ). Hodgson's essay in this volume discusses the views of a number of these authors.

Eccles ( 1970 ) and Kane ( 1996 ) make explicit use of such amplifications in their accounts of free will (see also Hodgson's discussion of Eccles in this volume).

See, e.g., Cosmelli, Lachaux, and Thompson ( 2007 ); Diesmann, Gewaltig, and Aertsen ( 1999 ); Freeman ( 1999 , 2000 ); Kaneko, Tsuda, and Ikegami ( 1994 , 103–89); Lehnertz, Elger, Arnhold, and Grassberger ( 2000 ); Van Orden, Holden, and Turvey ( 2003 , 2005 ); Vandervert ( 1997 ); see also chapter 27 , Walter's essay, in this volume.

For example, Brecht, Schneider, Sakmann, and Margrie ( 2004 ) found evidence that a change in a single neuron can evoke altered behavior in rats. Furthermore, general modeling studies strongly indicate a role for sensitive dependence in neural dynamics (Atmanspacher and Rotter 2008 , 310–12).

Ted Honderich ( 1988 , 269–304), among others, relies on these interpretive difficulties as an important line of argument against the relevance of quantum effects for questions about free will and determinism.

One possible way to resolve the impasse regarding the determinism/indeterminism of quantum mechanics is through nonlocality, as in the violations of Bell's inequalities (see Hodgson's essay this volume, specifically the section “Nonlocality”). On the one hand, special relativity forbids any faster-than-light signaling. On the other hand, in experiments quantum mechanics consistently violates the Bell inequalities. Roughly, these experiments seem to call into question a commonsense assumption about locality: If a pair of simultaneous measurements is made in two distant regions of an extended system, the two measurements should be independent of each other by special relativity's faster-than-light signaling prohibition. Thus, it appears that these violations have to be due to the nonlocal nature of quantum mechanics (Gisin 2008 , 2009a   2009b ; Popescu and Rohrlich 1994 ). However, if quantum mechanics is nonlocal, the only way to maintain the no faster-than-light signaling restriction of special relativity is if the outcomes of quantum measurement events are indeterministic (Gisin 2008 , 3–5; Popescu and Rohrlich 1994 ).

There are two other lines of argument regarding the relevance of quantum mechanics to free will worth mentioning, one by Roger Penrose (1991, 1994 , 1997 ), the other by Henry Stapp ( 1993 ). Penrose argues that conscious acts of thinking and choosing are tied directly to the processes of resolving quantum superpositions of potentialities to one actuality requiring a kind of large-scale quantum coherence be present in the brain. Stapp's approach focuses on the nonlocal correlations exhibited in the Einstein-Podolsky-Rosen experiments of Bell's inequalities (Einstein, Podolsky, and Rosen 1935 ; Aspect, Dalibard, and Roger 1982a , 1982b ; Bell 1987 , 14–21; Stapp 1993 , 5–9). Stapp ( 1993 , 2005 ) argues that nonlocal holism exhibited by these so-called EPR experiments implies quantum effects play an important role in consciousness and decision making in which QM provides the basis for a Heisenberg-Whiteheadean ontology for reality, which is neither dualist nor physicalist, but, instead, involves a kind of psychophysical interactionism implying an underlying unity. I should point out that Stapp's ( 1993 , 91–92) view is ultimately deterministic in that he construes quantum indeterminism to be due to our ignorance.

Whether these properties are strongly emergent in the sense of footnote 6 or in some weaker sense is usually left open.

Although I am unaware of anyone actively exploring this direction in the literature, there is a proposal related to, and indeed influenced by the Brussels-Austin work, by John Polkinghorne ( 1991 , 34–48) that the randomness in macroscopic chaotic models and systems be interpreted as representing a genuine indeterminism rather than merely a measure of our ignorance. He shares a deep skepticism that the interpretive difficulties of quantum mechanics raised in the section of this essay entitled “Routes for Amplification: Chaos and Nonlinear Dynamics,” will be overcome soon, casting doubt, in his mind, on whether this is the right source for the openness or indeterminism he thinks important to the free will and action we experience (40–41). Polkinghorne argues that the physical world must possess openness (causal creases or joints, if you will) for human free will to operate in the sense we experience (as well as for God to be active in the world; see Polkinghorne 1988 , 1989 ). In essence, the sensitivity to small changes exhibited by the systems and models studied in chaotic dynamics, complexity theory, and nonequilibrium statistical mechanics is taken to represent an opening at the ontic level in the physical order for human choice to cause physical changes (e.g., bodily changes such as beginning to walk). However, the sensitivity upon which Polkinghorne relies would also be open to quantum influences, whether deterministic or indeterministic, according to the sensitive dependence arguments discussed in the section “Routes for Amplification: Chaos and Nonlinear Dynamics,” above. (For a critical discussion of views like Polkinghorne's, see Bishop 2011 , §3.2).

Not the least of which were the pendulum and spring-driven clocks. The most notable dissenter from the view that the mechanical picture implied determinism was Isaac Newton (almost all of his followers essentially equated the mechanical picture with determinism).

Though the earliest articulation of this vision seems to be found in Leibniz ( 1924 , 129).

An excellent example of this is Dennett's 2003 . See Bishop ( 2009 ).

Indeed, a strongly idealized version of the faithful model assumption, the perfect model scenario, is needed, but also runs into difficulties regarding drawing inferences about the systems one is modeling (Judd and Smith 2001 ).

For discussion, see Bishop and Atmanspacher (in this volume) and Bishop ( 2010a ).

For example, Atmanspacher and Rotter ( 2008 ) describe how the attribution of determinism/indeterminism to neural dynamics is contextual and perhaps of great practical value, whereas context-free ontological conclusions about the nature of such dynamics are hazardous at best.

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

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

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

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

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

University of Notre Dame

Notre Dame Philosophical Reviews

  • Home ›
  • Reviews ›

Our Fate: Essays on God and Free Will

Placeholder book cover

John Martin Fischer, Our Fate: Essays on God and Free Will , Oxford University Press, 2016, 243pp., $74.00 (hbk), ISBN 9780199311293.

Reviewed by Martijn Boot, Waseda University

Contemporary debates about free will are dominated by two questions: ‘Is causal determinism true?’ and ‘Is free will compatible with causal determinism?’ These issues are important with respect to the question of to what extent we are morally responsible for our choices and actions. Moral responsibility cannot be detached from some form of free will and self-determination. Parallel to the issue of (in)compatibilism of free will and causal determinism is the question of whether free will and moral responsibility are compatible with divine foreknowledge.

John Martin Fischer investigates the relationship between divine foreknowledge, human freedom and moral responsibility. His book is a collection of eleven previously published essays. Fischer starts with a new introductory article in which he summarizes, extends and applies elements of the analyses presented in the other essays.

Fischer mainly concentrates on the question of whether God’s foreknowledge is compatible with human freedom to do otherwise. But he also pays attention to the still more important question about the compatibility of God’s foreknowledge and human moral responsibility.

Fischer tries to show that the answer to the first question is negative: God’s foreknowledge excludes human freedom to do otherwise. This seems to imply that the answer to the second question is negative as well: God’s foreknowledge is incompatible with human moral responsibility. However, Fischer believes that this doesn’t follow, because, according to him, moral responsibility does not require freedom to do otherwise. I will describe and discuss some of the main arguments Fischer adduces in his book to support these views.

Are divine foreknowledge and freedom to do otherwise reconcilable?

Fischer points out that the arguments for the incompatibility of divine foreknowledge and human freedom to do otherwise are in important ways similar to arguments for the incompatibility of causal determinism and freedom to do otherwise. Both kinds of arguments are related to ideas about the “fixity of the past”. Still there are also relevant differences, which have to be taken into account because the conclusions of the two kinds of arguments with respect to moral responsibility may differ.

Fischer discusses different families of arguments for the incompatibility of God’s foreknowledge and human freedom to do otherwise (‘the Arguments for Incompatibilism’), which are largely based on the premise of ‘the fixity of the past’ and on the so-called Transfer Principle. This principle shows circumstances in which doing X means doing Y , so that if we cannot do Y , we cannot do X . The incompatibilist uses the Transfer Principle by showing that we cannot do X (e.g. we cannot do otherwise) because we cannot do Y (e.g. we cannot bring it about that God held a false belief) . Fischer investigates several challenges to the Arguments for Incompatibilism, such as

  • ‘Scotism’: counterexamples to the Transfer Principle, attributed to Duns Scotus and developed by Anthony Kenny;
  • ‘Ockhamism’: based on ideas of William of Ockham who made a distinction between ‘hard’ (temporally nonrelational) and ‘soft’ (temporally relational) facts about the past;
  • ‘Molinism’: inspired by doctrines of the Jesuit philosopher Luis de Molina about God’s ‘middle knowledge’ — knowledge of what free agents would do in various situations.

Fischer defends the Arguments for Incompatibilism against these challenges. He argues that some versions of these Arguments are vulnerable to some challenges, while these challenges do not affect other versions. For instance, Kenny’s counterexamples to one of the versions of the Transfer Principle do not apply to another version. Besides, the Arguments for Incompatibilism do not exclusively depend on the Transfer Principle. So, one could reject some specific Arguments for Incompatibilism without thereby having to reject others. According to Fischer, Molinism provides an interesting model of divine providence but is not a response to the Arguments for Incompatibilism. He shows that, with respect to the issue of incompatibilism, it begs the question and doesn’t help with the specific problem of reconciling God’s foreknowledge with human freedom.

In addition to the previously published articles, Fischer offers some new reflections on related issues. He defends the theological incompatibilist’s argument against the challenge that it begs the question. Further he tries to show that a rejection of Ockhamism does not depend on the claim that God’s beliefs concern hard instead of soft facts about the past. Finally, he argues (against William Hasker, Patrick Todd and others) that, even in a causally indeterministic world (a world in which events are not causally determined), God can know with certainty that some future event will occur. A causally indeterministic world does not prevent an ordinary human being from having a justified and more or less certain belief in what a particular person, given his personal characteristics, will do in a future choice situation. God’s foreknowledge may be partly conceived in a similar way. But unlike a human being, whose beliefs are fallible, God knows that what He believes is true, because he knows that He is omniscient. God can thus ‘bootstrap’ his first-order belief to a second order of certain knowledge (Fischer calls this approach of God’s foreknowledge the Bootstrapping View). In this perspective God’s foreknowledge may be compatible with a causally indeterministic world. This is important with respect to the question of whether God’s foreknowledge can be reconciled with human moral responsibility. Indeed, if God’s foreknowledge would be inextricably bound up with a causally deterministic world, it would probably exclude human moral responsibility. Still, Fischer recognizes that, also if God’s foreknowledge is unrelated to causal determinism, it may be problematic to reconcile it with human moral responsibility, if it excludes the freedom to do otherwise. This problem forms the second main question to which the book pays attention.

Are God’s foreknowledge and moral responsibility reconcilable?

For a detailed discussion and defense of his theory of moral responsibility Fischer refers to previous publications. Many philosophers believe that the sort of freedom required for moral responsibility requires that the agent could have acted differently (the argument of the Principle of Alternative Possibilities). This would mean that incompatibility of God’s foreknowledge and the freedom to do otherwise implies that God’s foreknowledge is incompatible with moral responsibility. However, Fischer tries to make plausible that incompatibility of God’s foreknowledge and human freedom to do otherwise — a freedom that he calls ‘regulative control’ — does not imply that God’s foreknowledge is irreconcilable with human moral responsibility. Moral responsibility requires another kind of freedom than regulative control, namely ‘guidance control’. An individual exhibits guidance control to the extent that he acts from his own mechanism of practical reasoning and human deliberation, as a response to divergent reasons for alternative choices. An agent’s mechanism becomes his own when he ‘takes responsibility’. Thus guidance control consists of two important components: ownership and reasons-responsiveness .

Fischer adduces the following example — inspired by examples given by Harry Frankfurt — to demonstrate that somebody can be held morally responsible for his choice, although he could not have chosen otherwise.

Black has secretly inserted a chip in Jones’s brain. This enables Black to monitor and control Jones’s activities. If, in the presidential election, Jones were to show any inclination to vote for anyone other than the Democrat candidate, then the chip in Jones’s brain would intervene to ensure that he actually decides to vote for the Democrat. But if Jones decides on his own to vote for the Democrat, the chip does nothing. Suppose that Jones decides to vote for the Democrat on his own, just as he would have if Black had not inserted the chip in his head. It seems, upon first thinking about this case, that Jones can be held morally responsible for his choice, although he could not have done otherwise.

Fischer uses this Frankfurt-like approach to support his semicompatibilist view: he believes that divine foreknowledge is incompatible with free choices between alternative possibilities, but compatible with moral responsibility, because moral responsibility does not require freedom to do otherwise.

Some other philosophers, including Derk Pereboom, also do not believe that moral responsibility requires freedom to do otherwise, but they nevertheless hold that moral responsibility requires that the agent be the ‘ultimate source’ of his behavior. Causal determination is incompatible with the agent being the ultimate source of his behavior. Therefore, even if moral responsibility does not require freedom to do otherwise, causal determination seems incompatible with moral responsibility if the ultimate source requirement is right.

However, according to Fischer, his semicompatibilism is not vulnerable to the challenge of source-incompatibilism because, unlike causal determinism, the conception of divine foreknowledge does not exclude that the agent remains the source of his behavior, insofar as God’s foreknowledge is not conceived as causing human action. Still, Fischer points out that, also in the context of God’s foreknowledge and the Frankfurt Cases, something entirely external to the agent and out of his control is sufficient for his behavior. In this sense the source of his behavior is external to him. If, as Fischer believes, this is still compatible with guidance control — and, therefore, with moral responsibility — then this may mean that likewise the external source of the agent’s behavior in the case of causal determinism need not imply that moral responsibility is excluded. Therefore, Fischer argues, his Frankfurt-like approach casts doubt not only on the requirement of alternative possibilities for moral responsibility, but also on the requirement that, to be morally responsible, the agent must be the ultimate source of his behavior.

In sum, Fischer argues that it is at least plausible that, if God knows everything about our future, it follows that we are never free to do otherwise. But he further argues that it does not follow from this lack of regulative control that we lack guidance control of our future. If guidance control is the freedom-relevant condition for moral responsibility, then we can be morally responsible for our behavior, even if God knows everything we will ever do in advance.

Robert Kane notices in the Introduction of The Oxford Handbook of Free Will that contemporary debates about free will in the light of God’s foreknowledge “surpassed even medieval discussions in labyrinthine complexity.” Also Fischer’s profound analyses are complicated and not easily to follow for readers who are not specialized in the relevant issues. Therefore, the book is relevant and important mainly for specialists. Fischer’s argument is interesting not only for specialists who are interested in the (in)compatibility of God’s foreknowledge , the freedom to choose otherwise and moral responsibility but also for specialists who are interested in the (in)compatibility of causal determinism , the freedom to choose otherwise and moral responsibility. The reason is that the two issues have much ground in common, while there are also relevant differences, which Fischer elucidates.

Fischer’s approach consists of subtle conceptual analyses and rigorous evaluations of premises and the logical validity of arguments adduced in the free will debate. For instance, he discusses and defends the premise of the fixity of the past against challenges; he shows that the distinction between hard and soft facts may both reveal and conceal something, and that this distinction should be distinguished from the distinction between fixed and nonfixed facts; he points at confusions about the meaning of concepts such as various senses of ‘can’; he reveals logical fallacies such as equivocation and begging-the-question; he uncovers the fallacy of moving from one ‘language game’ to another (warned against by Wittgenstein).

Fischer recognizes that Frankfurt-type examples (which he adduces to make plausible that moral responsibility does not depend on freedom to do otherwise) are contentious and that he offers only a sketch of his theory of moral responsibility without a thorough defense. This may be one of the reasons why Fischer’s distinction between regulative and guidance control does not take away the doubt whether guidance control — in the absence of the possibility to do otherwise — is sufficient to make the agent morally responsible for his actions. The doubt especially applies to his suggestion that guidance control may be sufficient for moral responsibility even if causal determinism is true (causal determinism obtains if the initial state of the world together with the laws of nature entails every truth about what happens in the future).

Suppose Peter murders John, while he has guidance control (he ‘acts from his own suitably reasons-responsive mechanism’) but not regulative control. Not having regulative control means that Peter is not free to do otherwise and, thus, that it is impossible that Peter does not murder John. It is true that Peter acts from his own reasons-responsive mechanism and that he deliberately murders John. However, if causal determinism is true, then not only Peter’s actions but also his guidance control (and ‘every truth’ related to it, for instance, whether and how and with what results he exercises this control) are entirely determined by factors outside of him. The conditions sufficient for the necessary outcome of his deliberation and choice were already in place long before he even existed. As Fischer recognizes, the fixity of the past means that “the past is like the dog’s tail, and it is the tail that wags the dog”, not the other way round. Therefore, guidance control seems insufficient for rescuing moral responsibility, if the latter requires at least a minimum of self-determination. It seems inappropriate to blame somebody and hold him morally responsible for something entirely caused by external factors, which he, despite his guidance control, could not possibly influence, change or remove.

It is true that the unavoidability with respect to causal determinism differs from the unavoidability with respect to divine foreknowledge, insofar as God’s foreknowledge is not conceptualized as bringing about human action. However, as Fischer points out, it is questionable whether this difference makes a difference . Besides, Fischer admits that his ‘bootstrapping’ argument (which detaches God’s foreknowledge from causal determinism) is controversial. If we take into account these uncertainties, the conclusion seems justified that, if we are not free to do otherwise, it is implausible that divine foreknowledge and causal determinism are compatible with human moral responsibility.

Kane refers to Milton’s Paradise Lost , in which the angels, debating their freedom in the light of God’s foreknowledge, were lost in “endless mazes”. As he notices, this is “not a comforting thought for us mortals.” Fischer does not pose the question whether human beings are sufficiently capable of understanding the relationship between an omniscient God and human freedom, but he recognizes that every major view about God’s knowledge of the future “has at least a mystery associated with it, if not a significant problem.” This may mean that the problems under consideration are irresolvable, even in principle, due to a fundamental incompleteness of human knowledge with respect to the relation between God and man.

January 1, 2020

My Go-To Arguments for Free Will

Free will must exist if some of us have more of it than others

By John Horgan

thesis about free will

You chose to read this column, didn’t you? That’s proof of free will.

Yuri Arcurs Getty Images

This article was published in Scientific American’s former blog network and reflects the views of the author, not necessarily those of Scientific American

I often dwell on free will as the new year begins, and my resolutions ( quit caffeine, again , start  meditating, again ) are already wavering. Plus, late last semester I found myself trying, again, to convince my students to believe in free will. For these reasons, and just because I feel like it, I decided to jot down a few arguments for free will. 

By free will, I mean a capacity for deliberate, conscious decisions. Choices. Free will is variable. The more choices you have, the more free will you have. Our choices are constrained by all sorts of factors, physical, biological, social, economic, political, even romantic. My choices, for example, are often overruled by those of my willful girlfriend “Emily,” but that’s okay, because I’m with her by choice. Choices are never  entirely  free, but that doesn’t mean we lack them.

I’m not going to invoke quantum mechanics, information theory or arcane philosophical reasoning. I find slick, technical defenses of free will almost as unpersuasive as slick, technical denials. My arguments will leave many questions unanswered. Did we discover free will or invent it? I don’t know, both, perhaps. Do non-human animals possess it? Maybe, maybe not, but I know  we  have it. All right, enough throat-clearing, here are my arguments:

On supporting science journalism

If you're enjoying this article, consider supporting our award-winning journalism by subscribing . By purchasing a subscription you are helping to ensure the future of impactful stories about the discoveries and ideas shaping our world today.

Just Because Physics Can’t Account for Free Will Doesn’t Mean It Doesn’t Exist.  Free-will deniers tend to be  hard-core materialists , who think reality, ultimately, consists of particles pushed and pulled by fundamental forces. This hyper-reductive worldview can’t account for choice. Or  consciousness, for that matter , or  beauty ,  morality ,  meaning  and other trappings of the human condition. That doesn’t mean these things are somehow illusory. It just means materialistic science, which does a splendid job explaining protons and planets,  remains baffled by us .

