nathalie lees free will/illusion of choice illustration

The clockwork universe: is free will an illusion?

A growing chorus of scientists and philosophers argue that free will does not exist. Could they be right?

T owards the end of a conversation dwelling on some of the deepest metaphysical puzzles regarding the nature of human existence, the philosopher Galen Strawson paused, then asked me: “Have you spoken to anyone else yet who’s received weird email?” He navigated to a file on his computer and began reading from the alarming messages he and several other scholars had received over the past few years. Some were plaintive, others abusive, but all were fiercely accusatory. “Last year you all played a part in destroying my life,” one person wrote. “I lost everything because of you – my son, my partner, my job, my home, my mental health. All because of you, you told me I had no control, how I was not responsible for anything I do, how my beautiful six-year-old son was not responsible for what he did … Goodbye, and good luck with the rest of your cancerous, evil, pathetic existence.” “Rot in your own shit Galen,” read another note, sent in early 2015. “Your wife, your kids your friends, you have smeared all there [sic] achievements you utter fucking prick,” wrote the same person, who subsequently warned: “I’m going to fuck you up.” And then, days later, under the subject line “Hello”: “I’m coming for you.” “This was one where we had to involve the police,” Strawson said. Thereafter, the violent threats ceased.

It isn’t unheard of for philosophers to receive death threats. The Australian ethicist Peter Singer, for example, has received many, in response to his argument that, in highly exceptional circumstances, it might be morally justifiable to kill newborn babies with severe disabilities. But Strawson, like others on the receiving end of this particular wave of abuse, had merely expressed a longstanding position in an ancient debate that strikes many as the ultimate in “armchair philosophy”, wholly detached from the emotive entanglements of real life. They all deny that human beings possess free will. They argue that our choices are determined by forces beyond our ultimate control – perhaps even predetermined all the way back to the big bang – and that therefore nobody is ever wholly responsible for their actions. Reading back over the emails, Strawson, who gives the impression of someone far more forgiving of other people’s flaws than of his own, found himself empathising with his harassers’ distress. “I think for these people it’s just an existential catastrophe,” he said. “And I think I can see why.”

The difficulty in explaining the enigma of free will to those unfamiliar with the subject isn’t that it’s complex or obscure. It’s that the experience of possessing free will – the feeling that we are the authors of our choices – is so utterly basic to everyone’s existence that it can be hard to get enough mental distance to see what’s going on. Suppose you find yourself feeling moderately hungry one afternoon, so you walk to the fruit bowl in your kitchen, where you see one apple and one banana. As it happens, you choose the banana. But it seems absolutely obvious that you were free to choose the apple – or neither, or both – instead. That’s free will: were you to rewind the tape of world history, to the instant just before you made your decision, with everything in the universe exactly the same, you’d have been able to make a different one.

Nothing could be more self-evident. And yet according to a growing chorus of philosophers and scientists, who have a variety of different reasons for their view, it also can’t possibly be the case. “This sort of free will is ruled out, simply and decisively, by the laws of physics,” says one of the most strident of the free will sceptics, the evolutionary biologist Jerry Coyne. Leading psychologists such as Steven Pinker and Paul Bloom agree, as apparently did the late Stephen Hawking, along with numerous prominent neuroscientists, including VS Ramachandran, who called free will “an inherently flawed and incoherent concept” in his endorsement of Sam Harris’s bestselling 2012 book Free Will, which also makes that argument. According to the public intellectual Yuval Noah Harari, free will is an anachronistic myth – useful in the past, perhaps, as a way of motivating people to fight against tyrants or oppressive ideologies, but rendered obsolete by the power of modern data science to know us better than we know ourselves, and thus to predict and manipulate our choices.

Arguments against free will go back millennia, but the latest resurgence of scepticism has been driven by advances in neuroscience during the past few decades. Now that it’s possible to observe – thanks to neuroimaging – the physical brain activity associated with our decisions, it’s easier to think of those decisions as just another part of the mechanics of the material universe, in which “free will” plays no role. And from the 1980s onwards, various specific neuroscientific findings have offered troubling clues that our so-called free choices might actually originate in our brains several milliseconds, or even much longer, before we’re first aware of even thinking of them.