Don’t let  mean reductionists  bully you into agreeing with them. They’re  not as smart as they think they are . In fact, anyone who argues strenuously against free will is a walking, talking contradiction. Their disbelief, like my belief, is a choice stemming from (in their case, faulty) reason, another mind-based capacity irreducible to physics and chemistry.

This Sentence Is Proof of Free Will.  And this one. And this one. This whole column constitutes proof that free will exists. Free will is an idea, a packet of meaning, that cannot be reduced to mere physics. The  idea  of free will, not its instantiation in my brain, provoked me, my physical self, to type this column. I’m not  compelled  to write it. I have lots of other things to do, like watch  Couples Therapy  on Showtime with Emily, or find a birthday present for her. (No matter what I choose, she probably won’t like it. She’s very choosy.)

No, I  choose  to write this column, because I want others to share my belief in free will. It matters to me. Once I decide to write the column, then I must decide how to write it. That process entails countless choices. They are constrained, limited, by factors such as time, my verbal skills and knowledge, my sense of what readers will like and so on. Like I said, just because free will is never  entirely  free doesn’t mean it doesn’t exist.

You Reading This Sentence Is Proof Too.  You don’t have to read this column, do you? Of course not. You choose to read it, freely. More proof of free will! If you’re irritated by the substance or style of this column, and you jump to Twitter to find something more amusing, that’s another choice! More proof!

Libet’s Experiments Are Bogus.  Decades ago, psychologist Benjamin Libet monitored subjects’ neural activity while they chose to hit a button, and he discovered a burst of activity preceding the conscious decision to push the button by a split second. Free-will deniers seized upon Libet’s experiments as evidence that our  brains  make decisions, and our conscious choices are mere afterthoughts. Hence, no free will.

First of all, deciding when to push a button is not remotely analogous to genuine choices, like whether to get married, have kids, get divorced. The Libet experiments are profoundly flawed, as psychologist Steve Taylor points out in a  recent  Scientific American  column . The question is, why did anyone ever take seriously the claim that Libet had disproved free will? Why do smart people accept such flimsy evidence? I honestly don’t get it, unless it’s that some intellectuals enjoy attacking beliefs that give others comfort. Many adamant free-will deniers are also adamant atheists.

Free Will Must Exist If Some People Have More of It Than Others.  You have more free will—more ability to see, weigh and make choices--now than when you were a baby. Right? You have more than if you were suffering from Alzheimer’s disease, or addicted to heroin, or imprisoned for manslaughter. If you are black or female, gay or transgender, and living in a democracy,  you have more choices and hence free will  than you would have 50, 100 or 200 years ago. If some people and societies have more choices than others—and they obviously do--free will must exist.

This is my best, slam-dunk argument for free will. That doesn’t mean it always or even usually works. My students can be so stubborn! But I feel good making this argument, it convinces me. Usually. To be honest, I have  doubts about free will now and then. Sometimes I feel like I’m sleepwalking through life. I’m a confabulating somnambulist, a bundle of reflexes, twitches and compulsions with no self-knowledge, let alone self-control. I’m not even sure I really choose to be with Emily!

In these dark times, I give myself pro-free-will pep talks, like this column. Free will is like Tinkerbell. If you don’t believe in her, she dies. Maybe that’s what I’m really doing with this column, trying to keep free will alive. Please help me. The more we believe in free will, the more we have, and the more likely we are to change the world for the better.

Further Reading :

Mind-Body Problems   (free online book, also available as  Kindle e-book  and  paperback ). For reflections on free will, see especially chapters on  Douglas Hofstadter ,  Owen Flanagan  and  Deirdre McCloskey .

Meta-Post: Posts on the Mind-Body Problem  (includes previous columns on free will)

Free Will Is Real

Is Scientific Materialism "Almost Certainly False"?

  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.

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

Arguments for Incompatibilism

We believe that we have free will and this belief is so firmly entrenched in our daily lives that it is almost impossible to take seriously the thought that it might be mistaken. We deliberate and make choices, for instance, and in so doing we assume that there is more than one choice we can make, more than one action we are able to perform. When we look back and regret a foolish choice, or blame ourselves for not doing something we should have done, we assume that we could have chosen and done otherwise. When we look forward and make plans for the future, we assume that we have at least some control over our actions and the course of our lives; we think it is at least sometimes up to us what we choose and try to do.

Determinism is a highly general claim about the universe: very roughly, that everything that happens, including everything you choose and do, is determined by facts about the past together with the laws. Determinism isn’t part of common sense, and it is not easy to take seriously the thought that it might, for all we know, be true. The incompatibilist believes that if determinism turned out to be true, our belief that we have free will would be false. The compatibilist denies that the truth of determinism would have this drastic consequence. According to the compatibilist, the truth of determinism is compatible with the truth of our belief that we have free will. The philosophical problem of free will and determinism is the problem of deciding who is right: the compatibilist or the incompatibilist.

Much of the philosophical interest in the free will/determinism problem is motivated by concerns about moral responsibility because, it is generally agreed, having free will is a necessary condition of being morally responsible. So if determinism precludes free will, it also precludes moral responsibility. But it’s important to distinguish questions about free will (whether we have it, what it amounts to, whether it is compatible with determinism, whether it is compatible with other things we believe true) from questions about moral responsibility. Someone might believe that we have free will and that free will is compatible with determinism while also believing, for other reasons, that no one is ever morally responsible. And someone might believe that we don’t have free will (because of determinism or something else) while also believing, against conventional wisdom, that we are nevertheless morally responsible. What one believes about determinism and moral responsibility will depend, in large part, on what one believes about various matters within the scope of ethics rather than metaphysics. Among other things, it will depend on what one takes moral responsibility to be (P. Strawson 1962; G. Strawson 1986, 1994; Scanlon 2008; Watson 1996, 2004; Wolf 1990). For these reasons it is important not to conflate the question of the compatibility of free will and determinism with the question of whether moral responsibility is compatible with determinism.

It must be acknowledged that a change in definitions has crept into the literature, and many contemporary theorists understand ‘compatibilism’ and ‘incompatibilism’ as claims about moral responsibility (or “moral freedom” or the freedom that “grounds” or “explains” “moral responsibility”) rather than claims about free will (Pereboom 1995, 2001, 2014. See Vihvelin 2011 for discussion). This entry will not follow that usage.

This encyclopedia entry is about the free will/determinism problem; more specifically, it is about arguments for the claim that free will is incompatible with determinism. Since so much of the contemporary free will literature is dominated by concerns about moral responsibility, some of the arguments considered will be arguments for the thesis that if determinism is true, we are never morally responsible because we never satisfy the “freedom requirement” for being morally responsible for our actions. But the focus, in this entry, will be on the question of whether free will (or acting with free will) is compatible with determinism.

1. Preliminaries

2. two reasons for thinking that free will is incompatible with determinism, 3.1 no forking paths argument.

  • 3.2 Manipulation Arguments

4. Sourcehood Arguments

5. choice and the consequence argument, other internet resources, related entries.

In the literature, “determinism” is sometimes used as an umbrella term for a variety of different claims which have traditionally been regarded as threats to free will. Given this usage, the thesis that we are calling “determinism” (nomological determinism, also sometimes called ‘causal determinism’) is just one of several different kinds of determinism, and the free will/determinism problem we will be discussing is one of a family of related problems. For instance, logical determinism is the thesis that the principle of bivalence holds for all propositions, including propositions about the future, and the problem of free will and logical determinism is the problem of deciding whether our belief that we have free will is compatible with the existence of truths about all our future actions. (It should be noted that on this point there is almost universal agreement among philosophers that the answer is “yes”.) For more on logical determinism, see the entry on fatalism ; Taylor 1962; D. Lewis 1976; Merricks 2009; Fischer & Todd 2011; Van Cleve 2019; Vihvelin 2013 and 2020. Theological determinism is the thesis that God exists and has infallible knowledge of all true propositions including propositions about our future actions; the problem of free will and theological determinism is the problem of understanding how, if at all, we can have free will if God (who cannot be mistaken) knows what we are going to do. For more on theological determinism, see the entries on fatalism and divine foreknowledge and Fischer & Todd 2015.

In this entry, we will be restricting our attention to arguments for the incompatibility of free will and nomological determinism, but it is important to understand one preliminary point. Nomological and logical determinism are very different kinds of claims. Logical determinism is a claim about truth ; nomological determinism is a claim about the natural laws . Logical determinism doesn’t say anything about causation or the laws; it simply says that, regardless of whether we ever come to know them or not, there are timelessly true propositions about the past, present, and future, including our future actions. Logical determinism doesn’t entail nomological determinism; it might be true even if nomological determinism is false. But nomological determinism says (roughly) that facts about the past together with facts about the laws determine all the facts about the future. So if nomological determinism is true, there are true propositions about all our future actions.

Why does this matter? For the following reason. A common first response to logical determinism is the fatalist’s response:

If there are truths about what I will do in the future, then I must do whatever I will do, and so I have no free will.

While this response has a powerful intuitive grip (as can be seen from much of the popular and even scientific discussion of time travel), it is generally agreed, by philosophers, that the fatalist is making a mistake. Different diagnoses have been given of the “fatalist fallacy” or other mistake in modal reasoning, but the basic point is simple. Truth is not the same as necessity; it isn’t the same as logical necessity, metaphysical necessity, or even the relative necessity of unavoidability or lack of ability or power. The existence of a detailed set of truths about my future actions is consistent with my ability to do things other than the things I actually do.

It is of course possible to agree that the existence of truths about all our future actions is compatible with free will while denying that the existence of nomologically determined truths about all our future actions is compatible with free will. But an argument is needed for this conclusion, an argument which doesn’t rely on fatalist reasoning or an appeal to fatalist intuitions.

For comparisons between arguments for incompatibilism and arguments for fatalism, see van Inwagen 1983, Mackie 2003, Perry 2004, and Vihvelin 2008 and 2013.

At a first approximation, nomological determinism (henceforth “determinism”), is a contingent and empirical claim about the laws of nature: that they are deterministic rather than probabilistic, and that they are all-encompassing rather than limited in scope. At a second approximation, laws are deterministic if they entail exceptionless regularities (e.g., that all \(F\)s are \(G\)s, that all \(ABCD\)s are \(E\)s) and laws are probabilistic if they say that \(F\)s have an objective chance \(N\) (less than 1) of being \(G\)s (note that so-called “statistical laws” need not be probabilistic laws; see Armstrong 1983, Loewer 1996a). The laws of nature are all-encompassing if deterministic or probabilistic laws apply to everything in the universe, without any exceptions. If, on the other hand, some individuals or some parts of some individuals (e.g., the nonphysical minds of human beings) or some of the behaviors of some of the individuals (e.g., the free actions of human beings) do not fall under either deterministic or probabilistic laws, then the laws are limited rather than all-encompassing.

For a more precise articulation of determinism, the contemporary literature offers us two main choices.

Determinism is standardly defined in terms of entailment, along these lines: A complete description of the state of the world at any time together with a complete specification of the laws entails a complete description of the state of the world at any other time. (Hoefer 2002, Mele 2009, Beebee 2013; see also van Inwagen 1983, Ginet 1990, and the Encyclopedia entry on causal determinism ).

Alternatively, following D. Lewis 1973, we might understand determinism as the thesis that our world is governed by a set of natural laws which is such that any two possible worlds with our laws which are exactly alike at any time are also exactly alike at every other time (see also Earman 1986). This second definition of determinism is stronger than the first; if a possible world is deterministic according to the Lewis/Earman definition, it is deterministic according to the Entailment definition, but not vice versa.

Let’s call a possible world ‘deterministic’ iff the thesis of determinism is true at that world; ‘non-deterministic’ iff the thesis of determinism is false at that world. There are two very different ways in which a world might be non-deterministic. A world might be non-deterministic because at least some of its fundamental laws are probabilistic, or a world might be non-deterministic because it has no laws or because its laws are not all-encompassing. Let’s call worlds which are non-deterministic in only the first way ‘probabilistic worlds’ and let’s call worlds which are non-deterministic in the second way ‘lawless’ or ‘partly lawless’ worlds.

Determinism is a thesis about the statements or propositions that are the laws of our world; it says nothing about whether these statements or propositions are knowable by finite beings, let alone whether they could, even in principle, be used to predict all future events. (For more on the relation between determinism and predictability, see the Encyclopedia entry on Causal Determinism .)

Determinism, understood according to either of the two definitions above, is neutral with respect to different philosophical accounts of lawhood, ranging from the so-called “naïve regularity” account (Swartz 1986) to broadly Humean or “best system” accounts (D. Lewis 1973; Earman 1986; Loewer 1996a; Beebee 2000; Schaffer 2008) to various kinds of necessitarian accounts (Armstrong 1983; Carroll 2008). In the book that set the stage for much of the contemporary discussion of the free will/determinism problem, van Inwagen 1983 appealed only to two relatively uncontroversial assumptions about the laws of nature: that the propositions that are laws have this status independent of the history or present state of scientific knowledge and that the laws impose limits on our abilities. Most of the participants in the current debate agree, but there are dissenters who think that determinism must be defined in a metaphysically stronger way, as a kind of necessitation relation between particular events (Mumford & Anjum 2014). For critique, see Mackie 2014a and Franklin 2014.

Determinism (understood according to either of the two definitions above) is not a thesis about causation; it is not the thesis that causation is always a relation between events, and it is not the thesis that every event has a cause. If the fundamental laws turn out to be probabilistic rather than deterministic, this doesn’t mean that there is no causation; it just means that we have to revise our theories of causation to fit the facts. And this is what philosophers of causation have done; there are probabilistic versions of lawful entailment theories of causation, of counterfactual theories of causation, and so on, for all major theories of causation (see the entries on the metaphysics of causation and counterfactual theories of causation ). It is now generally accepted that it might be true that every event has a cause even if determinism is false and thus some events lack sufficient causes.

More controversially, it might be true that every event has a cause even if our world is neither deterministic nor probabilistic. If there can be causes without laws (if a particular event, object, or person can be a cause, for instance, without instantiating a law), then it might be true, even at a lawless or partly lawless world, that every event has a cause (Anscombe 1981; van Inwagen 1983).

It’s less clear whether determinism entails the thesis that every event has a cause. Whether it does depends on what the correct theory of causation is; in particular, it depends on what the correct theory says about the relation between causation and law.

What is clear, however, is that we should not make the assumption, almost universally made in the older literature, that the thesis that every event has a cause is equivalent to the thesis of determinism. This is an important point, because some of the older arguments in the literature against incompatibilism assume that the two claims are equivalent (Hobart 1934).

In the older literature, it was assumed that determinism is the working hypothesis of science, and that to reject determinism is to be against science. This no longer seems plausible. Some people think that quantum physics has shown determinism to be false. This remains controversial (Albert 1992; Loewer 1996b; P. Lewis 2016), but it is now generally agreed that we can reject determinism without accepting the view that the behavior of human beings falls outside the scope of natural laws. If naturalism is the thesis that human behavior can be explained in the same kind of way—in terms of events, natural processes, and laws of nature—as everything else in the universe, then we can reject determinism without rejecting naturalism.

Note, finally, that determinism neither entails physicalism nor is entailed by it. There are possible worlds where determinism is true and physicalism false; e.g., worlds where minds are nonphysical things which nevertheless obey deterministic laws (van Inwagen 1998). And there are possible worlds (perhaps our own) where physicalism is true and determinism is false.

So much for determinism. What about free will? How should we understand the disagreement between the compatibilist and the incompatibilist?

There’s lots of room for argument about how, exactly, we should understand our commonsense beliefs about ourselves as persons with free will. (Are we born with free will? If not, when do we acquire it, and in virtue of what abilities or powers do we have it? What is the difference between acting intentionally and acting with free will?) Luckily we don’t have to answer these questions in order to say what is at issue between the compatibilist and the incompatibilist. Let’s define the free will thesis as the thesis that at least one non-godlike (human-like) creature has free will, leaving it open what that amounts to. The free will thesis is a minimal claim about free will; it would be true if one person in the universe acted with free will (acted freely, acted while possessing free will) on one occasion. We won’t assume that the free will thesis is true or even possibly true, but let a free will world be any possible world where the free will thesis is true. Since non-determinism is the negation of determinism, and since determinism is a contingent thesis, we can divide the set of possible worlds into two non-overlapping subsets: deterministic worlds and non-deterministic worlds.

Given this apparatus, we could define incompatibilism and compatibilism in the following way: incompatibilism is the thesis that no deterministic world is a free will world . (Equivalently, incompatibilism is the claim that necessarily, if determinism is true, then the free will thesis is false.) And we could define compatibilism as the denial of incompatibilism; that is, as the claim that some deterministic worlds are free will worlds. (Equivalently, compatibilism is the claim that possibly, determinism and the free will thesis are both true.)

This way of defining compatibilism is unproblematic. There are compatibilists who are agnostic about the truth or falsity of determinism, so a compatibilist need not be a soft determinist (someone who believes that it is in fact the case that determinism is true and we have free will). And a compatibilist might believe that we don’t have free will for reasons independent of determinism. But all compatibilists believe that it is at least possible that determinism is true and we have free will. So all compatibilists are committed to the claim that there are deterministic worlds that are free will worlds.

But this definition of incompatibilism has a surprising consequence. Suppose, as some philosophers have argued, that we lack free will because free will is conceptually or metaphysically impossible, at least for nongodlike creatures like us (Taylor 1962; G. Strawson 1986, 1994). If these philosophers are right, there are no free will worlds. And if there are no free will worlds, it follows that there are no deterministic free will worlds. So if free will is conceptually or metaphysically impossible, at least for creatures like us, it follows that incompatibilism (as we have just defined it) is true. But this doesn’t seem right. If it is conceptually or metaphysically impossible for us to have free will, then we lack free will regardless of whether determinism is true or false. And if that is so, then the incompatibilist cannot say the kind of things she has traditionally wanted to say: that the truth or falsity of determinism is relevant to the question of whether or not we have free will, that if determinism were true, then we would lack free will because determinism is true, and so on.

If we want to avoid this counter-intuitive result, there is a remedy. Instead of understanding compatibilism and incompatibilism as propositions that are contradictories, we can understand them as propositions that are contraries. That is, we can understand compatibilism and incompatibilism as claims that can’t both be true, but that can both be false. Compatibilism and incompatibilism are both false if a third claim, impossibilism, is true. Impossibilism is the thesis that free will is conceptually or metaphysically impossible for non-godlike creatures like us.

If we accept this three-fold classification, we can define our terms as follows: Impossibilism is the thesis that there are no free will worlds. Incompatibilism is the thesis that there are free will worlds but no deterministic world is a free will world. Compatibilism is the thesis that there are free will worlds and free will worlds include deterministic worlds. (For some objections to this three-fold classification see McKenna 2010 and Mickelson 2015a. For defense, see Vihvelin 2008 and 2013.)

The term ‘impossibilism’ is being coined; however, the position it describes is recognized in the literature under a variety of names: the “no free will either way” view, “non-realism”, “illusionism”, “pessimism”. Theorists who defend impossibilism include G. Strawson 1986 and 1994, and Smilansky 2000. Another kind of impossibilist is the fatalist (Taylor 1962).

In the older literature, there were just two kinds of incompatibilists—hard determinists and libertarians. A hard determinist is an incompatibilist who believes that determinism is in fact true (or, perhaps, that it is close enough to being true so far as we are concerned, in the ways relevant to free will) and because of this we lack free will (Holbach 1770; Wegner 2003). A libertarian is an incompatibilist who believes that we in fact have free will and this entails that determinism is false, in the right kind of way (van Inwagen 1983). Traditionally, libertarians have believed that “the right kind of way” requires that agents have a special and mysterious causal power not had by anything else in nature: a godlike power to be an uncaused cause of changes in the world (Chisholm 1964). Libertarians who hold this view are committed, it seems, to the claim that free will is possible only at worlds that are at least partly lawless, and that our world is such a world (but see O’Connor 2000, Clarke 2003 and Steward 2012). But in the contemporary literature there are incompatibilists who avoid such risky metaphysical claims by arguing that free will is possible at worlds where some of our actions have indeterministic event causes (Kane 1996, 1999, 2008, 2011a; Ekstrom 2000; Balaguer 2010; Franklin 2018) or that free will is possible at worlds where some of our actions are uncaused (Ginet 1990). Note that none of these three kinds of incompatibilists (agent-causation theorists, indeterministic event-causation theorists, non-causal theorists) need be libertarians. They may reserve judgment about the truth or falsity of determinism and therefore reserve judgment about whether or not we in fact have free will. They might also be hard determinists because they believe that determinism is in fact true. But what they do believe—what makes them incompatibilists—is that it is possible for us to have free will and that our having free will depends on a contingent fact about the laws that govern the universe: that they are indeterministic in the right kind of way. (See the entry on incompatibilist theories of free will ).

Given these definitions and distinctions, we can now take the first step towards clarifying the disagreement between compatibilists and incompatibilists. Both sides agree that it is conceptually and metaphysically possible for us to have free will; their disagreement is about whether any of the possible worlds where we have free will are deterministic worlds. The compatibilist says ‘yes’; the incompatibilist says ‘no’. Arguments for incompatibilism must, then, be arguments for the claim that necessarily, if determinism is true, we lack the free will we might otherwise have.

A common first response to determinism is to think that it means that our choices make no difference to anything that happens because earlier causes have pre-determined or “fixed” our entire future (Nahmias 2011). It is easy to think that determinism implies that we have a destiny or fate that we cannot avoid, no matter what we choose or decide and no matter how hard we try.

Man, when running over, frequently without his own knowledge, frequently in spite of himself, the route which nature has marked out for him, resembles a swimmer who is obliged to follow the current that carries him along; he believes himself a free agent because he sometimes consents, sometimes does not consent, to glide with the stream, which, notwithstanding, always hurries him forward. (Holbach 1770 [2002]: 181; see also Wegner 2003)

It is widely agreed, by incompatibilists as well as compatibilists, that this is a mistake. Empirical discoveries about our brain and behavior might tell us that we don’t have as much conscious control as we think we have (Wegner 2003; Libet 1999). (For critique of arguments claiming that recent scientific research has shown that “conscious will is an illusion”, see Mele 2009, some of the essays in Sinnott-Armstrong & Nadel 2011, Mele 2015, and Roskies & Nahmias 2016.) And there are worries, arising from certain versions of physicalism, that our mental states don’t have the causal powers we think they have (Kim 1998). But these threats to free will have nothing to do with determinism. Determinism might imply that our choices and efforts have earlier sufficient causes; it does not imply that we don’t make choices or that our choices and efforts are causally impotent. Determinism is consistent with the fact that our deliberation, choices and efforts are part of the causal process whereby our bodies move and cause further effects in the world. And a cause is the kind of thing that “makes a difference” (Sartorio 2005, Menzies 2017, List 2019). If I raise my hand because I chose to do so, then it’s true, ceteris paribus , that if my choice had not occurred, my hand-raising would not have occurred.

Putting aside this worry, we may classify arguments for incompatibilism as falling into one of two main varieties:

  • Arguments for the claim that determinism would make it impossible for us to cause and control our actions in the right kind of way.
  • Arguments for the claim that determinism would deprive us of the power or ability to do or choose otherwise .

Arguments of the first kind focus on the notions of self, causation, and responsibility; the worry is that determinism rules out the kind of causation that we invoke when we attribute actions to persons (“It was Suzy who broke the vase”) and make judgments of moral responsibility (“It wasn’t her fault; Billy pushed her”). Someone who argues for incompatibilism in this way may concede that the truth of determinism is consistent with the causal efficacy of our deliberation, choices, and attempts to act. But, she insists, determinism implies that the only sense in which we are responsible for what we do is the sense in which a dog or young child is responsible. Moral responsibility requires something more than this, she believes. Moral responsibility requires autonomy or self-determination: that our actions are caused and controlled by, and only by , our selves. To use a slogan popular in the literature: We act freely and are morally responsible only if we are the ultimate source of our actions.

Each of us, when we act, is a prime mover unmoved. In doing what we do, we cause certain events to happen, and nothing—or no one—causes us to cause these events to happen. (Chisholm 1964: 32) Free will…is the power of agents to be the ultimate creators or originators and sustainers of their own ends or purposes…when we trace the causal or explanatory chains of action back to their sources in the purposes of free agents, these causal chains must come to an end or terminate in the willings (choices, decisions, or efforts) of the agents, which cause or bring about their purposes. (Kane 1996: 4)

Arguments of the second kind focus on the notion of choice. To have a choice, it seems, is to have genuine options or alternatives—different ways in which we can act. The worry is that determinism entails that what we do is, always, the only thing we can do, and that because of this we never really have a choice about anything , as opposed to being under the (perhaps inescapable) illusion that we have a choice. Someone who argues for incompatibilism in this way may concede that the truth of determinism is consistent with our making choices, at least in the sense in which a dog or young child makes choices, and consistent also with our choices being causally effective. But, she insists, this is not enough for free will; we have free will only if we have a genuine choice about what actions we perform, and we have a genuine choice only if there is more than one action we are able to perform.

A person has free will if he is often in positions like these: he must now speak or be silent, and he can now speak and can now remain silent; he must attempt to rescue a drowning child or else go for help, and he is able to attempt to rescue the child and able to go for help; he must now resign his chairmanship or else lie to the members; and he has it within his power to resign and he has it within his power to lie. (van Inwagen 1983: 8)

van Inwagen is giving what he takes to be uncontroversial examples of persons who have free will; he isn’t saying that free will just is the ability to perform more than one overt action. Our choices include choices among purely mental actions (to pay attention to a lecture or to spend the time deciding what to cook for dinner) as well as choices about the actions we perform by moving our bodies. If the incompatibilist claims that determinism robs us of free will by robbing us of choice, this must be understood as the claim that determinism has the consequence that we are never able to do anything other than what we actually do, where “do” includes deciding, choosing, and other instances of mental agency as well as the things we do by moving our bodies.

We might question whether arguments based on self-determination and arguments based on choice are independent ways of arguing for incompatibilism for the following reason: I cause and control my actions in the self-determining way required for moral responsibility only if my actions are the product of my free will and my actions are the product of my free will only if I have the ability to do (choose to do, decide to do, intend to do, try to do) otherwise. If determinism has the consequence that I never have the ability to do otherwise, it also has the consequence that I never cause my actions in the self-determining way required for moral responsibility (Kane 1996).

At one time, this link between moral responsibility, self-determination, and the ability to do otherwise was common ground between compatibilists and incompatibilists. That is, everyone agreed that a person is morally responsible only if she has the right kind of control over what she does, and everyone assumed that a person has the right kind of control over something she does only if she is able to do (or at least decide, choose, intend, or try) otherwise . Given this assumption, anyone hoping to defend the claim that moral responsibility is compatible with determinism had to first show that the ability to do otherwise is compatible with determinism. During the heyday of ordinary language philosophy in the middle years of the last century, there was a further assumption: that the ability to do otherwise is compatible with determinism only if sentences like “\(S\) could have done \(X\)” (where \(X\) is something \(S\) did not do) were equivalent in meaning to counterfactual conditionals like “if \(S\) had chosen (or tried, wanted, preferred, etc.) to do \(X\), \(S\) would have done \(X\)”. This debate was crippled by the fact that it took place at a time when counterfactuals were still poorly understood, before the advent of the Lewis-Stalnaker possible worlds semantics (D. Lewis 1973). There were counterexamples to the analyses proposed, and there was a growing consensus that the prospects for a successful “Conditional Analysis” were dim (Austin 1956; Chisholm 1964; Lehrer 1968 and 1976). (For an argument that this pessimism was premature, see Vihvelin 2004 and 2013. For an argument that a compatibilist doesn’t need to defend a Conditional Analysis, see Lehrer 1976.)

It was against this background that Harry Frankfurt proposed his famous counterexample to the “Principle of Alternate Possibilities” (a person is morally responsible for what he has done only if he could have done otherwise). Frankfurt wanted to defend the claim that moral responsibility is compatible with determinism without having to defend the claim that the ability to do otherwise is compatible with determinism. His strategy took the form of an ingenious thought experiment that was supposed to show that no matter how you understand ability to do otherwise—whether you are a compatibilist or an incompatibilist—you should agree that the possession of this ability is not a necessary condition of being morally responsible (Frankfurt 1969).

There were two steps to the thought experiment. In the first step he invited you to imagine a person, Jones, who has free will, and who acts freely and who satisfies all the conditions you think necessary and sufficient for moral responsibility. You may imagine Jones in one of the scenarios van Inwagen describes, faced with a choice to speak or be silent, to try to rescue the child or go for help, to resign his chairmanship or to lie, and to imagine that Jones deliberates and decides, for his own reasons, in favor of one of his contemplated alternatives, and then successfully acts on his decision. In the second step you are invited to add to the story the existence of a powerful being, Black, who takes a great interest in what Jones does, including how he deliberates and decides. You may fill in the details however you like, but you must imagine that Black has the power to interfere with Jones in a way that ensures that Jones does exactly what Black wants him to do. And you must also imagine that Black is paying very close attention and is prepared to intervene instantly, should it be necessary, to stop Jones from doing something Black doesn’t want him to do. But, as it turns out, it wasn’t necessary. By lucky co-incidence, Jones did exactly what Black wanted him to do. (He even deliberated and decided the way Black wanted him to deliberate and decide.) So Black remained on the sidelines and only watched.

Because Black never laid a finger on Jones, or interfered in any way, it seems that Jones is as morally responsible in the second step of the story as he is in the first step. But it also seems that the existence of the powerful Black suffices to make it the case that Jones is unable to do otherwise (or, as Frankfurt put it, that he lacks “alternate possibilities”.)

Frankfurt’s story has strong intuitive force, and many were convinced that it shows that we were mistaken when we thought that a person has the right kind of control over what she does—the kind of control that is necessary for moral responsibility—only if she is able to do otherwise. These philosophers, who later became known as “Source” compatibilists or “semi-compatibilists”, took the moral to be that it is not relevant, so far as moral responsibility is concerned , whether determinism has the consequence that we don’t have the free will we think we have, that we never have any genuine choices, that we are never able to do otherwise. They said that what matters, so far as moral responsibility is concerned, is only what happens in the “actual sequence” (Fischer 1994; Fischer & Ravizza 1998; Sartorio 2016).

But a little reflection should make us wonder whether facts about the “actual sequence” can come apart from facts about free will in the way envisaged by Frankfurt. How can it be true that Black succeeds, not just in limiting the range of Jones’ alternative courses of action, but in making it the case that Jones has no alternatives? How can Black, sitting on the sidelines, deprive Jones of the ability to deliberate, decide, or try otherwise?

We can easily grant that Black has the power to directly manipulate Jones’ body so Jones loses the ability to move his body in any way other than the ways that Black wants. We might even grant that Black has the power to directly manipulate Jones’ mind/brain, so Jones loses the ability to “move” his mind in any way other than the ways that Black wants. But Black never exercises his power. There is a difference between the existence of a power and the exercise of a power. It is a modal fallacy to reason from “Black has the power to bring it about that Jones lacks the ability to do otherwise” to “Jones lacks the ability to do otherwise”. The truth about Jones is not that Black robs him of the ability to do otherwise; it is the more complicated truth that Black puts him at constant risk of losing the ability to do otherwise. Jones’ free will and his moral responsibility are dependent on the luck of his deliberations and intentions co-coinciding with those of Black (Vihvelin 2013; for some other criticisms of Frankfurt’s argument see Lamb 1993, Alvarez 2009, and Steward 2009).

Frankfurt’s argument has been enormously influential, though perhaps not in the way he intended. His thought experiment was a failure; while most compatibilists were convinced, most incompatibilists were not. (Compatibilists who were not convinced include Smith 1997, 2004; Campbell 2005; Fara 2008; Vihvelin 2000 and 2013.) These incompatibilists insisted, though not for the reason given above, that Black does not succeed in robbing Jones of all his freedom; there is something that remains up to Jones (Widerker 1995; Ginet 1996; Kane 1996). This response, famously dubbed the “flickers of freedom” reply, was criticized by defenders of Frankfurt’s argument on the ground that the “alternative possibilities” retained by Jones are not sufficiently “robust” (Fischer 1994, 2003; Fischer & Ravizza 1998). The critics of the argument rejected this charge, arguing that Jones retains a morally relevant ability to do otherwise, thus resurrecting the very debate that Frankfurt had hoped to undermine. Other defenders of Frankfurt’s argument told more complicated stories and argued that, even if Frankfurt’s story does not succeed, these new stories show that a person may be morally responsible for her action despite lacking any ability to do otherwise (Mele & Robb 1998; Pereboom 2003). More than fifty years after the publication of Frankfurt’s article, the debate continues. For a sample of some of this vast literature, see Widerker & McKenna 2003.

Frankfurt’s aim was to make it easier to defend the claim that moral responsibility is compatible with determinism and there is a case to be made that he was successful or, at least, successful against the “Source” incompatibilists who were convinced by his argument (Fischer 2003, Sartorio 2016). But there has been a cost. Our interest in free will is not limited to our interest in moral responsibility. Compatibilism has historically been charged with being “a wretched subterfuge” and a “quagmire of evasion”. Determinism seems prima facie incompatible with genuine choice and the ability to do otherwise, and while there was a time when the philosophical consensus was that the incompatibilist is guilty of some simple confusion or mistake akin to the fatalist’s mistake, this is no longer the case, thanks to the Consequence argument (to be discussed later). The influence of Frankfurt’s argument (and his other essays, especially Frankfurt 1971) has been so great that most compatibilists have turned their attention away from the free will/determinism problem and focused exclusively on problems about moral responsibility (and related topics, like blame, desert, and punishment). The term ‘compatibilism’ is now often used to refer to a thesis about moral responsibility—that the freedom sufficient for moral responsibility is compatible with determinism (Pereboom 1995), that the existence of actions free in the sense required for moral responsibility is compatible with determinism (Markosian 1999), that the unique ability of persons to exercise control in the manner necessary for moral responsibility is compatible with determinism (McKenna 2004b [2015]). The literature on the traditional problem of free will and determinism is dominated by incompatibilists. There is a growing consensus that the incompatibilist is right: if our universe is a deterministic one, we never have the ability to choose and do anything other than what we actually do.

Before we ask whether this pessimism about the compatibility of free will with determinism is warranted, we should pause to ask whether there really is a substantive disagreement between compatibilists and incompatibilists. When an incompatibilist says that determinism would rob us of the free will we think we have, including genuine choices and the ability to do otherwise, and when the compatibilist denies this, are they asserting and denying the same proposition? Or is the incompatibilist asserting one thing while the compatibilist is denying something else?

Some of the things said in the literature suggest that there is no substantive debate. For instance, some philosophers contrast a “strong incompatibilist ability” with a “weak compatibilist ability” (or “libertarian free will” with “compatibilist free will”) and write in a way that suggests that they think that the only substantive question about free will is whether it is or entails the “strong” incompatibilist ability to do otherwise (since the “weak” compatibilist ability to do otherwise is, by definition , compatible with determinism). And one leading semantic proposal might seem to support the claim that there is no real dispute. According to the Lewis/Kratzer, proposal ‘can’ (and other modal words, including ‘is able’, ‘has the power’, ‘is free to’) means ‘compossible with relevant facts F’, and the relevant facts are determined by the context together with the intentions of the speaker (D. Lewis 1976, Kratzer 1977). For a different kind of contextualist proposal see Hawthorne 2001; for criticism, see Feldman 2004. Whittle 2021 is a book-length defense of a contextualist account of freedom and responsibility.

So, for instance, it might be true, given one context, that I can speak Finnish (I’m a native Estonian speaker, so it wouldn’t take that long for me to learn Finnish) and false, given another context (don’t take me to Finland as your interpreter). There is no contradiction in saying that I both can and can’t speak Finnish, so long as we understand that what we are saying is that my speaking Finnish is compossible with one set of facts (my ability to easily learn the language) and not compossible with another, more inclusive set of facts (my current inability to speak it). Given this understanding of ‘can, it might seem that there is no genuine disagreement between the compatibilist and the incompatibilist. When the compatibilist asserts that deterministic agents are often able to do otherwise, she has in mind contexts (the contexts of ordinary speakers, in daily life) in which the relevant facts \(F\) are restricted to some relatively local facts about the agent and her surroundings. When the incompatibilist denies that any deterministic agent is ever able to do otherwise, she has in mind contexts (the contexts of philosophers discussing the free will/determinism problem) in which the relevant facts include all the facts about the laws and the past . So the proposition denied by the incompatibilist is not the proposition asserted by the compatibilist.

The leading contemporary incompatibilist, van Inwagen, rejects any suggestion that the compatibilist and the incompatibilist have a merely verbal disagreement, either because they are using different senses of ‘can’ (‘ability’, ‘power’, “free will” etc.) or because they are focusing on different contexts of utterance. The debate, he says, is about whether determinism has the consequence that no one is ever able to do otherwise (equivalently, that no one ever has it in their power to do otherwise) given what ordinary speakers mean, in the contexts in which they use these words. The contexts to which he is referring are the contexts of deliberation and choice in which we consider our options, while believing that we are able to pursue each of them. When a compatibilist asserts, and an incompatibilist denies, that a person at a deterministic world is sometimes able to do otherwise, they mean exactly the same thing by ‘is able to’; they mean what ordinary English speakers mean (in these contexts). The proposition asserted by the compatibilist is the proposition denied by the incompatibilist. Citing David Lewis as his example of a compatibilist opponent, van Inwagen says that he and Lewis cannot both be right. One of them is wrong, but neither is muddled or making a simple mistake (van Inwagen 2008).

In what follows, we will assume that the debate about free will (including, but not necessarily limited to, genuine choice and the ability to do otherwise) and determinism is a substantive debate, and not one that can be dissolved by appeal to different senses or contexts of utterance. We will now turn to the arguments.

3. Arguments based on Intuition

Some arguments for incompatibilism don’t fall into either of the two varieties described above—arguments that determinism is incompatible with ultimate sourcehood and arguments that determinism is incompatible with choice and the ability to do otherwise. These are arguments that appeal primarily to our intuitions. There are many variations on this way of arguing for incompatibilism, but the basic structure of the argument is usually something like this:

If determinism is true, we are like: billiard balls, windup toys, playthings of external forces, puppets, robots, victims of a nefarious neurosurgeon who controls us by directly manipulating the brain states that are the immediate causes of our actions. Billiard balls (windup toys, etc.) have no free will. So if determinism is true, we don’t have free will.

Most of these intuition-based arguments are not very good. Even if determinism entails that there is something we have in common with things which lack free will, it doesn’t follow that there are no relevant differences. Billiard balls, toys, puppets, and simple robots lack minds, and having a mind is a necessary condition of having free will. And determinism doesn’t have the consequence that all our actions are caused by irresistible desires that are, like the neurosurgeon’s direct manipulations, imposed on us by external forces outside our control. (For criticism of “intuition pumps”, see Dennett 1984. For discussion of cases involving more subtle kinds of manipulation, see Section 3.2 .)

With this caveat in mind, let’s take a closer look at the two most influential intuition-based arguments.

The No Forking Paths argument (van Inwagen 1983; Fischer 1994; Ekstrom 2000) begins by appealing to the idea that whenever we make a choice we are doing (or think we are doing) something like what a traveler does when faced with a choice between different roads. The only roads the traveler is able to choose are roads which are a continuation of the road she is already on. By analogy, the only choices we are able to make are choices which are a continuation of the actual past and consistent with the laws of nature . If determinism is false, then making choices really is like this: one “road” (the past) behind us, two or more different “roads” (future actions consistent with the laws) in front of us. But if determinism is true, then our journey through life is like traveling (in one direction only) on a road which has no branches. There are other roads, leading to other destinations; if we could get to one of these other roads, we could reach a different destination. But we can’t get to any of these other roads from the road we are actually on. So if determinism is true, our actual future is our only possible future ; we are never able to choose or do anything other than what we actually do. (See also Flint 1987 and Warfield 2003 for discussion of a related argument that appeals to the metaphor of our freedom to “add” to the list of truths about the world.)