Despite the criticism that this is all just armchair philosophy, the truth is that the stakes could hardly be higher. Were free will to be shown to be nonexistent – and were we truly to absorb the fact – it would “precipitate a culture war far more belligerent than the one that has been waged on the subject of evolution”, Harris has written. Arguably, we would be forced to conclude that it was unreasonable ever to praise or blame anyone for their actions, since they weren’t truly responsible for deciding to do them; or to feel guilt for one’s misdeeds, pride in one’s accomplishments, or gratitude for others’ kindness. And we might come to feel that it was morally unjustifiable to mete out retributive punishment to criminals, since they had no ultimate choice about their wrongdoing. Some worry that it might fatally corrode all human relations, since romantic love, friendship and neighbourly civility alike all depend on the assumption of choice: any loving or respectful gesture has to be voluntary for it to count.

Peer over the precipice of the free will debate for a while, and you begin to appreciate how an already psychologically vulnerable person might be nudged into a breakdown, as was apparently the case with Strawson’s email correspondents. Harris has taken to prefacing his podcasts on free will with disclaimers, urging those who find the topic emotionally distressing to give them a miss. And Saul Smilansky, a professor of philosophy at the University of Haifa in Israel, who believes the popular notion of free will is a mistake, told me that if a graduate student who was prone to depression sought to study the subject with him, he would try to dissuade them. “Look, I’m naturally a buoyant person,” he said. “I have the mentality of a village idiot: it’s easy to make me happy. Nevertheless, the free will problem is really depressing if you take it seriously. It hasn’t made me happy, and in retrospect, if I were at graduate school again, maybe a different topic would have been preferable.”

Smilansky is an advocate of what he calls “illusionism”, the idea that although free will as conventionally defined is unreal, it’s crucial people go on believing otherwise – from which it follows that an article like this one might be actively dangerous. (Twenty years ago, he said, he might have refused to speak to me, but these days free will scepticism was so widely discussed that “the horse has left the barn”.) “On the deepest level, if people really understood what’s going on – and I don’t think I’ve fully internalised the implications myself, even after all these years – it’s just too frightening and difficult,” Smilansky said. “For anyone who’s morally and emotionally deep, it’s really depressing and destructive. It would really threaten our sense of self, our sense of personal value. The truth is just too awful here.”

T he conviction that nobody ever truly chooses freely to do anything – that we’re the puppets of forces beyond our control – often seems to strike its adherents early in their intellectual careers, in a sudden flash of insight. “I was sitting in a carrel in Wolfson College [in Oxford] in 1975, and I had no idea what I was going to write my DPhil thesis about,” Strawson recalled. “I was reading something about Kant’s views on free will, and I was just electrified. That was it.” The logic, once glimpsed, seems coldly inexorable. Start with what seems like an obvious truth: anything that happens in the world, ever, must have been completely caused by things that happened before it. And those things must have been caused by things that happened before them – and so on, backwards to the dawn of time: cause after cause after cause, all of them following the predictable laws of nature, even if we haven’t figured all of those laws out yet. It’s easy enough to grasp this in the context of the straightforwardly physical world of rocks and rivers and internal combustion engines. But surely “one thing leads to another” in the world of decisions and intentions, too. Our decisions and intentions involve neural activity – and why would a neuron be exempt from the laws of physics any more than a rock?

So in the fruit bowl example, there are physiological reasons for your feeling hungry in the first place, and there are causes – in your genes, your upbringing, or your current environment – for your choosing to address your hunger with fruit, rather than a box of doughnuts. And your preference for the banana over the apple, at the moment of supposed choice, must have been caused by what went before, presumably including the pattern of neurons firing in your brain, which was itself caused – and so on back in an unbroken chain to your birth, the meeting of your parents, their births and, eventually, the birth of the cosmos.

An astronomical clock in Prague, Czech Republic.

But if all that’s true, there’s simply no room for the kind of free will you might imagine yourself to have when you see the apple and banana and wonder which one you’ll choose. To have what’s known in the scholarly jargon as “contra-causal” free will – so that if you rewound the tape of history back to the moment of choice, you could make a different choice – you’d somehow have to slip outside physical reality. To make a choice that wasn’t merely the next link in the unbroken chain of causes, you’d have to be able to stand apart from the whole thing, a ghostly presence separate from the material world yet mysteriously still able to influence it. But of course you can’t actually get to this supposed place that’s external to the universe, separate from all the atoms that comprise it and the laws that govern them. You just are some of the atoms in the universe, governed by the same predictable laws as all the rest.

It was the French polymath Pierre-Simon Laplace, writing in 1814, who most succinctly expressed the puzzle here: how can there be free will, in a universe where events just crank forwards like clockwork? His thought experiment is known as Laplace’s demon, and his argument went as follows: if some hypothetical ultra-intelligent being – or demon – could somehow know the position of every atom in the universe at a single point in time, along with all the laws that governed their interactions, it could predict the future in its entirety. There would be nothing it couldn’t know about the world 100 or 1,000 years hence, down to the slightest quiver of a sparrow’s wing. You might think you made a free choice to marry your partner, or choose a salad with your meal rather than chips; but in fact Laplace’s demon would have known it all along, by extrapolating out along the endless chain of causes. “For such an intellect,” Laplace said, “nothing could be uncertain, and the future, just like the past, would be present before its eyes.”

It’s true that since Laplace’s day, findings in quantum physics have indicated that some events, at the level of atoms and electrons, are genuinely random, which means they would be impossible to predict in advance, even by some hypothetical megabrain. But few people involved in the free will debate think that makes a critical difference. Those tiny fluctuations probably have little relevant impact on life at the scale we live it, as human beings. And in any case, there’s no more freedom in being subject to the random behaviours of electrons than there is in being the slave of predetermined causal laws. Either way, something other than your own free will seems to be pulling your strings.

B y far the most unsettling implication of the case against free will, for most who encounter it, is what it seems to say about morality: that nobody, ever, truly deserves reward or punishment for what they do, because what they do is the result of blind deterministic forces (plus maybe a little quantum randomness). “For the free will sceptic,” writes Gregg Caruso in his new book Just Deserts, a collection of dialogues with his fellow philosopher Daniel Dennett, “it is never fair to treat anyone as morally responsible.” Were we to accept the full implications of that idea, the way we treat each other – and especially the way we treat criminals – might change beyond recognition.

Consider the case of Charles Whitman. Just after midnight on 1 August 1966, Whitman – an outgoing and apparently stable 25-year-old former US Marine – drove to his mother’s apartment in Austin, Texas, where he stabbed her to death. He returned home, where he killed his wife in the same manner. Later that day, he took an assortment of weapons to the top of a high building on the campus of the University of Texas, where he began shooting randomly for about an hour and a half. By the time Whitman was killed by police, 12 more people were dead, and one more died of his injuries years afterwards – a spree that remains the US’s 10th worst mass shooting.

Within hours of the massacre, the authorities discovered a note that Whitman had typed the night before. “I don’t quite understand what compels me to type this letter,” he wrote. “Perhaps it is to leave some vague reason for the actions I have recently performed. I don’t really understand myself these days. I am supposed to be an average reasonable and intelligent young man. However, lately (I can’t recall when it started) I have been a victim of many unusual and irrational thoughts [which] constantly recur, and it requires a tremendous mental effort to concentrate on useful and progressive tasks … After my death I wish that an autopsy would be performed to see if there is any visible physical disorder.” Following the first two murders, he added a coda: “Maybe research can prevent further tragedies of this type.” An autopsy was performed, revealing the presence of a substantial brain tumour, pressing on Whitman’s amygdala, the part of the brain governing “fight or flight” responses to fear.

As the free will sceptics who draw on Whitman’s case concede, it’s impossible to know if the brain tumour caused Whitman’s actions. What seems clear is that it certainly could have done so – and that almost everyone, on hearing about it, undergoes some shift in their attitude towards him. It doesn’t make the killings any less horrific. Nor does it mean the police weren’t justified in killing him. But it does make his rampage start to seem less like the evil actions of an evil man, and more like the terrible symptom of a disorder, with Whitman among its victims. The same is true for another wrongdoer famous in the free-will literature, the anonymous subject of the 2003 paper Right Orbitofrontal Tumor with Paedophilia Symptom and Constructional Apraxia Sign, a 40-year-old schoolteacher who suddenly developed paedophilic urges and began seeking out child pornography, and was subsequently convicted of child molestation. Soon afterwards, complaining of headaches, he was diagnosed with a brain tumour; when it was removed, his paedophilic urges vanished. A year later, they returned – as had his tumour, detected in another brain scan.