This is intuitively powerful, since it’s natural to think of our future as being “open” in the branching way suggested by the road analogy and to associate this kind of branching structure with freedom of choice. But several crucial assumptions have been smuggled into this picture: assumptions about time and causation and assumptions about possibility. The assumptions about time and causation needed to make the analogy work include the following: that we “move” through time in something like the way that we move down a road, that our “movement” is necessarily in one direction only, from past to future, that the past is necessarily “fixed” or beyond our control in some way that the future is not. These assumptions are all controversial; on some theories of time and causation (the four-dimensionalist theory of time, a theory of causation that permits time travel and backwards causation), they are all false (D. Lewis 1976; Horwich 1987; Sider 2001; Hoefer 2003).

The assumption about possibility is that possible worlds are concrete spatio-temporal things (in the way that roads are) and that worlds can overlap (literally share a common part) in the way that roads can overlap. But most possible worlds theorists reject the first assumption and nearly everyone rejects the second assumption (D. Lewis 1986).

Determinism (without these additional assumptions) does not have the consequence that our “journey” through life is like moving down a road; the contrast between non-determinism and determinism is not the contrast between traveling on a branching road and traveling on a road with no branches.

As an argument for incompatibilism, the appeal to the metaphor of the branching roads (“the garden of forking paths”) fails. If we strip away the metaphors, the main premise of the argument turns into the claim that we have genuine choices between alternative course of action only if our choosing and doing otherwise is compossible with the actual past and the actual laws . But this claim is none other than a statement of what the incompatibilist believes and the compatibilist denies.

If the intuitions to which the No Forking Paths argument appeals nevertheless continue to engage us, it is because we think that our range of possible choices is constrained by two factors: the laws and the past. We can’t change or break the laws; we can’t causally affect the past. (Even if backwards causation is logically possible, it is not within our power.) These beliefs—about the laws and the past—are the basis of the most influential contemporary argument for incompatibilism: the Consequence argument. More of this later.

3.2 Manipulation and Design Arguments

We turn now to a family of arguments that work by appealing to our intuitive response to cases involving two persons, whom we will call “Victim” and “Producer”. Producer designs or manipulates Victim (in some of the stories, in the way the maker of a robot designs his robot or a god creates a human being; in other stories, by employing techniques of behavioral engineering or neural manipulation). Producer’s purpose is to ensure either that Victim performs a specific action (Mele 1995, 2006, 2019; Rosen 2002; Pereboom 1995, 2001, 2008, 2014) or that he will have the kind of psychology and motivational structure which will ensure or make probable that he performs certain kinds of actions and leads a certain kind of life. (See Kane 1996 for discussion of Huxley’s Brave New World and Skinner’s Walden Two .)

We are supposed to agree that Victim is not morally responsible because he acts unfreely and that he acts unfreely because of Producer’s role in the causation of his actions. Victim performs the actions he performs because that’s what he was designed or more directly manipulated to do, and it was Producer who made him be that way or do those things.

The argument then goes as follows:

  • Victim doesn’t act freely and, for that reason, is not morally responsible for what he does.
  • If determinism is true, there is no relevant difference between Victim and any normal case of apparently free and morally responsible action.
  • Therefore, if determinism is true no one ever acts freely or is morally responsible for what he does.

We are supposed to accept premise 1 on the grounds of our intuitive response to the story about Victim. The argument for premise 2 is that if determinism is true, then we are like Victim with respect to the fact that we are merely the proximate causes of our actions. We do what we do because of the way we are (our psyche or “design” together with the total mix of our thoughts, desires, and other psychological states at the time of action) and the causes of these psychological characteristics ultimately come from outside us, from forces and factors beyond our control . The only difference between us (in this imagined scenario in which determinism is true) and Victim is that our psychological features are not the causal upshot of the work of a single Producer who had a specific plan for us. But this fact about the remote causes of our actions—that they are caused by a variety of natural causes rather than the intentional acts of a single agent—is not relevant to questions about our freedom and responsibility. Or so it is argued, by the advocates of Manipulation arguments.

Manipulation arguments may be seen as the Source incompatibilist’s response to Frankfurt’s thought experiment. Frankfurt was trying to show that determinism isn’t as bad as we might think. In his story, Black was a stand-in for determinism, and Frankfurt was trying to convince us that the facts about Black are consistent with the facts, as we know them, about how we actually deliberate, decide, and act, and these facts are the only facts that matter, so far as moral responsibility is concerned. So even if Jones lacks the ability to do otherwise, he is still morally responsible. The Manipulation argument says, in effect:

Frankfurt’s thought experiment is faulty. Determinism isn’t a powerful agent, standing by, in the background, like Black. Determinism is part of the “actual sequence”. Let me tell you a story to make this clear…

And then Producer is introduced, and we are told that he has a plan concerning the action or actions of another person, Victim, the power to enforce his plan, and moreover, unlike Black , he does enforce it.

It would be a mistake, however, to think that manipulation of one person by another automatically undermines freedom. In real life, we know that we may be manipulated by others to do things we would not have done, but for their arguments or other ways of persuading us to change our minds. Absent further reasons, we don’t think that this kind of manipulation robs us of free will or the ability to act freely in the way required for moral responsibility. We think that we could have resisted the argument (or the sales pitch or the subtle pressures exerted by our manipulative friend or colleague) and we might blame ourselves later for not doing so.

The question, then, is whether there is a case that can serve the purposes of a manipulation argument: a case where Victim lacks the freedom that is a necessary condition of moral responsibility while not being different, in any relevant way, from a normal agent in normal circumstances at a deterministic world (that is, from someone who we think acts freely and is morally responsible for what she does).

There are cases and cases, and many of the ones in the literature are under-described. The first three Plums of Pereboom 2001 are an example. Depending on how the details are spelled out, our verdict about Plum might be that he is unfree and not morally responsible because he is different, in relevant ways that don’t require the falsity of determinism, from a normal agent in normal circumstances (Fischer 2004; Mickelson 2010; Sripada 2012). Alternatively, the story might be fleshed out in a way that supports the judgment that Plum is not different, in any relevant way, from a normal agent (deterministic or indeterministic) in normal circumstances. But this leaves it open to the compatibilist to take the hard-line reply (McKenna 2004a, 2008; Jeppsson 2020) that since the normal deterministic agent is morally responsible, so is Plum. Opinions vary as to whether the intuitive cost of the hard-line reply is too great (Pereboom 2008; McKenna 2014).

Consider, next, cases of the Brave New World variety—cases where children are subjected to intensive behavioral engineering from birth, in a way intended to make them accept their assigned roles in a rigidly hierarchical society. Everything depends on the details, but it is surely not implausible to think that the subjects of some Brave New World cases lack a morally significant freedom because their cognitive, evaluational, and volitional capacities have been stunted or impaired in certain ways:

they are incapable of effectively envisaging or seeing the significance of certain alternatives, of reflecting on themselves and on the origins of their motivations. (Watson 1987)

Wolf 1990 argues, on the basis of similar cases, that the ability that grounds our freedom and responsibility is the unimpaired capacity to choose and act in accordance with “the True and the Good” (see also Nelkin 2011). But determinism doesn’t have the consequence that everyone’s cognitive, evaluational and volitional capacities are impaired in these sorts of ways.

There are cases where Victim is under the direct control of Producer in a way that makes it true that Victim is not morally responsible for what she does because she no longer has the kind of causal control that is a necessary condition acting freely. Producer might, for instance, directly manipulate Victim’s limbs so that Victim finds her body moving, puppetlike, in ways she does not intend. Or Producer might directly manipulate the brain states that are the neural realizers of Victim’s first order desires so Victim is, like Frankfurt’s unwilling addict, acting intentionally but “against her will” (Frankfurt 1971). But determinism doesn’t have the consequence that we never act intentionally and it doesn’t have the consequence that we always act contrary to our second order volitions or that we are always subject to psychological compulsions, addictive or “irresistible” desires, phobias, and other pathological aversions.

Defenders of Manipulation arguments claim, however, that the argument works even if these kinds of cases are set aside. They say that the intuitive force of the argument depends only on the fact that deterministically caused actions are ultimately caused, as are Victim’s, by factors and forces outside the agent’s control. They say that the argument succeeds even in cases where Producer designs a Victim with unimpaired capacities and a normal psychology, perhaps the kind we’d like our children to have, perhaps a rationally egoistic psychology of a kind we would prefer our children not to have. They also say that the argument succeeds even when Producer is such a sophisticated designer of Victim that Victim has a past history that satisfies the requirements of those compatibilist accounts of free agency that include a historical condition. (For a helpful account of the difference between historical and nonhistorical compatibilist accounts in the context of Manipulation arguments, see McKenna 2004a. For argument that the best explanation for our belief that manipulated agents act unfreely is the belief that these agents don’t have it in their power to do otherwise, see Berofsky 2012. For argument that Manipulation arguments fail given a Strawsonian account of responsibility, see Latham and Tierney 2022.)

The best example of this kind of case is Mele’s Ernie. Ernie was created as a zygote \(Z\) in Mary by goddess Diana because Diana wanted a certain event to take place 30 years later, and she was able to use her knowledge of the deterministic laws and the state of the entire universe to deduce that placing a zygote with precisely \(Z\)’s constitution in Mary would produce a normal (or better than normal: ideally self-controlled, rational, etc.) adult who would, 30 years later, judge, on the basis of rational deliberation, that it is best to do \(A\) and who would \(A\) on the basis of her judgment, thereby bringing about \(E\) (Mele 2006, 2019).

To many people, it seems intuitively clear that Ernie acts unfreely and is for that reason not morally responsible for what he does. But if that’s right, it looks like this version of the Manipulation argument succeeds. For consider this: Ernie has an atom-for-atom duplicate, Bert, a normal guy in every way, exactly like Ernie (ideally self-controlled, rational, etc.) except for the fact that he was not created by a goddess. Bert finds himself in circumstances exactly like the ones Ernie is in, and Bert also judges, on the basis of rational deliberation, that it is best to do \(A\), and he also acts on the basis of his judgment, thereby bringing about \(E\). There appears to be no relevant difference between Ernie and Bert. Therefore, Bert also acts unfreely and is also not morally responsible for what he does. But Bert (like Ernie) is normal in every way, and we can also stipulate that he (like Ernie) satisfies all plausible compatibilist conditions (historical as well as nonhistorical) for being a free and morally responsible agent. If Bert acts unfreely, so does every deterministic agent on every occasion. Therefore the kind of freedom necessary for moral responsibility is incompatible with determinism.

Mele claims that the case of Ernie is an improvement on earlier Manipulation cases in two ways. First, it is a case where it is clear that the causes of Ernie’s action are deterministic rather than the kinds of causes that might be found at a non-deterministic world. Second, it is a case where it is clear that there is no relevant difference between Ernie and any case of apparently free and responsible action at a deterministic world.

Deery and Nahmias 2017 appeal to interventionist theories of causation to argue that Mele is wrong about the second claim. Tierney and Glick 2020 criticise Deery and Nahmias. (See also Herdova 2020.)

Suppose, however, that there is no relevant difference between Ernie and Bert. It does not follow that the argument succeeds. If there really is no freedom-relevant difference between Bert and Ernie, why should we reason from the unfreedom of Ernie to the unfreedom of Bert rather than the other way around, from the freedom of Bert to the freedom of Ernie?

After all, our grounds for saying that there is no relevant difference between the two is that the historical facts about Ernie’s creation (that he was created by a goddess, with the powers of a Laplacian predictor, with certain intentions, and so on) are not relevant to the question of whether Ernie acts freely or unfreely 30 years later. If they are not relevant, they don’t provide us with reasons for thinking that Ernie is unfree. By contrast, we do have reasons for thinking that Bert acts freely and is morally responsible for what he does; he satisfies the ordinary conditions we use in real life, as well as all the conditions of the best compatibilist accounts on offer. If we started out with reasons for thinking Bert free and responsible (either because we are already compatibilists or because we have never thought about determinism), the Zygote argument hasn’t given us any reason to change our minds. (For further elaboration on this critique, including some helpful counter-thought experiments, see Fischer 2011, 2021; Kearns 2012.)

A defender of the Zygote argument might respond by claiming that the intuitions that favor the unfreedom and lack of responsibility of Ernie are stronger than the intuitions that favor the freedom and responsibility of Bert. But this is problematic. Intuitions are not always evidentially trustworthy and there are, as we have already noted, reasons for being wary of our intuitive response to descriptions of deterministic worlds: we have a natural tendency to confuse “it will be” with “it must be”. There are further reasons for not giving evidential credence to our intuitions about Ernie’s unfreedom: he was created by a goddess, with the kind of foreknowledge that no human being could have, for the express purpose of performing a certain action 30 years later. Perhaps our intuitions are explained (though not justified) by the belief that being created in this way robs Ernie of the freedom required for responsibility. The first premise of the Zygote argument must be defended by something more than appeal to intuition.

(For some recent empirical work on intuitions about free will, determinism, and moral responsibility, see Nahmias, Morris, Nadelhoffer, and Turner 2006, Nahmias 2011 and Murray & Nahmias 2014. For a critique of the use of intuitions in Manipulation arguments, see Vihvelin 2013 and Spitzley 2015.)

Defenders of the Zygote argument don’t dispute these points. The Manipulation argument works only if the second premise is true, and the second premise says that there is no relevant difference between Victim (in this case, Ernie) and any normal deterministic case of apparently free and responsible action (in this case, Bert). Ernie differs from Bert with respect to certain historical facts about his creation: the fact that he was created by a goddess with foreknowledge, and intentions about his future. So none of these facts can be counted relevant, even if they affect our intuitions . But one historical fact is relevant, according to the defender of the argument: the fact that the deterministic causes of Ernie’s action trace back thirty years to conditions outside his control. The claim, then, is that Ernie acts unfreely and without responsibility because determinism is true. But this claim was supposed to be the conclusion of the argument, not the premise.

There is one remaining possibility. Perhaps our intuitions are best explained by our belief that Ernie doesn’t have it in his power to do anything other than act according to Diana’s plan. (We think it unfair to blame someone who wasn’t able to avoid doing what he did.) If that’s the case, then the back story of Ernie’s creation functions as a way of making vivid one of the possible ways that an otherwise normal agent at a deterministic world might have got to be the way that he is at a certain moment in his life. Should we reason from our belief that Ernie isn’t, on that occasion, able to do otherwise to the belief that Bert and every normal deterministic agent also lacks this ability, or should we reason in the opposite direction, from Bert’s ability to do otherwise to Ernie’s? This requires further discussion, which we will take up in the section on the Consequence argument. (See also Mickelson 2015b and Huoranskzi 2021.)

What has this boy to do with it? He was not his own father; he was not his own mother; he was not his own grandparents. All of this was handed to him. He did not surround himself with governesses and wealth. He did not make himself. And yet he is to be compelled to pay (Darrow 1924: 65).

Let’s turn now to arguments for incompatibilism based on the idea that a free and responsible action is an action that is caused and controlled by its agent in a special “the buck stops here” way, a way that is incompatible with deterministic event-causation.

The most popular instance of this kind of argument is an argument that we will call “the desperate defense attorney’s argument” (Darrow 1924). The defense attorney’s argument is simple:

  • My client is responsible for his crime only if he “made himself”—that is, only if he caused himself to be the kind of person he is.
  • My client did not make himself.
  • Therefore my client is not responsible for his crime.

The defense attorney is trying to persuade the jurors that his client is not responsible for his action, but not for any of the standard excusing conditions—insanity, accident, mistaken belief, duress, mental handicap, and so on. Nor does he claim that there is anything that distinguishes his client from any of the rest of us. His argument is that his client is not responsible because he did not make himself. But none of us has made ourselves (at least not from scratch)—we are all the products of heredity and environment. So if we accept the defense attorney’s argument, it appears that we are committed to the conclusion that no one is ever responsible for anything .

It’s not clear that this is an argument for incompatibilism. It’s an argument for incompatibilism only if it’s an argument for hard determinism—that is, if it’s an argument for the thesis that determinism is true and because of this we are never responsible for anything. Let’s take a closer look.

What’s the argument for premise (2)? After all, we do make our selves, at least in the garden-variety way in which we make other things: we plant gardens, cook dinners, build boats, write books and, over the course of our lives, re-invent, re-create, and otherwise “make something of ourselves”. Of course we don’t do any of these things “from scratch”, without help from anyone or anything else, but it’s impossible (or at least impossible for human beings) to make anything from scratch. The truth or falsity of determinism has no bearing on this point.

(See G. Strawson 1986 and 1994 for an argument for the impossibility of “true responsibility” that is a more sophisticated version of the defense attorney’s argument. See also Smilansky 2000. See Mele 1995 and Clarke 2005 for a critique of Strawson’s argument.)

If we pressed our defense attorney (or brought in a philosopher to help him out), we might get the following reply: The kind of garden-variety self-making possible at a deterministic world is not good enough for the kind of moral responsibility required for deserved blame and punishment. Granted, we can never have complete control over the actions we perform because of our choices (Nagel 1979), and this limits the control we have over our self-making . But we are morally responsible for our actions only if we have at least some control over our self-making, and we have control over our self-making only if we have control over the choices that are the causes of the actions whereby we make our selves. And we have control over these choices only if we cause our choices and no one and nothing causes us to make them (Chisholm 1964).

The defense attorney (or philosopher) is defending premise (2) by arguing for a certain interpretation of premise (1)—that our responsibility for our actions requires that we have “made ourselves” in the sense that, over the course of our lives, we have frequently been the first cause of the choices that result in actions and thus eventually (albeit often in ways we can neither predict nor control) to changes in our selves. In arguing this way, he has shifted the focus of the argument from the obviously impossible demand that the freedom required for moral responsibility requires entirely self-made selves to the intuitively appealing and at least not obviously impossible demand that the freedom required for moral responsibility requires what Robert Kane has called “ultimate responsibility” (that we are the ultimate sources or first causes of at least some of our choices—those choices that are the causes of “self-forming actions”). (See Kane 1996, 1999, 2008, 2011a. See also Chisholm 1964; Clarke 1993, 1996, 2003; O’Connor 1995a, 2000, 2011; Pereboom 2001, 2009b, 2014; Steward 2012).

This brings us to the philosopher’s version of the defense attorney’s argument. (For variations on this kind of argument, see Kane 1996, 1999, 2008, 2011a and Pereboom 2001, 2005, 2014.)

  • We act freely (in the way necessary for moral responsibility) only if we are the ultimate sources (originators, first causes) of at least some of our choices.
  • If determinism is true, then everything we do is ultimately caused by events and circumstances outside our control.
  • If everything we do is ultimately caused by events and circumstances outside our control, then we are not the ultimate sources (originators, first causes) of any of our choices.
  • Therefore, if determinism is true, we are not the ultimate sources of any of our choices.
  • Therefore, if determinism is true, we never act freely and we are never morally responsible.

Premise (2) follows from the definition of determinism (at least given two widely accepted assumptions: that there is causation in a deterministic universe and that causation is a transitive relation). (For some doubts about the latter assumption, see Hall 2000.) Premise (3) is clearly true. So if we want to reject the conclusion, we must reject Premise (1).

Compatibilists have argued against (1) in two different ways. On the positive side, they have argued that we can give a satisfactory account of the (admittedly elusive) notion of self-determination without insisting that self-determination requires us to be the first causes of our choices (Frankfurt 1971, 1988; Watson 1975, 1987, 2004; Dennett 1984; Wolf 1990; Fischer 1994; Fischer & Ravizza 1998; Bok 1998; Nelkin 2011; Sartorio 2016). On the negative side, compatibilists have challenged (1) by arguing that it is of no help to the incompatibilist: if we accept (1), we are committed to the conclusion that free will and moral responsibility are impossible , regardless of whether determinism is true or false.