If you find the presence of a brain tumour in these cases in any way exculpatory, though, you face a difficult question: what’s so special about a brain tumour, as opposed to all the other ways in which people’s brains cause them to do things? When you learn about the specific chain of causes that were unfolding inside Charles Whitman’s skull, it has the effect of seeming to make him less personally responsible for the terrible acts he committed. But by definition, anyone who commits any immoral act has a brain in which a chain of prior causes had unfolded, leading to the act; if that weren’t the case, they’d never have committed the act. “A neurological disorder appears to be just a special case of physical events giving rise to thoughts and actions,” is how Harris expresses it. “Understanding the neurophysiology of the brain, therefore, would seem to be as exculpatory as finding a tumour in it.” It appears to follow that as we understand ever more about how the brain works, we’ll illuminate the last shadows in which something called “free will” might ever have lurked – and we’ll be forced to concede that a criminal is merely someone unlucky enough to find himself at the end of a causal chain that culminates in a crime. We can still insist the crime in question is morally bad; we just can’t hold the criminal individually responsible. (Or at least that’s where the logic seems to lead our modern minds: there’s a rival tradition , going back to the ancient Greeks, which holds that you can be held responsible for what’s fated to happen to you anyway.)

Illustration for Guardian long read 27 April 2021

For Caruso, who teaches philosophy at the State University of New York, what all this means is that retributive punishment – punishing a criminal because he deserves it, rather than to protect the public, or serve as a warning to others – can’t ever be justified. Like Strawson, he has received email abuse from people disturbed by the implications. Retribution is central to all modern systems of criminal justice, yet ultimately, Caruso thinks, “it’s a moral injustice to hold someone responsible for actions that are beyond their control. It’s capricious.” Indeed some psychological research, he points out, suggests that people believe in free will partly because they want to justify their appetite for retribution. “What seems to happen is that people come across an action they disapprove of; they have a high desire to blame or punish; so they attribute to the perpetrator the degree of control [over their own actions] that would be required to justify blaming them.” (It’s no accident that the free will controversy is entangled in debates about religion: following similar logic, sinners must freely choose to sin, in order for God’s retribution to be justified.)

Caruso is an advocate of what he calls the “public health-quarantine” model of criminal justice, which would transform the institutions of punishment in a radically humane direction. You could still restrain a murderer, on the same rationale that you can require someone infected by Ebola to observe a quarantine: to protect the public. But you’d have no right to make the experience any more unpleasant than was strictly necessary for public protection. And you would be obliged to release them as soon as they no longer posed a threat. (The main focus, in Caruso’s ideal world, would be on redressing social problems to try stop crime happening in the first place – just as public health systems ought to focus on preventing epidemics happening to begin with.)

It’s tempting to try to wriggle out of these ramifications by protesting that, while people might not choose their worst impulses – for murder, say – they do have the choice not to succumb to them. You can feel the urge to kill someone but resist it, or even seek psychiatric help. You can take responsibility for the state of your personality. And don’t we all do that, all the time, in more mundane ways, whenever we decide to acquire a new professional skill, become a better listener, or finally get fit?

But this is not the escape clause it might seem. After all, the free will sceptics insist, if you do manage to change your personality in some admirable way, you must already have possessed the kind of personality capable of implementing such a change – and you didn’t choose that . None of this requires us to believe that the worst atrocities are any less appalling than we previously thought. But it does entail that the perpetrators can’t be held personally to blame. If you’d been born with Hitler’s genes, and experienced Hitler’s upbringing, you would be Hitler – and ultimately it’s only good fortune that you weren’t. In the end, as Strawson puts it, “luck swallows everything”.

G iven how watertight the case against free will can appear, it may be surprising to learn that most philosophers reject it: according to a 2009 survey, conducted by the website PhilPapers, only about 12% of them are persuaded by it. And the disagreement can be fraught, partly because free will denial belongs to a wider trend that drives some philosophers spare – the tendency for those trained in the hard sciences to make sweeping pronouncements about debates that have raged in philosophy for years, as if all those dull-witted scholars were just waiting for the physicists and neuroscientists to show up. In one chilly exchange, Dennett paid a backhanded compliment to Harris, who has a PhD in neuroscience, calling his book “remarkable” and “valuable” – but only because it was riddled with so many wrongheaded claims: “I am grateful to Harris for saying, so boldly and clearly, what less outgoing scientists are thinking but keeping to themselves.”

What’s still more surprising, and hard to wrap one’s mind around, is that most of those who defend free will don’t reject the sceptics’ most dizzying assertion – that every choice you ever make might have been determined in advance. So in the fruit bowl example, a majority of philosophers agree that if you rewound the tape of history to the moment of choice, with everything in the universe exactly the same, you couldn’t have made a different selection. That kind of free will is “as illusory as poltergeists”, to quote Dennett. What they claim instead is that this doesn’t matter: that even though our choices may be determined, it makes sense to say we’re free to choose. That’s why they’re known as “compatibilists”: they think determinism and free will are compatible. (There are many other positions in the debate, including some philosophers, many Christians among them, who think we really do have “ghostly” free will; and others who think the whole so-called problem is a chimera, resulting from a confusion of categories, or errors of language.)

To those who find the case against free will persuasive, compatibilism seems outrageous at first glance. How can we possibly be free to choose if we aren’t, in fact, you know, free to choose? But to grasp the compatibilists’ point, it helps first to think about free will not as a kind of magic, but as a mundane sort of skill – one which most adults possess, most of the time. As the compatibilist Kadri Vihvelin writes, “we have the free will we think we have, including the freedom of action we think we have … by having some bundle of abilities and being in the right kind of surroundings.” The way most compatibilists see things, “being free” is just a matter of having the capacity to think about what you want, reflect on your desires, then act on them and sometimes get what you want. When you choose the banana in the normal way – by thinking about which fruit you’d like, then taking it – you’re clearly in a different situation from someone who picks the banana because a fruit-obsessed gunman is holding a pistol to their head; or someone afflicted by a banana addiction, compelled to grab every one they see. In all of these scenarios, to be sure, your actions belonged to an unbroken chain of causes, stretching back to the dawn of time. But who cares? The banana-chooser in one of them was clearly more free than in the others.

“Harris, Pinker, Coyne – all these scientists, they all make the same two-step move,” said Eddy Nahmias, a compatibilist philosopher at Georgia State University in the US. “Their first move is always to say, ‘well, here’s what free will means’” – and it’s always something nobody could ever actually have, in the reality in which we live. “And then, sure enough, they deflate it. But once you have that sort of balloon in front of you, it’s very easy to deflate it, because any naturalistic account of the world will show that it’s false.”

Daniel Dennett in Stockholm, Sweden.

Consider hypnosis. A doctrinaire free will sceptic might feel obliged to argue that a person hypnotised into making a particular purchase is no less free than someone who thinks about it, in the usual manner, before reaching for their credit card. After all, their idea of free will requires that the choice wasn’t fully determined by prior causes; yet in both cases, hypnotised and non-hypnotised, it was. “But come on, that’s just really annoying ,” said Helen Beebee, a philosopher at the University of Manchester who has written widely on free will, expressing an exasperation commonly felt by compatibilists toward their rivals’ more outlandish claims. “In some sense, I don’t care if you call it ‘free will’ or ‘acting freely’ or anything else – it’s just that it obviously does matter, to everybody, whether they get hypnotised into doing things or not.”

Granted, the compatibilist version of free will may be less exciting. But it doesn’t follow that it’s worthless. Indeed, it may be (in another of Dennett’s phrases) the only kind of “free will worth wanting”. You experience the desire for a certain fruit, you act on it, and you get the fruit, with no external gunmen or internal disorders influencing your choice. How could a person ever be freer than that?

Thinking of free will this way also puts a different spin on some notorious experiments conducted in the 80s by the American neuroscientist Benjamin Libet, which have been interpreted as offering scientific proof that free will doesn’t exist. Wiring his subjects to a brain scanner, and asking them to flex their hands at a moment of their choosing, Libet seemed to show that their choice was detectable from brain activity 300 milliseconds before they made a conscious decision. (Other studies have indicated activity up to 10 seconds before a conscious choice.) How could these subjects be said to have reached their decisions freely, if the lab equipment knew their decisions so far in advance? But to most compatibilists, this is a fuss about nothing. Like everything else, our conscious choices are links in a causal chain of neural processes, so of course some brain activity precedes the moment at which we become aware of them.

From this down-to-earth perspective, there’s also no need to start panicking that cases like Charles Whitman’s might mean we could never hold anybody responsible for their misdeeds, or praise them for their achievements. (In their defence, several free will sceptics I spoke to had their reasons for not going that far, either.) Instead, we need only ask whether someone had the normal ability to choose rationally, reflecting on the implications of their actions. We all agree that newborn babies haven’t developed that yet, so we don’t blame them for waking us in the night; and we believe most non-human animals don’t possess it – so few of us rage indignantly at wasps for stinging us. Someone with a severe neurological or developmental impairment would surely lack it, too, perhaps including Whitman. But as for everyone else: “ Bernie Madoff is the example I always like to use,” said Nahmias. “Because it’s so clear that he knew what he was doing, and that he knew that what he was doing was wrong, and he did it anyway.” He did have the ability we call “free will” – and used it to defraud his investors of more than $17bn.