The challenge to (1) takes the form of a dilemma: Either determinism is true or it’s not. If determinism is true, then my choices are ultimately caused by events and conditions outside my control, so I am not their first cause and therefore, if we accept (1), I am neither free nor responsible. If determinism is false, then something that happens inside me (something that I call “my choice” or “my decision”) might be the first event in a causal chain leading to a sequence of body movements that I call “my action”. But since this event has no sufficient cause, whether or not it happens is not under my control any more than the spinning of a roulette wheel inside my brain is under my control. Therefore, if determinism is false, I am not the first cause or ultimate source of my choices and, if we accept (1), I am neither free nor responsible (Ayer 1954; Wolf 1990).

In order to defend (1) against the “determined or random” dilemma, above, the incompatibilist has to offer a positive account of the puzzling claim that persons are the first causes of their choices. The traditional incompatibilist answer is that this claim must be taken literally, at face value. We—agents, persons, enduring things—are causes with a very special property: we initiate causal chains, but nothing and no one causes us to do this. Like God, we are uncaused causers, or first movers. If Joe deliberately throws a rock, which breaks a window, then the window’s breaking (an event) was caused by Joe’s throwing the rock (another event), which was caused by Joe’s choice (another event). But Joe’s choice was not caused by any further event, not even the event of Joe’s thinking it might be fun to throw the rock; it was caused by Joe himself . And since Joe is not an event, he is not the kind of thing which can be caused. (Or so it is argued, by agent-causalists. See Chisholm 1964; O’Connor 1995a, 2000, 2011; Pereboom 2001.)

Many philosophers think that agent-causation is either incoherent or impossible, due to considerations about causation. What sense does it make to say that a person, as opposed to a change in a person, or the state of a person at a time, is a cause? (Bok 1998). (See also Clarke 2003 for a detailed and sympathetic examination of the metaphysics of agent-causation, which ends with the conclusion that there are, on balance, reasons to think that agent-causation is impossible.)

Others (van Inwagen 2000; Mele 2006) have argued that even if agent-causation is possible, it would not solve the problem of transforming an undetermined event into one which is in our control in the way that our free choices must be. And others have argued that if agent-causation is possible, it is possible at deterministic as well as non-deterministic worlds (Markosian 1999; Nelkin 2011; Franklin 2016).

Some incompatibilists have responded to the “determined or random” dilemma in a different way: by appealing to the idea of probabilistic causation (Kane 1996, 1999, 2008, 2011a). If our choices are events which have probabilistic causes (e.g., our beliefs, desires, and other reasons for acting), then it no longer seems plausible to say that we have no control over them. We make choices for reasons, and our reasons cause our choices, albeit indeterministically. Kane’s reply may go some way towards avoiding the second (no control) horn of the dilemma. But it doesn’t avoid the first horn. If our choices are caused by our reasons, then our choices are not the first causes of our actions. And our reasons are presumably caused, either deterministically or probabilistically, so they are not the first causes of our actions either. But then our actions are ultimately caused by earlier events over which we have no control and we are not the ultimate sources of our actions.

We think that we make choices, and we think that our choices typically make a difference to our future. We think that there is a point to deliberation: how we deliberate—what reasons we consider—makes a difference to what we choose and thus to what we do. We also think that when we deliberate there really is more than one choice we are able to make, more than one action we are able to perform, and more than one future which is, at least partly, in our power to bring about.

Our beliefs about our powers with respect to the future contrast sharply with our beliefs about our lack of power with respect to the past. We don’t think we have any choice about the past. We don’t deliberate about the past; we think it irrational to do anything aimed at trying to change or affect the past (“You had your chance; you blew it. It’s too late now to do anything about it”). Our beliefs about our options, opportunities, alternatives, possibilities, abilities, powers, and so on, are all future-directed. We may summarize this contrast by saying that we think that the future is “open” in some sense that contrasts with the non-openness or “fixity” of the past.

Although we don’t think we (now) have a choice about the past, we have beliefs about what was possible for us in the past. When called upon to defend what we did, or when we blame or reproach ourselves, or simply wonder whether we did the right thing (or the sensible thing, the rational thing, and so on), we evaluate our action by comparing it to what we believe were our other possible actions, at that time . We blame, criticize, reproach, regret, and so on, only insofar as we believe we had alternatives. And if we later discover that we were mistaken in believing that some action \(X\) was among our alternatives, we think it is irrational to criticize or regret our failure to do \(X\).

Is determinism compatible with the truth of these beliefs? In particular, is it compatible with the belief that we are often able to choose and do more than one action?

Incompatibilists have traditionally said “No”. And it’s not hard to see why. If we think of ‘can’ in the “open future” way suggested by the commonsense view, then it’s tempting to think that the past is necessary in some absolute sense. And it’s natural to think that we are able to do otherwise only if we can do otherwise given this “fixed” past ; that is, only if our doing otherwise is a possible continuation of the actual past. If we follow this train of thought, we will conclude that we are able to do otherwise only if our doing otherwise is a possible continuation of the past consistent with the laws. But if determinism is true, there is only one possible continuation of the past consistent with the laws. And thus we get the incompatibilist conclusion. If determinism is true, our actual future is our only possible future. What we actually do is the only thing we are able to do.

But this argument is too quick. It rests on assumptions about the nature of time which are arguably at odds with what physics tells us. (Hoefer 2003 [2016], Ismael 2016.) There is an alternative explanation for our beliefs about the “open” future as opposed to the “fixed” past—the direction of causation. Causes are always earlier than their effects. Our deliberation causes our choices, which cause our actions. But not the other way around. Our choices cause future events; they never cause past events. Why causation works this way is a deep and difficult question, but the leading view, among philosophers of science, is that the temporal asymmetry of causation is a fundamental but contingent fact about our universe. If things were different enough—if we could travel backwards in time—then we would sometimes have an ability that we don’t actually have—the ability to causally affect past persons and things (Horwich 1987; D. Lewis 1976). If this is right, then we don’t need to suppose that the past is metaphysically or absolutely necessary in order to explain the fixed past/open future contrast. The past could have been different . But, given the way things actually are, there is nothing that we are able to do that would causally affect the past .

This alternative explanation of our commonsense belief about the contrast between fixed past and open future allows the compatibilist to say the kind of things that compatibilists have traditionally wanted to say: The ‘can’ of our freedom of will and freedom of action is the ‘can’ of causal and counterfactual dependence. Our future is open because it depends, causally and counterfactually, on our choices, which in turn depend, causally and counterfactually, on our deliberation and on the reasons we take ourselves to have. (At least in the normal case, where there is neither external constraint nor internal compulsion or other pathology.) If our reasons were different (in some appropriate way), we would choose otherwise, and if we chose otherwise, we would do otherwise. And our reasons can be different, at least in the sense that we, unlike simpler creatures and young children, have the ability to critically evaluate our reasons (beliefs, desires, values, principles, and so on) and that we have, and at least sometimes exercise, the ability to change our reasons (Bok 1998; Dennett 1984; Fischer 1994; Fischer & Ravizza 1998; Frankfurt 1971, 1988; Lehrer 1976, 1980, 2004; Watson 1975, 1987, 2004; Wolf 1990; Smith 1997, 2004; Pettit & Smith 1996; Nelkin 2011; Vihvelin 2004, 2013). All this is compatible with determinism. So the truth of determinism is compatible with the truth of our commonsense belief that we really do have a choice about the future, that we really can choose and do other than what we actually do.

Incompatibilists think that this, and any compatibilist account of the ‘can’ of freedom of choice, is, and must be , mistaken. The Consequence Argument (Ginet 1966, 1983, 1990; van Inwagen 1975, 1983, 2000; Wiggins 1973; Lamb 1977) is widely regarded as the best argument for this conclusion. In the remainder of this section we will take a closer look at van Inwagen’s version of this important and influential argument.

In An Essay on Free Will (1983), van Inwagen presents three formal arguments which, he says, are intended as three versions of the same basic argument, which he characterized as follows:

If determinism is true, then our acts are the consequence of laws of nature and events in the remote past. But it’s 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. (1983: 56)

We will begin by looking at the third version of the argument (the Rule Beta argument). The Rule Beta argument uses a modal sentential operator which van Inwagen defines as follows: ‘\(\mathbf{N}p\)’ abbreviates ‘\(p\) and no one has, or ever had, any choice about whether \(p\)’. Van Inwagen tells us that the logic of ‘\(\mathbf{N}\)’ includes these two inference rules, where \(\Box p\) asserts that it is logically necessity that \(p\):

Alpha : From \(\Box p\), we may infer \(\mathbf{N}p\).

Beta : From \(\mathbf{N}p\) and \(\mathbf{N}(p \supset q)\), we may infer \(\mathbf{N}q\).

In the argument below, ‘\(L\)’ is an abbreviation for a sentence expressing a conjunction of all the laws of nature; ‘\(H\)’ is a sentence expressing a true proposition about the total state of the world at some time in the distant past before any agents existed; ‘\(\Box\)’ is ‘it is logically necessary that’; ‘\(\supset\)’ is the material conditional, and ‘\(P\)’ is a dummy for which we may substitute any sentence which expresses a true proposition.

The argument is a conditional proof: Assume determinism and show that it follows that no one has, or ever had, a choice about any true proposition, including propositions about the apparently free actions of human beings.

Premises (1) and (2) follow from determinism. (3) follows from (2), by application of rule Alpha. Rule Alpha seems uncontroversial. (But see Spencer 2017).

Premises 4 and 6 also look uncontroversial. \(N\) necessity isn’t logical or metaphysical necessity. We can insist that the laws and the distant past could , in the broadly logical sense, have been different, so neither \(\Box H\) nor \(\Box L\) are true. But it still seems undeniably true that we have no choice about whether the laws and the distant past are the way they are; there is nothing that we are able to do that would make it the case that either the laws or the distant past are different from the way they actually are.

Rule Beta is the key to the argument. It’s what makes the difference between this version of the Consequence Argument and an argument widely agreed to be fallacious.

\(\Box(P \supset Q)\) \(P\) Therefore, \(\Box Q\)

An example of this invalid inference is an argument sometimes called “the fatalist fallacy”:

\(\Box\)(it’s true that it will rain tomorrow \(\supset\) it will rain tomorrow) It’s true that it will rain tomorrow Therefore, \(\Box\)(it will rain tomorrow)

Another example:

\(\Box((H \amp L) \supset P)\) \(H \amp L\) Therefore, \(\Box P\)

On the other hand, the following is a valid inference:

\(\Box P\) \(\Box(P \supset Q)\) Therefore, \(\Box Q\)

The necessity expressed by the ‘no choice about’ operator is not logical or metaphysical necessity. But it might nevertheless be similar enough for Beta to be a valid rule of inference. Or so argued van Inwagen, and gave examples:

\(\mathbf{N}\)(The sun explodes in the year 2000) \(\mathbf{N}\)(The sun explodes in the year 2000 \(\supset\) All life on earth ends in the year 2000) Therefore, \(\mathbf{N}\)(All life on earth ends in the year 2000)

An early response to the Consequence argument was to argue that Beta is invalid because a compatibilist account of the ability to do otherwise is correct (Gallois 1977; Foley 1979; Slote 1982; Flint 1987). For instance, if “\(S\) is able to do \(X\)” means “if \(S\) tried to do \(X\), \(S\) would do \(X\)”, then the premises of the argument are true (since even if \(S\) tried to change the laws or the past, she would not succeed), but the conclusion is false (since determinism is consistent with the truth of counterfactuals like “if \(S\) tried to raise her hand, she would”).

Incompatibilists were unmoved by this response, saying, in effect, that the validity of Beta is more plausible than the truth of any compatibilist account of ability to do otherwise. They pointed out that there was no agreement, even among compatibilists, about how such an account should go, and that the simplest accounts (so-called “Conditional Analyses”, originally proposed by Hume) had been rejected, even by compatibilists.

(For criticism of Simple Conditional Analyses, see Austin 1956; Chisholm 1964; Lehrer 1968, 1976; van Inwagen 1983. For defense of a compatibilist account of ability to do otherwise, see Moore 1912; Hobart 1934; Kapitan 1991, 1996, 2011; Lehrer 1980, 2004; Bok 1998; Smith 1997, 2004; Campbell 2005; Perry 2004, 2008, 2010; Vihvelin 2004, 2013; Fara 2008; Schlosser 2017; Menzies 2017; List 2019.)

More recently, van Inwagen has conceded that Beta is invalid (van Inwagen 2000). McKay and Johnson (1996) showed that Beta entails Agglomeration:

\(\mathbf{N}p\) \(\mathbf{N}q\) Therefore, \(\mathbf{N}(p \amp q)\)

Agglomeration is uncontroversially invalid. To see this, let ‘\(p\)’ abbreviate ‘The coin does not land heads’, let ‘q’ abbreviate ‘The coin does not land tails’, and suppose that it’s a fair coin which isn’t tossed but someone could have tossed it (McKay and Johnson 1996).

(For counterexamples to Beta, see Widerker 1987, Huemer 2000, and Carlson 2000.)

Van Inwagen proposed to repair the Consequence argument by replacing ‘\(\mathbf{N}\)’ with ‘\(\mathbf{N}\)*’, where ‘\(\mathbf{N}*p\)’ says “\(p\) and no one can, or ever could, do anything such that if she did it, \(p\) might be false”. Agglomeration is valid for ‘\(\mathbf{N}\)*’, and thus this particular objection to the validity of Beta does not apply.

It has also been suggested (Finch & Warfield 1998) that the Consequence argument can be repaired by keeping ‘\(\mathbf{N}\)’ and replacing Beta with Beta 2:

Beta 2 : From \(\mathbf{N}p\) and \(\Box(p \supset q)\), we may infer \(\mathbf{N}q\)

This would yield the following argument:

Other ways of repairing the argument have been proposed by O’Connor 1993 and Huemer 2000.

So it still looks as though the compatibilist is in trouble. For it seems plausible to suppose that there is nothing that we are able to do that might make it the case that either \(H\) or \(L\) is false. And it seems plausible to suppose that we have no choice about whether \((H \amp L)\). We need to dig deeper to criticize the argument.

David Lewis tells us to think of the argument as a reductio (Lewis 1981). A compatibilist is someone who claims that the truth of determinism is compatible with the existence of the kinds of abilities that we assume we have in typical choice situations. Let’s call these ‘ordinary abilities’. The Consequence argument, as Lewis articulates it, says that if we assume that a deterministic agent has ordinary abilities, we are forced to credit her with incredible abilities as well.

Here, with some modifications, is Lewis’s statement of the argument:

Pretend that determinism is true, and that I did not raise my hand (at that department meeting, to vote on that proposal) but had the ability to do so. If I had exercised my ability—if I had raised my hand—then either the remote past or the laws of physics would have been different (would have to have been different). But if that’s so, then I have at least one of two incredible abilities—the ability to change the remote past or the ability to change the laws. But to suppose that I have either of these abilities is absurd. So we must reject the claim that I had the ability to raise my hand.

This counterfactual version of the Consequence argument nicely highlights a point that the rule Beta version glosses over. The argument relies on a claim about counterfactuals. The argument says that if determinism is true, then at least one of these counterfactuals is true:

Different Past: If I had raised my hand, the remote past would have been different.

Different Laws: If I had raised my hand, the laws would have been different.

Both these counterfactuals strike many people as incredible. But there is a reason for that—we are not used to thinking in terms of determinism and we are not accustomed to counterfactual speculation about what would have been the case, beforehand , if anything at a deterministic world had happened in any way other than the way it actually happened.

On the other hand, we are good at evaluating counterfactuals, or at least some counterfactuals, and we are especially good at evaluating those counterfactuals that we entertain in contexts of choice, when we ask questions about the causal upshots of our contemplated actions. (What would happen if… I struck this match, put my finger in the fire, threw this rock at that window, raised my hand?) And when we contemplate our options, we take for granted the existence of many facts—including facts about the laws and the past.

In other words, when we evaluate counterfactuals in real life, we do so by considering imaginary situations which are very like the situation we are actually in, and we do not suppose that there are any gratuitous departures from actuality. And to suppose a difference in the past or the laws seems like a gratuitous difference.

So it is no surprise that when our attention is directed to Different Past and Different Laws , these counterfactuals strike us as incredible, or at least odd . But that doesn’t mean that they are false, and if determinism is true, then either Different Past or Different Laws is true.

So the first point is that we all need a theory of counterfactuals, and if determinism is true, the true counterfactuals will include either Different Past or Different Laws .

The second point is that the details of the correct compatibilist solution to the free will/determinism problem will turn on the details of the correct theory of counterfactuals.

(For similar criticisms of the Consequence argument, see Fischer 1983, 1988; Horgan 1985; Watson 1987; Vihvelin 1988, 2008, 2013; Rummens 2021.)

If Lewis’s theory of counterfactuals (D. Lewis 1973, 1979) is correct, or even more or less correct (Schaffer 2004), then the relevant counterfactuals about the past and laws, at a deterministic world, are:

Almost the Same Past: If I had raised my hand, the past would have been exactly the same until a time shortly before the time of my decision to raise my hand.

Slightly Different Laws: If I had raised my hand, the laws would have been ever so slightly different in a way that permitted a divergence from the lawful course of actual history shortly before the time of my decision to raise my hand.

On the other hand, if Lewis’s theory is wrong, and counterfactuals are always evaluated by holding the laws constant, then the relevant counterfactuals, at a deterministic world, are:

Same Laws : If I had raised my hand, the laws would still have been exactly the same.

Different Past : If I had raised my hand, past history would have been at least somewhat different all the way back to the Big Bang.

(For critique of Lewis’s theory, and defense of a theory of counterfactuals that holds the laws fixed, see Bennett 1984; Dorr 2016; Vihvelin 2017b.)

We’ve got to choose. We all need a theory of counterfactuals, and our theory should provide the correct verdicts for the uncontroversially true counterfactuals at deterministic as well as indeterministic worlds. Our choice is limited to a theory that accepts Different Laws or Different Past . Which theory we choose has nothing to do with the free will/determinism problem and everything with how we evaluate counterfactuals.

We can now explain the essence of Lewis’s reply to the counterfactual version of the Consequence argument in a way that doesn’t require you to accept Lewis’s theory of counterfactuals.

The argument trades on an equivocation between two counterfactuals.

There is a corresponding equivocation between two ability claims:

The problem with the argument, says Lewis, is that it equivocates between these two ability claims. To count as a reductio against the compatibilist, the argument must establish that the compatibilist is committed to (A2) . But the compatibilist is committed only to (C1) and thus only to (A1) . The compatibilist is committed only to saying that if determinism is true, we have abilities we would exercise only if the past (or the laws) had been different in the appropriate ways. And while this may sound odd, it is no more incredible than the claim that the successful exercise of our abilities depends, not only on us, but also on the co-operation of things not in our control: the good or bad luck of our immediate surroundings. Since we are neither superheroes nor gods, we are always in this position, regardless of the truth or falsity of determinism.

The Consequence Argument was intended as an argument from premises that we must all accept—premises about our lack of control over the past and the laws—to the conclusion that if determinism is true, we don’t have the free will common sense says we have. The counterfactual version of the argument claims that if we attribute ordinary abilities to deterministic agents, we are forced to credit them with incredible past or law-changing abilities as well. But no such incredible conclusion follows. All that follows is something that we must accept anyway, as the price of our non-godlike nature: that the exercise of our abilities always depends, in part, on circumstances outside our control. (See also Fischer 1983, 1988, 1994; Watson 1987; Nelkin 2001; Vihvelin 2008, 2011, 2013, 2017; Kapitan 1991, 2011; Carlson 2000.)

If the aim of the Consequence argument was to show that no compatibilist account of ‘could have done otherwise’ can succeed, then Lewis is surely right; the reductio fails. The distinction between (A1) and (A2) permits the compatibilist to avoid making incredible claims about the powers of free determined agents. On the other hand, the incompatibilist surely has a point when she complains that it is difficult to believe that anyone has the ability described by (A1) . We believe that our powers as agents are constrained by the past and by the laws. One way to understand this belief is compatible with determinism: we lack causal power over the past and the laws . But it’s natural to understand the constraint in a different, simpler way: we are able to do only those things which are such that our doing of them does not counterfactually require a difference in either the past or the laws. And this leads more or less directly to the incompatibilist conclusion that if determinism is true, then we are never able to do otherwise.

This brings us back to our starting point. Our common sense web of beliefs about ourselves as deliberators, choosers, and agents includes the belief that the future is open in some sense that the past is not. It also includes the belief that our abilities and powers are constrained by the laws. One way of understanding these beliefs leads to incompatibilism; another way does not. Which one is right?

The Consequence argument is an attempt to provide an argument in defense of the incompatibilist’s way of understanding these common sense beliefs. Even if it fails as a reductio, it has been successful in other ways. It has made it clear that the free will/determinism problem is a metaphysical problem and that the underlying issues concern questions about our abilities and powers, as well as more general questions about the nature of causation, counterfactuals, and laws of nature. Can the abilities or powers of choosers and agents be understood as a kind of natural capacity or disposition? Is there a viable incompatibilist alternative? How should we understand counterfactuals about the alternative actions and choices of agents at deterministic worlds? Is the compatibilist proposal about the way in which the laws and past constrain us defensible? Are incompatibilists committed to the defense of a particular view about the nature of laws of nature? Are they committed to the rejection of a Humean view, for instance?

Insofar as the Consequence argument has pointed us in the direction of these deep and difficult underlying metaphysical questions, it represents a significant step forward in the discussion of one of the most intractable problems of philosophy. (For discussion of some of these issues, see D. Lewis 1979; Dennett 1984, 2003; Hoefer 2002; Berofsky 2003, 2012; Beebee & Mele 2002; Beebee 2000; Schaffer 2004, 2008; Vihvelin 1990, 2004, 2013, 2017; Perry 2004, 2008, 2010; van Inwagen 2004a; Fara 2008; Holton 2009; Wilson 2014; Clarke 2015; Ismael 2016; van Inwagen 2017; Spencer 2017; Franklin 2018; List 2019, Esfeld 2021; Loew and Huttemann 2022)

  • Albert, David Z., 1992, Quantum Mechanics and Experience , Cambridge, MA: Harvard University Pres..
  • Alvarez, Maria, 2009, “Actions, Thought Experiments, and the ‘Principle of Alternate Possibilities’”, Australasian Journal of Philosophy , 85(1): 61–81. doi:10.1080/00048400802215505
  • Anscombe, G. Elizabeth M., 1981, “Causality and Determination”, in Metaphysics and the Philosophy of Mind , (Collected philosophical papers of G.E.M. Anscombe, v. 2), Minneapolis, MN: University of Minnesota Press: 133–147. Originally her inaugural lecture at Cambridge University, 1971.
  • Armstrong, David, 1983, What is a Law of Nature? , Cambridge: Cambridge University Press.
  • Austin, J.L., 1956, “Ifs and Cans”, Proceedings of the British Academy, vol. 42: 109–132. doi:10.1093/019283021X.003.0009
  • Ayer, A.J., 2009, “Freedom and Necessity”, in Pereboom 2009a: ch. 12.
  • Balaguer, Mark, 2010, Free Will as an Open Scientific Problem , Cambridge, MA: MIT Press.
  • Beebee, Helen, 2000, “The Non-Governing Conception of Laws”, Philosophy and Phenomenological Research , 61(3): 571–594. doi:10.2307/2653613
  • –––, 2002, “Reply to Huemer on the Consequence Argument”, The Philosophical Review , 111(2): 235–241. doi:10.1215/00318108-111-2-235
  • –––, 2003, “Local Miracle Compatibilism”, Noûs , 37(2): 258–277. doi:10.1111/1468-0068.00438
  • –––, 2013, Free Will: An Introduction , New York: Palgrave Macmillan.
  • Beebee, Helen and Alfred Mele, 2002, “Humean Compatibilism”, Mind , 111(442): 201–23. doi:10.1093/mind/111.442.201
  • Bennett, Jonathan, 1984, “Counterfactuals and Temporal Direction”, The Philosophical Review , 93(1): 57–––91. doi:10.2307/2184413
  • –––, 2003, A Philosophical Guide to Conditionals , Oxford: Clarendon Press. doi:10.1093/0199258872.001.0001
  • Berofsky, Bernard, 2003, “Classical Compatibilism: Not Dead Yet”, in Widerker & McKenna 2003: 107–126.
  • –––, 2006a, “The Myth of Source”, Acta Analytica , 21(4): 3–18.
  • –––, 2006b, “Global Control and Freedom”, Philosophical Studies , 131(2): 419–445. doi:10.1007/s11098-004-7490-1
  • –––, 2012, Nature’s Challenge to Free Will , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199640010.001.0001
  • Bok, Hilary, 1998, Freedom and Responsibility , Princeton, NJ: Princeton University Press.
  • Brand, Myles and Douglas Walton (eds.), 1976, Action Theory: Proceedings of the Winnipeg Conference on Human Action, held at Winnipeg, Manitoba, Canada, 9–11 May 1975 , Dordrecht: Reidel.
  • Campbell, Joseph Keim, 2005, “Compatibilist Alternatives”, Canadian Journal of Philosophy , 35(3): 387–406. doi:10.1080/00455091.2005.10716595
  • –––, 2007, “Free Will and the Necessity of the Past”, Analysis , 67(2): 105–111. doi:10.1093/analys/67.2.105
  • Campbell, Joseph Keim, Michael O’Rourke, and David Shier (eds.), 2004, Freedom and Determinism , MIT Press: Bradford Books.
  • Campbell, Joseph Keim, Michael O’Rourke, and Matthew H. Slater (eds.), 2011, Carving Nature at its Joints: Natural Kinds in Metaphysics and Science , (Topics in Contemporary Philosophy, v. 8), Cambridge, MA: MIT Press. doi:10.7551/mitpress/9780262015936.001.0001
  • Capes, Justin A., 2012, “Action, Responsibility, and Ability to Do Otherwise”, Philosophical Studies , 158(1): 1–15. doi:10.1007/s11098-010-9662-5
  • –––, 2019, “What the Consequence Argument is an Argument For”, Thought: A Journal of Philosophy , 8(1): 50–56. doi:10.1002/tht3.404
  • Carlson, Erik, 2000, “Incompatibilism and the Transfer of Power Necessity”, Noûs , 34(2): 277–290. doi:10.1111/0029-4624.00211
  • Carroll, John, 2008, “Nailed to Hume’s Cross?”, in Sider, Hawthorne, & Zimmerman 2008: 67–81.
  • Chisholm, Roderick M., 1964, “Human Freedom and the Self”, The Lindley Lectures , Department of Philosophy, University of Kansas. Reprinted in Watson 2003: ch. 1 and Pereboom 2009a: ch. 14.
  • Clarke, Randolph, 1993, “Towards a Credible Agent-Causal Account of Free Will”, Noûs , 27(2): 191–203. Reprinted in Watson 2003: ch. 14 and Russell & Deery 2013: ch. 11. doi:10.2307/2215755
  • –––, 2003, Libertarian Accounts of Free Will , New York: Oxford University Press. doi:10.1093/019515987X.001.0001
  • –––, 2005, “On an Argument for the Impossibility of Moral Responsibility”, Midwest Studies in Philosophy , 29: 13–24. doi:10.1111/j.1475-4975.2005.00103.x
  • –––, 2015, “Abilities to Act”, Philosophy Compass , 10(12): 893–904. doi: 10.1111/phc3.12299
  • Cross, Charles, 1986, “‘Can’ and the Logic of Ability”, Philosophical Studies , 50(1): 53–64. doi:10.1007/BF00355160
  • Cutter, Brian, 2017, “What is the Consequence Argument an Argument For?”, Analysis , 77 (2): 278–287. doi: 10.1093/analys/anx052
  • Darrow, Clarence, 1924, The Trial of Leopold and Loeb, Attorney for the Damned: Clarence Darrow in the Courtroom , Arthur Weinberg (ed.), Chicago: The University of Chicago Press, 16–86.
  • Deery, Oisin, 2015, “The Fall from Eden: Why Libertarianism Isn’t Justified by Experience”, Australasian Journal of Philosophy , 93(2): 319–334. doi:10.1080/00048402.2014.968596
  • –––, 2015, “Why People Believe in Indeterminist Free Will”, Philosophical Studies , 172(8): 2033–2054. doi:10.1007/s11098-014-0396-7
  • Deery, Oisin, Matt Bedke, and Shaun Nichols, 2013, “Phenomenal Abilities: Incompatibilism and the Experience of Agency” in D. Shoemaker 2013: 126–150. doi:10.1093/acprof:oso/9780199694853.003.0006.
  • Deery, Oisin and Eddy Nahmias, 2017, “Defeating Manipulation Arguments: Interventionist Causation and Compatibilist Sourcehood”, Philosophical Studies , 174(5): 1255–1276. doi:10.1007/s11098-016-0754-8
  • Dennett, Daniel, 1984, Elbow Room: The Varieties of Free Will Worth Wanting , Cambridge, MA: MIT Press.
  • –––, 2003, Freedom Evolves, New York: Viking Penguin.
  • Dorr, Cian, 2016, “Against Counterfactual Miracles”, The Philosophical Review , 125(2): 241–286. doi:10.1215/00318108-3453187
  • Double, Richard, 1989, “Puppeteers, Hypnotists, and Neurosurgeons”, Philosophical Studies , 56(2): 163–173. Reprinted in Russell & Deery 2013: ch. 14. doi:10.1007/BF00355940
  • Earman, John, 1986, A Primer on Determinism , Dordrecht: D. Reidel Publishing Company.
  • –––, 2004, “Determinism: What We Have Learned and What We Still Don’t Know”, in Campbell, O’Rourke, & Shier 2004: 21–46.
  • Ekstrom, Laura, 2000, Free Will: A Philosophical Study , Boulder, CO: Westview Press.
  • Esfeld, Michael, 2021, “Super-Humeanism and Free Will”, Synthese , 198(7): 6245–6258. doi: 10.1007/s11229-019-02460-x
  • Fara, Michael, 2008, “Masked Abilities and Compatibilism”, Mind , 117(468): 843–865. doi:10.1093/mind/fzn078
  • Feldman, Richard, 2004, “Freedom and Contextualism”, in Campbell, O’Rourke, & Shier 2004: 255–276.
  • Finch, Alicia, 2013, “On Behalf of the Consequence Argument: Time, Modality, and the Nature of Free Action”, Philosophical Studies , 163(1): 151–170. doi:10.1007/s11098-011-9791-5
  • Finch, Alicia and Ted Warfield, 1998, “The MIND Argument and Libertarianism”, Mind , 107(427): 515–528. doi:10.1093/mind/107.427.515
  • Fischer, John Martin, 1983, “Incompatibilism”, Philosophical Studies , 43(1): 121–37. doi:10.1007/BF01112527
  • –––, 1986, “Power Necessity”, Philosophical Topics , 14(2): 77–91. doi:10.5840/philtopics19861424
  • –––, 1988, “Freedom and Miracles”, Noûs , 22(2): 235–252. doi:10.2307/2215861
  • –––, 1994, The Metaphysics of Free Will: An Essay on Control , Oxford: Blackwell.
  • –––, 2003, “Frankfurt-Style Compatibilism”, in Watson 2003: ch. 10.
  • –––, 2004, “Responsibility and Manipulation”, The Journal of Ethics , 8(2): 145–77. doi:10.1023/B:JOET.0000018773.97209.84
  • –––, 2009, “Ultimacy and Alternative Possibilities”, Philosophical Studies , 144(1): 15–20. doi:10.1007/s11098-009-9370-1
  • –––, 2011, “The Zygote Argument Re-mixed”, Analysis , 71(2): 267–272. doi:10.1093/analys/anr008
  • –––, 2021, “Initial Design, Manipulation, and Moral Responsibility”, Criminal Law and Philosophy 15: 255–270. doi:10.1007/s11572-021-09561-0
  • Fischer, John Martin and Mark Ravizza, 1998, Responsibility and Control: A Theory of Moral Responsibility , Cambridge: Cambridge University Press.
  • Fischer, John Martin and Patrick Todd, 2011, “The Truth about Freedom: A Reply to Merricks”, The Philosophical Review , 120(1): 97–115. doi:10.1215/00318108-2010-025
  • ––– (eds.), 2015, Freedom, Foreknowledge, and Fatalism , New York: Oxford University Press.
  • Flint, Thomas P., 1987, “Compatibilism and the Argument from Unavoidability”, Journal of Philosophy , 74(8): 423–40. doi:10.2307/2027000
  • Foley, Richard, 1979, “Compatibilism and Control over the Past”, Analysis , 39(2): 70–74. doi:10.1093/analys/39.2.70
  • Frankfurt, Harry G., 1969, “Alternate Possibilities and Moral Responsibility”, Journal of Philosophy , 66(23): 820–39. Reprinted in Frankfurt 1988, Watson 2003: ch. 8, Pereboom 2009a: ch. 15, and Russell & Deery 2013: ch 7. doi:10.2307/2023833
  • –––, 1971, “Freedom of the Will and the Concept of a Person”, Journal of Philosophy , 68(1): 5–20. Reprinted in Frankfurt 1988, Watson 2003: ch. 16, Pereboom 2009a: ch. 16, and Russell & Deery 2013: ch. 13. doi:10.2307/2024717
  • –––, 1988, The Importance of What we Care About , Cambridge: Cambridge University Press.
  • –––, 2002, “Reply to John Martin Fischer”, in Sarah Buss and Lee Overton (eds.), Contours of Agency: Essays on Themes from Harry Frankfurt , Cambridge, MA: MIT Press.
  • Franklin, Christopher Evan, 2014, “Powers, Necessity, and Determinism”, Thought , 3(3): 225–229. doi:10.1002/tht3.139
  • –––, 2016, “If Anyone Should Be an Agent-Causalist, then Everyone Should Be an Agent-Causalist”, Mind , 125(500): 1101–1131. doi:10.1093/mind/fzv177
  • –––, 2018, A Minimal Libertarianism: Free Will and the Promise of Reduction , New York: Oxford University Press.
  • Gallois, Andre, 1977, “Van Inwagen on Free Will and Determinism”, Philosophical Studies , 32(1): 99–105. doi:10.1007/BF00373718
  • Ginet, Carl, 1966, “Might We Have No Choice?” in Lehrer 1966: 87–104.
  • –––, 1983, “In Defense of Incompatibilism”, Philosophical Studies , 44(3): 391–400.
  • –––, 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: 403–417. Reprinted in Widerker & McKenna 2003: 75–90. doi:10.2307/2216254
  • Glick, David and Hannah Tierney, 2020, “Desperately Seeking Sourcehood”, Philosophical Studies , 177(4): 953–970. doi:10.1007/s11098-018-1215-3
  • Graham, Peter A, 2008, “A Defense of Local Miracle Compatibilism”, Philosophical Studies , 140(1): 65–82. doi:10.1007/s11098-008-9226-0
  • Hall, Ned, 2000, “Causation and the Price of Transitivity”, Journal of Philosophy , 97(4): 198–222. doi:10.2307/2678390
  • Hausmann, Marco, 2018, “The Consequence Argument Ungrounded”, Synthese , 195 (11): 4931–4950. doi: 10.1007/s11229-017-1436-6
  • Hawthorne, John, 2001, “Freedom in Context”, Philosophical Studies , 104(1): 63–79. doi:10.1023/A:1010398805497
  • Herdova, Marcela, 2021, “The Importance of Being Ernie”, Thought: A Journal of Philosophy , 10 (4): 257–263. doi:10.1002/tht3.503
  • Hobart, R.E., 1934, “Free Will as Involving Determination and Inconceivable without It”, Mind , 63(169): 1–27. Reprinted in van Inwagen & Zimmerman 2008: 420–432.
  • Hoefer, Carl, 2002, “Freedom from the Inside Out”, in Time, Reality, and Experience , Craig Callender (ed.), Cambridge: Cambridge University Press.
  • –––, 2003 [2016], “Causal Determinism”, Stanford Encyclopedia of Philosophy (Spring 2016 edition), Edward N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2016/entries/determinism-causal/ >
  • Holbach, Baron, 1770, “The Illusion of Free Will”, from The System of Nature ( Système de la nature ). Translated by H.D. Robinson. Reprinted in The Experience of Philosophy , Daniel Kolak and Raymond Martin (eds.), Oxford: Oxford University Press, 2002, 5th edition: 176–181.
  • Holliday, Wesley H., 2012, “Freedom and the Fixity of the Past”, The Philosophical Review , 121(2): 179–207. doi:10.1215/00318108-1539080
  • –––, 2017, “Freedom and Modality”, in Keller 2017: 149–156. doi:10.1093/acprof:oso/9780198715702.003.0009
  • Holton, Richard, 2009, Willing, Wanting, Waiting , Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199214570.001.0001
  • Horgan, Terence, 1985, “Compatibilism and the Consequence Argument”, Philosophical Studies , 47(3): 339–56. doi:10.1007/BF00355208
  • Horwich, Paul, 1987 , Asymmetries in Time , Cambridge, MA: MIT Press.
  • Huemer, Michael, 2000, “Van Inwagen’s Consequence Argument”, The Philosophical Review , 109(4): 525–544. doi:10.1215/00318108-109-4-525
  • –––, 2004, “Elusive Freedom? A Reply to Helen Beebee”, The Philosophical Review , 113(3): 411–416. doi:10.1215/00318108-113-3-411
  • Huoranszki, Ferenc, 2021, “Physical Determinism, Zygote Manipulation and Responsible Agency”, Philosophia , 49 (4): 1525–1540. doi:10.1007/s11406-020-00307-1
  • Ismael, Jenann, 2016, How Physics Makes Us Free , Oxford: Oxford University Press.
  • Jaster, Romy, 2020, Agents’ Abilities , Berlin: de Gruyter.
  • Jeppsson, Sofia, 2020, “The Agential Perspective: a Hard-Line Reply to the Four-Case Manipulation Argument”, Philosophical Studies , 177(7): 1935–1951. doi: 10.1007/s11098-019-01292-2
  • Kane, Robert, 1996, The Significance of Free Will , Oxford: Oxford University Press. doi:10.1093/0195126564.001.0001
  • –––, 1999, “Responsibility, Luck, and Chance: Reflections on Free Will and Indeterminism”, Journal of Philosophy , 96(5): 217–240. Reprinted in Watson 2003: ch. 15 and Russell & Deery 2013: ch. 10. doi:10.2307/2564666
  • –––, 2011a, “Rethinking Free Will: New Perspectives on an Ancient Problem”, in Kane 2011b: 381–401.
  • ––– (ed.), 2011b,, The Oxford Handbook of Free Will , second edition, Oxford: Oxford University Press. doi:10.1093/oxfordhb/9780195399691.001.0001
  • Kapitan, Tomis, 1991, “Ability and Cognition: a Defense of Compatibilism”, Philosophical Studies , 63(2): 231–43. doi:10.1007/BF00381690
  • –––, 2011, “A Compatibilist Reply to the Consequence Argument”, in Kane 2011b: 131–146.
  • Kearns, Stephen, 2012, “Aborting the Zygote Argument”, Philosophical Studies , 160(3): 379–389. doi:10.1007/s11098-011-9724-3
  • Keller, John A. (ed.), 2017, Being, Freedom, and Method: Themes from the Philosophy of Peter van Inwagen , New York: Oxford University Press. doi:0.1093/acprof:oso/9780198715702.001.0001
  • Kenny, Anthony, 1975, Will, Freedom, and Power , Oxford: Blackwell.
  • Kim, Jaegwon, 1998, Mind in a Physical World: An Essay on the Mind-Body Problem and Mental Causation , Cambridge, MA: Bradford Books, MIT Press.
  • Kratzer, Angelia, 1977, “What ‘must’ and ‘can’ must and can mean”, Linguistics and Philosophy , 1(3): 337–355. doi:10.1007/BF00353453
  • Lamb, James W., 1977, “On a Proof of Incompatibilism”, The Philosophical Review , 86(1): 20–35. doi:10.2307/2184160
  • –––, 1993, “Evaluative Compatibilism and the Principle of Alternative Possibilities”, Journal of Philosophy , 90(1): 497–516. doi:10.2307/2941025
  • Lampert, Fabio and Pedro Merlussi, 2021, “Counterfactuals, Counteractuals, and Free Choice”, Philosophical Studies , 178(2): 445–469. doi:10.1007/s11098-020-01440-z
  • Latham, Andrew J. and Hannah Tierney, 2022, “Defusing Existential and Universal Threats to Compatibilism: A Strawsonian Dilemma for Manipulation Arguments”, Journal of Philosophy , 119(3): 144–161. doi:10.5840/jphil202211939
  • Lehrer, Keith, 1968 “Cans Without Ifs”, Analysis , 29(1): 29–32. doi:10.1093/analys/29.1.29
  • –––, 1976, “’Can’ in Theory and Practice: A Possible Worlds Analysis”, in Brand & Walton 1976.
  • –––, 1980, “Preferences, Conditionals, and Freedom”, in van Inwagen 1980: 187–201. doi:10.1007/978-94-017-3528-5_11
  • –––, 2004, “Freedom and the Power of Preference”, in Campbell, O’Rourke, & Shier 2004: 47–69.
  • Lewis, David, 1973, Counterfactuals , Cambridge, MA: Harvard University Press.
  • –––, 1976, “The Paradoxes of Time Travel”, American Philosophical Quarterly , 13(2): 145–52
  • –––, 1979, “Counterfactual Dependence and Time’s Arrow”, Noûs , 13(4): 455–476. doi:10.2307/2215339
  • –––, 1981, “Are We Free to Break the Laws?”, Theoria , 47(3): 113–21. Reprinted in Watson 2003: ch. 6. doi:10.1111/j.1755-2567.1981.tb00473.x
  • –––, 1986, “Against Overlap”, in his On the Plurality of Worlds , Oxford: Blackwell.
  • Lewis, Peter J., 2016, Quantum Ontology: A Guide to the Metaphysics of Quantum Mechanics , New York: Oxford University Press. doi:10.1093/acprof:oso/9780190469825.001.0001
  • Libet, Benjamin, 1999, “Do We Have Free Will?”, Journal of Consciousness Studies , 6(8–9): 47–57. Reprinted in Russell and Deery 2013, ch. 24 and Sinnott-Armstrong & Nadel 2011: 1–10.
  • List, Christian, 2019, Why Free Will is Real , Cambridge, MA: Harvard University Press.
  • Loew, Christian and Andreas Hüttemann, 2022, “Are We Free to Make the Laws?”, Synthese , 200 (1): 1–16. doi: 10.1007/s11229-022-03460-0
  • Loewer, Barry, 1996a, “Humean Supervenience”, Philosophical Topics , 24(1): 101–27. doi:10.5840/philtopics199624112
  • –––, 1996b, “Freedom from Physics: Quantum Mechanics and Free Will”, Philosophical Topics , 24(2): 91–112. doi:10.5840/philtopics19962428
  • Looper, Brian, 2021, “What Freedom in a Deterministic World Must Be,” Mind , 130(519): 863–885. doi: 10.1093/mind/fzaa105
  • Lycan, William G., 2003, “Free Will and the Burden of Proof”, Royal Institute of Philosophy Supplements , 53: 107–122. doi:10.1017/S1358246100008298
  • Mackie, Penelope, 2003, “Fatalism, Incompatibilism, and the Power to Do Otherwise”, Noûs , 37(4): 672–689. doi:10.1046/j.1468-0068.2003.00455.x
  • –––, 2014a, “Mumford and Anjum on Incompatibilism, Powers, and Determinism”, Analysis , 74(4): 593–603. doi:10.1093/analys/anu088
  • –––, 2014b, “Counterfactuals and the Fixity of the Past”, Philosophical Studies , 168(2): 397–415. doi:10.1007/s11098-013-0135-5
  • Maier, John, 2013, “The Agentive Abilities”, Philosophy and Phenomenological Research , 90(1): 113–134. doi:10.1111/phpr.12038
  • –––, 2018, “Ability, Modality, and Genericity”, Philosophical Studies , 175 (2): 411–428. doi:10.1007/s11098-017-0874-9
  • Mandelkern, Matthew, Ginger Schultheis and David Boylan, 2017, “Agentive Modals”, The Philosophical Review , 126(3): 301–343. doi:10.1215/00318108-3878483
  • Markosian, Ned, 1999, “A Compatibilist Version of the Theory of Agent Causation”, Pacific Philosophical Quarterly , 80(3): 257–277. doi:10.1111/1468-0114.00083
  • McKay, Thomas J. and David Johnson, 1996, “A Reconsideration of an Argument against Compatibilism”, Philosophical Topics , 24(2): 113–122. doi:10.5840/philtopics199624219
  • McKenna, Michael, 2004a, “Responsibility and Globally Manipulated Agents”, Philosophical Topics , 32(1–2): 169–192, John Fischer (ed.). Reprinted in Russell & Deery 2013: ch. 18. doi:10.5840/philtopics2004321/222
  • –––, 2004b [2015], “Compatibilism”, Stanford Encyclopedia of Philosophy (Winter 2015 edition), Edward N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/win2016/entries/compatibilism/ >
  • –––, 2008, “A Hard-line Reply to Pereboom’s Four-Case Manipulation Argument”, Philosophy and Phenomenological Research , 77(1): 142–159. doi:10.1111/j.1933-1592.2008.00179.x
  • –––, 2010, “Whose Argumentative Burden, which Incompatibilist Arguments?—Getting the Dialectic Right”, Australasian Journal of Philosophy , 88(3): 429–443. doi:10.1080/00048400903233811
  • –––, 2014, “Revisiting the Manipulation Argument: A Hard-Liner Takes it on the Chin”, Philosophy and Phenomenological Research , 89 (2), 467–484. doi: 10.1111/phpr.12076
  • Mele, Alfred R., 1995, Autonomous Agents: From Self-Control to Autonomy , New York: Oxford University Press. doi:10.1093/0195150430.001.0001
  • –––, 2006, Free Will and Luck , New York: Oxford University Press. doi:10.1093/0195305043.001.0001
  • –––, 2008, “Manipulation, Compatibilism, and Moral Responsibility”, The Journal of Ethics , 12(3–4): 263–286. doi:10.1007/s10892-008-9035-x
  • –––, 2009, Effective Intentions: The Power of the Conscious Will , New York: Oxford University Press. doi:10.1093/acprof:oso/9780195384260.001.0001
  • ––– (ed.), 2015, Surrounding Free Will: Philosophy, Psychology, Neuroscience , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199333950.001.0001
  • –––, 2019, Manipulated Agents: A Window to Moral Responsibility , New York: Oxford University Press.
  • Mele, Alfred R. and David Robb, 1998, “Rescuing Frankfurt-Style Cases”, The Philosophical Review , 107(1): 97–112. doi:10.2307/2998316
  • Menzies, Peter, 2017, “The Consequence Argument Disarmed: An Interventionist Perspective” in Making a Difference: Essays on the Philosophy of Causation , edited by Helen Beebee, Christopher Hitchcock, and Huw Price, 307–330. Oxford: Oxford University Press.
  • Merricks, Trenton, 2009, “Truth and Freedom”, The Philosophical Review , 118(1): 29–257. doi:10.1215/00318108-2008-028
  • Mickelson, Kristin [Demetriou, Kristin], 2010, “The Soft-Line Solution to Pereboom’s Four-Case Argument”, Australasian Journal of Philosophy , 88(4): 595–617. doi:10.1080/00048400903382691
  • –––, 2015a, “A Critique of Vihvelin’s Three-Fold Classification”, Canadian Journal of Philosophy , 45(1): 85–99. doi:10.1080/00455091.2015.1009321
  • –––, 2015b, “The Zygote Argument is Invalid: Now What?”, Philosophical Studies , 172(11): 2911–2929. doi:10.1007/s11098-015-0449-6
  • Moore, G.E., 1912, “Free Will” in his Ethics , Oxford: Oxford University Press, Chapter 6.
  • Mumford, Stephen and Rani Lill Anjum, 2014, “A New Argument Against Compatibilism”, Analysis , 74(1): 20–25. doi:10.1093/analys/ant095
  • Murray, Dylan and Eddy Nahmias, 2014, “Explaining Away Incompatibilist Intuitions”, Philosophy and Phenomenological Research , 88(2): 434–467. doi:10.1111/j.1933-1592.2012.00609.x
  • Nagel, Thomas, 1979, “Moral Luck”, in his Mortal Questions , New York: Cambridge University Press. Reprinted in Russell & Deery 2013: ch. 1.
  • Nahmias, Eddy, 2004, “The Phenomenology of Free Will”, Journal of Consciousness Studies , 11(7–8): 162–179. Reprinted in Russell & Deery 2013: ch. 25.
  • –––, 2011, “Intuitions about Free Will, Determinism, and Bypassing”, in Kane 2011b: 555–573.
  • Nahmias, Eddy, Stephen Morris, Thomas Nadelhoffer, and Jason Turner, 2006, “Is Incompatibilism Intuitive?”, Philosophy and Phenomenological Research , 73(1): 28–53. doi:10.1111/j.1933-1592.2006.tb00603.x
  • Nelkin, Dana, 2001, “The Consequence Argument and the Mind Argument”, Analysis , 61(2): 107–115. Reprinted in Russell & Deery 2013: ch. 6. doi:10.1093/analys/61.2.107
  • Nelkin, Dana Kay, 2004, “Deliberative Alternatives”, Philosophical Topics , 32(1–2): 215–240. doi:10.5840/philtopics2004321/224
  • –––, 2011, Making Sense of Freedom and Responsibility , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199608560.001.0001
  • Nelkin, Dana K. and Samuel Rickless, 2002, “Warfield’s New Argument for Incompatibilism”, Analysis , 62(2): 104–107. doi:10.1093/analys/62.2.104
  • Oakley, Shane, 2006, “Defending Lewis’s Local Miracle Compatibilism”, Philosophical Studies , 130(2): 337–349. doi:10.1007/s11098-004-4677-4
  • O’Connor, Timothy, 1993, “On the Transfer of Necessity”, Noûs , 27(2): 204–218. doi:10.2307/2215756
  • –––, 1995a, “Agent Causation”, in O’Connor 1995b: 173–200. Reprinted in Watson 2003: ch. 13.
  • –––, 2000, Persons and Causes: the Metaphysics of Free Will , New York: Oxford University Press. doi:10.1093/019515374X.001.0001
  • –––, 2011, “Agent-Causal Theories of Freedom”, in Kane 2011b: 309–327.
  • Pereboom, Derk, 1995, “Determinism al Dente”, Noûs , 29(1): 21–45. Reprinted in Pereboom 2009a: ch. 22. doi:10.2307/2215725
  • –––, 2001, Living Without Free Will , Cambridge: Cambridge University Press.
  • –––, 2003, “Source Incompatibilism and Alternative Possibilities”, in Widerker & McKenna 2003: 185–200.
  • –––, 2005, “Defending Hard Incompatibilism”, Midwest Studies in Philosophy , 29: 228–247. doi:10.1111/j.1475-4975.2005.00114.x
  • –––, 2008, “A Hard-line Reply to the Multiple-Case Manipulation Argument”, Philosophy and Phenomenological Research , 77(1): 160–170. doi: 10.1111/j.1933-1592.2008.00192.x
  • ––– (ed.), 2009a, Free Will , second edition, (Hackett Readings in Philosophy), Indianapolis, IN: Hackett Publishing.
  • –––, 2009b, “Hard Incompatibilism and its Rivals”, Philosophical Studies , 144(1): 21–33. doi:10.1007/s11098-009-9371-0
  • –––, 2014, Free Will, Agency, and Meaning in Life , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199685516.001.0001
  • Pendergraft, Garrett, 2011, “The Explanatory Power of Local Miracle Compatibilism”, Philosophical Studies , 156(2): 249–266. doi:10.1007/s11098-010-9594-0
  • Perry, John, 2004, “Compatibilist Options”, in Campbell, O’Rourke, & Shier 2004: 231–254.
  • –––, 2008, “Can’t We All Just be Compatibilists?: A Critical Study of John Martin Fischer’s My Way ”, The Journal of Ethics , 12(2): 157–166. doi:10.1007/s10892-008-9030-2
  • –––, 2010, “Wretched Subterfuge: A Defense of the Compatibilism of Freedom and Natural Causation”, Dewey Lecture delivered before the 84th Annual Pacific Division meeting, Proceedings and Addresses of the American Philosophical Association , 84(2): 93–113.
  • Pettit Philip and Michael Smith, 1996, “Freedom in Belief and Desire”, Journal of Philosophy , 93(9): 429–449. doi:10.2307/2940892
  • Rosen, Gideon, 2002, “The Case for Incompatibilism”, Philosophy and Phenomenological Research , 64(3): 699–706. doi:10.1111/j.1933-1592.2002.tb00168.x
  • Roskies, Adina and Eddy Nahmias, 2016, “‘Local Determination’, even if we could find it, does not challenge free will: Commentary on Marcelo Fischborn”, Philosophical Psychology , 30(1–2): 185–197. doi:10.1080/09515089.2016.1248286
  • Rummens, Stefan, 2021, “The Counterfactual Structure of the Consequence Argument”, Erkenntnis , 86(3): 523–542. doi:10.1007/s10670-019-00117-2
  • Russell, Paul and Oisin Deery (eds.), 2013, The Philosophy of Free Will, Essential Readings from the Contemporary Debates , Oxford: Oxford University Press.
  • Sartorio, Carolina, 2005, “Causes as Difference-Makers”, Philosophical Studies , 123(1–2): 71–96. doi:10.1007/s11098-004-5217-y
  • –––, 2016, Causation and Free Will , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780198746799.001.0001
  • Scanlon, T.M., 2008, Moral Dimensions: Permissibility, Meaning, Blame , Cambridge, MA: Belknap Press of Harvard University Press.
  • Schaffer, Jonathan, 2004, “Counterfactuals, Causal Independence, and Conceptual Circularity”, Analysis , 64 (4): 299–309. doi:10.1111/j.0003-2638.2004.00501.x
  • –––, 2008, “Causation and Laws of Nature: Reductionism”, in Sider, Hawthorne, & Zimmerman 2008: 82–107.
  • Schlosser, Marcus, 2017, “Traditional Compatibilism Reformulated and Defended”, Journal of Philosophical Research , 42: 277–300. doi:10.5840/jpr2017629108
  • Schwarz, Wolfgang, 2020, “Ability and Possibility”, Philosophers’ Imprint , 20(6): 1–29.
  • Sehon, Scott, 2016, Free Will and Action Explanation , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780198758495.001.0001
  • Shabo, Seth, 2010, “Uncompromising Source Incompatibilism”, Philosophy and Phenomenological Research , 80(2): 349–383. doi:10.1111/j.1933-1592.2010.00328.x
  • Shoemaker, David (ed.), 2013, Oxford Studies in Agency and Responsibility , vol.1, Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199694853.001.0001
  • Sider, Theodore, 2001, Four-Dimensionalism: An Ontology of Persistence and Time , Oxford: Clarendon Press. doi:10.1093/019924443X.001.0001
  • Sider, Theodore, John Hawthorne, and Dean W. Zimmerman (eds.), 2008, Contemporary Debates in Metaphysics , Malden, MA: Blackwell Publishing.
  • Sinnott-Armstrong, Walter and Lynn Nadel, 2011, Conscious Will and Responsibility: A Tribute to Benjamin Libet , New York: Oxford University Press. doi:10.1093/acprof:oso/9780195381641.001.0001
  • Slote, Michael, 1982, “Selective Necessity and the Free Will Problem”, Journal of Philosophy , 79(1): 5–24. doi:10.2307/2026343
  • Smilansky, Saul, 2000, Free Will and Illusion , Oxford: Clarendon Press.
  • Smith, Michael, 1997, “A Theory of Freedom and Responsibility”, in Ethics and Practical Reason , Garrett Cullity & Berys Gaut (eds.), New York: Clarendon Press.
  • –––, 2004, “Rational Capacities”, in his Ethics and the A Priori: Selected Essays on Moral Psychology and Meta-Ethics , New York: Cambridge University Press. doi:10.1017/CBO9780511606977.008
  • Speak, Daniel, 2011, “The Consequence Argument Revisited”, in Kane 2011b: 115–128.
  • Spencer, Jack, 2017, “Able to Do the Impossible”, Mind , 126(502): 466–497. doi:10.1093/mind/fzv183
  • Spitzley, Jay, 2015, “Why Pereboom’s Four-Case Manipulation Argument is Manipulative”, Journal of Cognition and Neuroethics , 3(1): 363–382. [ Spitzley 2015 available online ]
  • Sripada, Chandra Sekhar, 2012, “What Makes a Manipulated Agent Unfree?”, Philosophy and Phenomenological Research , 85(3): 563–593. doi:10.1111/j.1933-1592.2011.00527.x
  • Steward, Helen, 2009, “Fairness, Agency, and the Flicker of Freedom”, Noûs , 43(1): 64–93. doi:10.1111/j.1468-0068.2008.01696.x
  • –––, 2012, A Metaphysics for Freedom , Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199552054.001.0001
  • Strawson, Galen, 1986, Freedom and Belief , Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199247493.001.0001
  • –––, 1994, “The Impossibility of Ultimate Moral Responsibility”, Philosophical Studies , 75(1–2): 5–24. Reprinted in Pereboom 2009a: 21, and Russell & Deery 2013: ch. 19. doi:10.1007/BF00989879
  • Strawson, Peter F., 1962, “Freedom and Resentment”, Proceedings of the British Academy , 48: 187–211. Reprinted in Watson 2003: ch. 4 and Pereboom 2009a: ch. 13.
  • Swartz, Norman, 1986, The Concept of Physical Law , Cambridge: Cambridge University Press.
  • Taylor, Richard, 1962, “Fatalism”, The Philosophical Review , 71(1): 56–66. doi:10.2307/2183681
  • Todd, Patrick, 2013, “Defending a (modified version of) the Zygote Argument”, Philosophical Studies , 164(1): 189–203. doi:10.1007/s11098-011-9848-5
  • Todd, Patrick, 2017, “Manipulation Arguments and the Freedom to do Otherwise”, Philosophy and Phenomenological Research , 95(2): 395–407. doi:10.1111/phpr.12298
  • Tognazzini, Neal A., 2014, “The Structure of a Manipulation Argument”, Ethics , 124(2): 358–369. doi:10.1086/673434
  • Tognazzini, Neal A. and John Martin Fischer, 2017, “Incompatibilism and the Fixity of the Past”, in Keller 2017: 140–148. doi:10.1093/acprof:oso/9780198715702.003.0008
  • Tomberlin, James (ed.), 2000, Action and Freedom , ( Philosophical Perspectives , a supplement to Noûs , vol. 14), Malden, MA: Blackwell Publishing.
  • Van Cleve, James, 2019, “Lewis and Taylor as Partners in Sin”, Acta Analytica , 34(2):165–175. doi: 10.1007/s12136-018-0367-2
  • van Inwagen, Peter, 1975, “The Incompatibility of Free Will and Determinism”, Philosophical Studies , 27(3): 185–199. Reprinted in Pereboom 2009a: ch. 17. doi:10.1007/BF01624156
  • ––– (ed.), 1980, Time and Cause: Essays Presented to Richard Taylor , Dordrecht: Reidel. doi:10.1007/978-94-017-3528-5
  • –––, 1983, An Essay on Free Will , Oxford: Clarendon Press.
  • –––, 1998, “The Mystery of Metaphysical Freedom”, in van Inwagen & Zimmerman 1998: 365–373.
  • –––, 2000, “Free Will Remains a Mystery”, in Tomberlin 2000: 1–19. Reprinted in van Inwagen 2017.
  • –––, 2004a, “Freedom to Break the Laws”, Midwest Studies in Philosophy , 28: 334–350. doi:10.1111/j.1475-4975.2004.00099. Reprinted in van Inwagen 2017.
  • –––, 2004b, “Van Inwagen on Free Will ”, in Campbell, O’Rourke, & Shier 2004: 213–230. Reprinted in van Inwagen 2017.
  • –––, 2008, “How to Think about the Problem of Free Will”, The Journal of Ethics , 12(3–4): 327–341. doi:10.1007/s10892-008-9038-7. Reprinted in van Inwagen 2017.
  • –––, 2011, “A Promising Argument”, in Kane 2011b: 475–481. Reprinted in van Inwagen 2017.
  • –––, 2017, Thinking About Free Will , Cambridge: Cambridge University Press.
  • van Inwagen, Peter and Dean W. Zimmerman (eds.), 2008, Metaphysics: The Big Questions , Oxford: Blackwell Publishing.
  • Vargas, Manuel, 2013, “How to Solve the Problem of Free Will”, in Russell & Deery 2013: ch. 21.
  • Vihvelin, Kadri, 1988, “The Modal Argument for Incompatibilism”, Philosophical Studies , 53(2): 227–244. doi:10.1007/BF00354642
  • –––, 1990, “Freedom, Necessity, and Laws of Nature as Relations between Universals”, Australasian Journal of Philosophy , 68(4): 371–381. doi: 10.1080/00048409012344381
  • –––, 2000, “Freedom, Foreknowledge, and the Principle of Alternate Possibilities”, Canadian Journal of Philosophy 30(1): 1–24. doi:10.1080/00455091.2000.10717523
  • –––, 2004, “Free Will Demystified: A Dispositional Account”, Philosophical Topics , 32(1–2): 427–450. Reprinted in Russell & Deery 2013: ch. 9. doi:10.5840/philtopics2004321/211
  • –––, 2008, “Compatibilism, Incompatibilism, and Impossibilism”, in Sider, Hawthorne, Zimmerman 2008.
  • –––, 2011, “How to Think about the Free Will/Determinism Problem”, in Campbell, O’Rourke, & Slater 2011: 313–340. doi:10.7551/mitpress/9780262015936.003.0014
  • –––, 2013, Causes, Laws, and Free Will: Why Determinism Doesn’t Matter , New York: Oxford University Press. doi:10.1093/acprof:oso/9780199795185.001.0001
  • –––, 2017, “How the Laws Constrain: Causation, Counterfactuals, and Free Will”, Free Will and Causation, Philosophie de la Connaissance , edited by Claudine Tiercelin, Paris: College de France Press.
  • –––, 2020, “Killing Time Again”, The Monist 103(3): 312–327. doi:10.1093/monist/onaa006
  • Warfield, Ted A., 2000, “Causal Determinism and Human Freedom are Incompatible: A New Argument for Incompatibilism”, in Tomberlin 2000: 167–180.
  • –––, 2003, “Compatibilism and Incompatibilism: Some Arguments”, in The Oxford Handbook of Metaphysics , Michael J. Loux and Dean W. Zimmerman (eds.), Oxford: Oxford University Press, 614–630.
  • Watson, Gary, 1975, “Free Agency”, Journal of Philosophy , 72(8): 205–220. Reprinted in Watson 2003: ch. 17 and Watson 2004: ch. 1. doi:10.2307/2024703
  • –––, 1987, “Free Action and Free Will”, Mind , 96(382): 145–72. Reprinted in Watson 2004: ch. 6. doi:10.1093/mind/XCVI.382.145
  • –––, 1996, “Two Faces of Responsibility”, Philosophical Topics , 24(2): 227–248. Reprinted in Watson 2004: ch. 9. doi:10.5840/philtopics199624222
  • –––, (ed.), 2003, Free Will , second edition, Oxford: Oxford University Press.
  • –––, 2004, Agency and Answerability: Selected Essays , Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199272273.001.0001
  • Wegner, Daniel, 2003, The Illusion of Conscious Will , Cambridge, MA: Bradford Books, MIT Press.
  • Westphal, Jonathan, 2006, “The Future and the Truth-Value Links: A Common Sense View”, Analysis , 66: 1–9. doi:10.1093/analys/66.1.1
  • –––, 2012, “Is There a Modal Fallacy in van Inwagen’s ‘First Formal Argument’?”, Analysis , 72(1): 36–41. doi:10.1093/analys/anr138
  • Whittle, Ann, 2016, “Ceteris Paribus, I Could Have Done Otherwise”, Philosophy and Phenomenological Research , 92(1): 73–85. doi:10.1111/phpr.12111
  • –––, 2021, Freedom and Responsibility in Context , Oxford: Oxford University Press.
  • Wilson, Jessica, 2014, “Hume’s Dictum and the Asymmetry of Counterfactual Dependence”, in Chance and Temporal Asymmetry , edited by Jessica M. Wilson, Oxford: Oxford University Press, 258–279.
  • Wolf, Susan, 1987, “Sanity and the Metaphysics of Responsibility”, in Responsibility, Character, and Emotions , Ferdinand David Schoeman (ed.), Cambridge: Cambridge University Press, 46–62. Reprinted in Russell & Deery 2013: ch. 15.
  • –––, 1990, Freedom within Reason , Oxford: Oxford University Press.
  • Widerker, David, 1987, “On an Argument for Incompatibilism”, Analysis , 47(1): 37–41. doi:10.1093/analys/47.1.37
  • –––, 1995, “Libertarianism and Frankfurt’s Attack on the Principle of Alternative Possibilities”, The Philosophical Review , 104(2): 247–261. Reprinted in Watson 2003: ch. 9. doi:10.2307/2185979
  • Widerker, David and Michael McKenna (eds.), 2003, Moral Responsibility and Alternative Possibilities , Burlington, VT: Ashgate Publishing Company.
  • Wiggins, David, 1973, “Towards a Reasonable Libertarianism”, in Essays on Freedom of Action , Ted Honderich (ed.), London: Routledge & Kegan Paul. Reprinted in Watson 2003: ch. 5.
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.

[Please contact the author with suggestions.]

abilities | action | agency | causation: counterfactual theories of | causation: the metaphysics of | compatibilism | decision theory: causal | determinism: causal | dispositions | fatalism | free will | free will: divine foreknowledge and | incompatibilism: (nondeterministic) theories of free will | laws of nature | moral responsibility | time

Copyright © 2022 by Kadri Vihvelin < vihvelin @ usc . 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

  • Share full article

Advertisement

Supported by

Guest Essay

The Supreme Court Got It Wrong: Abortion Is Not Settled Law

In an black-and-white photo illustration, nine abortion pills are arranged on a grid.

By Melissa Murray and Kate Shaw

Ms. Murray is a law professor at New York University. Ms. Shaw is a contributing Opinion writer.

In his majority opinion in the case overturning Roe v. Wade, Justice Samuel Alito insisted that the high court was finally settling the vexed abortion debate by returning the “authority to regulate abortion” to the “people and their elected representatives.”

Despite these assurances, less than two years after Dobbs v. Jackson Women’s Health Organization, abortion is back at the Supreme Court. In the next month, the justices will hear arguments in two high-stakes cases that may shape the future of access to medication abortion and to lifesaving care for pregnancy emergencies. These cases make clear that Dobbs did not settle the question of abortion in America — instead, it generated a new slate of questions. One of those questions involves the interaction of existing legal rules with the concept of fetal personhood — the view, held by many in the anti-abortion movement, that a fetus is a person entitled to the same rights and protections as any other person.

The first case , scheduled for argument on Tuesday, F.D.A. v. Alliance for Hippocratic Medicine, is a challenge to the Food and Drug Administration’s protocols for approving and regulating mifepristone, one of the two drugs used for medication abortions. An anti-abortion physicians’ group argues that the F.D.A. acted unlawfully when it relaxed existing restrictions on the use and distribution of mifepristone in 2016 and 2021. In 2016, the agency implemented changes that allowed the use of mifepristone up to 10 weeks of pregnancy, rather than seven; reduced the number of required in-person visits for dispensing the drug from three to one; and allowed the drug to be prescribed by individuals like nurse practitioners. In 2021, it eliminated the in-person visit requirement, clearing the way for the drug to be dispensed by mail. The physicians’ group has urged the court to throw out those regulations and reinstate the previous, more restrictive regulations surrounding the drug — a ruling that could affect access to the drug in every state, regardless of the state’s abortion politics.

The second case, scheduled for argument on April 24, involves the Emergency Medical Treatment and Labor Act (known by doctors and health policymakers as EMTALA ), which requires federally funded hospitals to provide patients, including pregnant patients, with stabilizing care or transfer to a hospital that can provide such care. At issue is the law’s interaction with state laws that severely restrict abortion, like an Idaho law that bans abortion except in cases of rape or incest and circumstances where abortion is “necessary to prevent the death of the pregnant woman.”

Although the Idaho law limits the provision of abortion care to circumstances where death is imminent, the federal government argues that under EMTALA and basic principles of federal supremacy, pregnant patients experiencing emergencies at federally funded hospitals in Idaho are entitled to abortion care, even if they are not in danger of imminent death.

These cases may be framed in the technical jargon of administrative law and federal pre-emption doctrine, but both cases involve incredibly high-stakes issues for the lives and health of pregnant persons — and offer the court an opportunity to shape the landscape of abortion access in the post-Roe era.

These two cases may also give the court a chance to seed new ground for fetal personhood. Woven throughout both cases are arguments that gesture toward the view that a fetus is a person.

If that is the case, the legal rules that would typically hold sway in these cases might not apply. If these questions must account for the rights and entitlements of the fetus, the entire calculus is upended.

In this new scenario, the issue is not simply whether EMTALA’s protections for pregnant patients pre-empt Idaho’s abortion ban, but rather which set of interests — the patient’s or the fetus’s — should be prioritized in the contest between state and federal law. Likewise, the analysis of F.D.A. regulatory protocols is entirely different if one of the arguments is that the drug to be regulated may be used to end a life.

Neither case presents the justices with a clear opportunity to endorse the notion of fetal personhood — but such claims are lurking beneath the surface. The Idaho abortion ban is called the Defense of Life Act, and in its first bill introduced in 2024, the Idaho Legislature proposed replacing the term “fetus” with “preborn child” in existing Idaho law. In its briefs before the court, Idaho continues to beat the drum of fetal personhood, insisting that EMTALA protects the unborn — rather than pregnant women who need abortions during health emergencies.

According to the state, nothing in EMTALA imposes an obligation to provide stabilizing abortion care for pregnant women. Rather, the law “actually requires stabilizing treatment for the unborn children of pregnant women.” In the mifepristone case, advocates referred to fetuses as “unborn children,” while the district judge in Texas who invalidated F.D.A. approval of the drug described it as one that “starves the unborn human until death.”

Fetal personhood language is in ascent throughout the country. In a recent decision , the Alabama Supreme Court allowed a wrongful-death suit for the destruction of frozen embryos intended for in vitro fertilization, or I.V.F. — embryos that the court characterized as “extrauterine children.”

Less discussed but as worrisome is a recent oral argument at the Florida Supreme Court concerning a proposed ballot initiative intended to enshrine a right to reproductive freedom in the state’s Constitution. In considering the proposed initiative, the chief justice of the state Supreme Court repeatedly peppered Nathan Forrester, the senior deputy solicitor general who was representing the state, with questions about whether the state recognized the fetus as a person under the Florida Constitution. The point was plain: If the fetus was a person, then the proposed ballot initiative, and its protections for reproductive rights, would change the fetus’s rights under the law, raising constitutional questions.

As these cases make clear, the drive toward fetal personhood goes beyond simply recasting abortion as homicide. If the fetus is a person, any act that involves reproduction may implicate fetal rights. Fetal personhood thus has strong potential to raise questions about access to abortion, contraception and various forms of assisted reproductive technology, including I.V.F.

In response to the shifting landscape of reproductive rights, President Biden has pledged to “restore Roe v. Wade as the law of the land.” Roe and its successor, Planned Parenthood v. Casey, were far from perfect; they afforded states significant leeway to impose onerous restrictions on abortion, making meaningful access an empty promise for many women and families of limited means. But the two decisions reflected a constitutional vision that, at least in theory, protected the liberty to make certain intimate choices — including choices surrounding if, when and how to become a parent.

Under the logic of Roe and Casey, the enforceability of EMTALA, the F.D.A.’s power to regulate mifepristone and access to I.V.F. weren’t in question. But in the post-Dobbs landscape, all bets are off. We no longer live in a world in which a shared conception of constitutional liberty makes a ban on I.V.F. or certain forms of contraception beyond the pale.

Melissa Murray, a law professor at New York University and a host of the Supreme Court podcast “ Strict Scrutiny ,” is a co-author of “ The Trump Indictments : The Historic Charging Documents With Commentary.”

Kate Shaw is a contributing Opinion writer, a professor of law at the University of Pennsylvania Carey Law School and a host of the Supreme Court podcast “Strict Scrutiny.” She served as a law clerk to Justice John Paul Stevens and Judge Richard Posner.

IMAGES

  1. Free Will

    thesis about free will

  2. Free Will Essay

    thesis about free will

  3. free will

    thesis about free will

  4. Problem of free will Essay Example

    thesis about free will

  5. Kostenloses Magic Thesis Statement

    thesis about free will

  6. Free Will Essay

    thesis about free will

VIDEO

  1. SHAUN LEE / GBPAUD

  2. Nadiem Nakarim’s Thesis-Free Graduation Policy

  3. Best Website For Thesis Research Paper || Useful Websites For Research Paper || Help For Thesis

  4. Determinism vs Free Will #philosophy #determinism #freewill

  5. Nietzsche Unveiled: The Truth about Free Will & Determinism

  6. The Problem of Free Will (And How To Solve It)

COMMENTS

  1. 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 ...

  2. Free Will

    Free Will and Determinism a. The Thesis of Causal Determinism. Most contemporary scholarship on free will focuses on whether or not it is compatible with causal determinism. Causal determinism is sometimes also called "nomological determinism." ... Free Will: A Philosophical Study (HarperCollins Publishers). Finch, Alicia and Ted Warfield ...

  3. Foreknowledge and Free Will

    Foreknowledge and Free Will. Fatalism is the thesis that human acts occur by necessity and hence are unfree. Theological fatalism is the thesis that infallible foreknowledge of a human act makes the act necessary and hence unfree. If there is a being who knows the entire future infallibly, then no human act is free.

  4. Compatibilism

    Compatibilism is the thesis that free will is compatible with determinism. Because free will is typically taken to be a necessary condition of moral responsibility, compatibilism is sometimes expressed as a thesis about the compatibility between moral responsibility and determinism. 1. Free Will and the Problem of Causal Determinism.

  5. PDF Introduction to the Problem of Free Will and Divine Causality Aquinas

    Moral freedom is as contentious a notion as metaphysical freedom is. In classical moral theory, moral freedom is the telos or end of metaphysical freedom and goes by names like freedom for excellence or moral self-possession or, with Socrates, freedom from enslavement to the passions—or, in the Gospels, freedom from sin.

  6. Why Free Will Is Real

    In Why Free Will Is Real, List does as advertised, advancing a novel, intriguing view of free will and making a thoughtful case for the thesis that free will, as he conceives of it, is real. This book is a pleasure to read. An original and challenging new contribution to contemporary debates about free will. After making a compelling case for ...

  7. Free Will, Determinism, and Moral Responsibility

    The first half of this thesis is a survey of the PSR, followed by consideration of arguments for and against the principle. This survey spans from the Ancient Greeks to the present day, and gives the reader a sense of the ways in which the PSR has been used both implicitly and explicitly throughout the history of philosophy. I argue that, while none of the arguments either for or against the ...

  8. Free Will (Philosophy)

    The former believe that free will and causal determinism are incompatible and that free will exists and the thesis of causal determinism is false. (To be sure, libertarians need not hold that causal determinism never applies to physical objects. They can maintain that humans have special causal powers, over which they have control, that at ...

  9. 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 ...

  10. The problem of free will and determinism

    The philosophical problem of free will and determinism is the problem of whether or not free will exists in light of determinism. Thus, it is crucial to be clear in defining what we mean by "free will" and "determinism.". As we will see, these turn out to be difficult and contested philosophical questions.

  11. Free Will and Theism: Connections, Contingencies, and Concerns

    The next two essays consider how creaturely free agency can be compatible with classic Christian teaching about God's intimate causal involvement in creation. W. Matthews Grant's concern is with the "doctrine of divine universal causality," on which God directly causes everything -- including human actions -- to exist as long as they exist. ...

  12. 4 Chaos, Indeterminism, and Free Will

    Because the question of how physical laws are related to the exercise of free will appears explicitly in free-will discussions (e.g., see the essays in this volume on the Consequence Argument; Kane 1996), my primary focus will be on physical determinism, the thesis that all physical events are determined to occur according to physical laws.This choice of focus only removes part of the ...

  13. Free will

    Free will is the capacity or ability to choose between different possible courses of action unimpeded. ... "theological determinism is the thesis that God exists and has infallible knowledge of all true propositions including propositions about our future actions," more minimal criteria designed to encapsulate all forms of theological determinism.

  14. Free Will

    Free Will. First published Mon Jan 7, 2002; substantive revision Thu Apr 14, 2005. "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.

  15. Psychoanalysis and Free Will

    ABSTRACT. This article is mainly concerned with the conception of free will in Freudian theory and ego psychology. There are a number of Freuds, not all consistent with each other, on this issue of free will: the Freud who views free will as an illusion, the Freud who identifies as a goal of psychoanalytic treatment the enhancement of the ego's freedom to choose, and the Freud who locates ...

  16. Our Fate: Essays on God and Free Will

    Our Fate: Essays on God and Free Will; Our Fate: Essays on God and Free Will John Martin Fischer, Our Fate: Essays on God and Free Will, Oxford University Press, 2016, 243pp., $74.00 (hbk), ISBN 9780199311293. Reviewed by . Martijn Boot, Waseda University. 2016.05.25.

  17. My Go-To Arguments for Free Will

    By free will, I mean a capacity for deliberate, conscious decisions. Choices. Free will is variable. The more choices you have, the more free will you have. Our choices are constrained by all ...

  18. Friedrich Nietzsche and free will

    Friedrich Nietzsche and free will. The 19th-century philosopher Friedrich Nietzsche is known as a critic of Judeo-Christian morality and religions in general. One of the arguments he raised against the truthfulness of these doctrines is that they are based upon the concept of free will, which, in his opinion, does not exist.

  19. Free will

    Jean-Paul Sartre (1905-80), for example, spoke of the individual "condemned to be free." The existence of free will is denied by some proponents of determinism, the thesis that every event in the universe is causally inevitable. Determinism entails that, in a situation in which people make a certain decision or perform a certain action ...

  20. Free Will Essay example

    Free Will Essay example. I want to argue that there is indeed free will. In order to defend the position that free will means that human beings can cause some of what they do on their own; in other words, what they do is not explainable solely by references to factors that have influenced them. My thesis then, is that human beings are able to ...

  21. Sample Essay on Free Will and Moral Responsibility

    Ultius. 17 May 2014. 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 ...

  22. Arguments for Incompatibilism

    The free will thesis is a minimal claim about free will; it would be true if one person in the universe acted with free will (acted freely, acted while possessing free will) on one occasion. We won't assume that the free will thesis is true or even possibly true, but let a free will world be any possible world where the free will thesis is ...

  23. (PDF) General Thesis on Free Will

    December 1999 · Israel Journal of Mathematics. H E A Eddy Campbell. N. E. Kechagias. In his PhD thesis, Arnon [1] builds a completion of the Dickson algebras which contains a "free root ...

  24. OATD

    Theses and dissertations, free to find, free to use. October 3, 2022. OATD is dealing with a number of misbehaved crawlers and robots, and is currently taking some steps to minimize their impact on the system. This may require you to click through some security screen. Our apologies for any inconvenience.

  25. Why Abortion Is Back at the Supreme Court

    Wade, Justice Samuel Alito insisted that the high court was finally settling the vexed abortion debate by returning the "authority to regulate abortion" to the "people and their elected ...