To the free will sceptics, this is all just a desperate attempt at face-saving and changing the subject – an effort to redefine free will not as the thing we all feel, when faced with a choice, but as something else, unworthy of the name. “People hate the idea that they aren’t agents who can make free choices,” Jerry Coyne has argued. Harris has accused Dennett of approaching the topic as if he were telling someone bent on discovering the lost city of Atlantis that they ought to be satisfied with a trip to Sicily. After all, it meets some of the criteria: it’s an island in the sea, home to a civilisation with ancient roots. But the facts remain: Atlantis doesn’t exist. And when it felt like it wasn’t inevitable you’d choose the banana, the truth is that it actually was.

I t’s tempting to dismiss the free will controversy as irrelevant to real life, on the grounds that we can’t help but feel as though we have free will, whatever the philosophical truth may be. I’m certainly going to keep responding to others as though they had free will: if you injure me, or someone I love, I can guarantee I’m going to be furious, instead of smiling indulgently on the grounds that you had no option. In this experiential sense, free will just seems to be a given.

But is it? When my mind is at its quietest – for example, drinking coffee early in the morning, before the four-year-old wakes up – things are liable to feel different. In such moments of relaxed concentration, it seems clear to me that my intentions and choices, like all my other thoughts and emotions, arise unbidden in my awareness. There’s no sense in which it feels like I’m their author. Why do I put down my coffee mug and head to the shower at the exact moment I do so? Because the intention to do so pops up, caused, no doubt, by all sorts of activity in my brain – but activity that lies outside my understanding, let alone my command. And it’s exactly the same when it comes to those weightier decisions that seem to express something profound about the kind of person I am: whether to attend the funeral of a certain relative, say, or which of two incompatible career opportunities to pursue. I can spend hours or even days engaged in what I tell myself is “reaching a decision” about those, when what I’m really doing, if I’m honest, is just vacillating between options – until at some unpredictable moment, or when an external deadline forces the issue, the decision to commit to one path or another simply arises.

This is what Harris means when he declares that, on close inspection, it’s not merely that free will is an illusion, but that the illusion of free will is itself an illusion: watch yourself closely, and you don’t even seem to be free. “If one pays sufficient attention,” he told me by email, “one can notice that there’s no subject in the middle of experience – there is only experience. And everything we experience simply arises on its own.” This is an idea with roots in Buddhism, and echoed by others, including the philosopher David Hume: when you look within, there’s no trace of an internal commanding officer, autonomously issuing decisions. There’s only mental activity, flowing on. Or as Arthur Rimbaud wrote, in a letter to a friend in 1871: “I am a spectator at the unfolding of my thought; I watch it, I listen to it.”

There are reasons to agree with Saul Smilansky that it might be personally and societally detrimental for too many people to start thinking in this way, even if it turns out it’s the truth. (Dennett, although he thinks we do have free will, takes a similar position, arguing that it’s morally irresponsible to promote free-will denial.) In one set of studies in 2008, the psychologists Kathleen Vohs and Jonathan Schooler asked one group of participants to read an excerpt from The Astonishing Hypothesis by Francis Crick, co-discoverer of the structure of DNA, in which he suggests free will is an illusion. The subjects thus primed to doubt the existence of free will proved significantly likelier than others, in a subsequent stage of the experiment, to cheat in a test where there was money at stake. Other research has reported a diminished belief in free will to less willingness to volunteer to help others, to lower levels of commitment in relationships, and lower levels of gratitude.

Unsuccessful attempts to replicate Vohs and Schooler’s findings have called them into question. But even if the effects are real, some free will sceptics argue that the participants in such studies are making a common mistake – and one that might get cleared up rather rapidly, were the case against free will to become better known and understood. Study participants who suddenly become immoral seem to be confusing determinism with fatalism – the idea that if we don’t have free will, then our choices don’t really matter, so we might as well not bother trying to make good ones, and just do as we please instead. But in fact it doesn’t follow from our choices being determined that they don’t matter. It might matter enormously whether you choose to feed your children a diet rich in vegetables or not; or whether you decide to check carefully in both directions before crossing a busy road. It’s just that (according to the sceptics) you don’t get to make those choices freely.

In any case, were free will really to be shown to be nonexistent, the implications might not be entirely negative. It’s true that there’s something repellent about an idea that seems to require us to treat a cold-blooded murderer as not responsible for his actions, while at the same time characterising the love of a parent for a child as nothing more than what Smilansky calls “the unfolding of the given” – mere blind causation, devoid of any human spark. But there’s something liberating about it, too. It’s a reason to be gentler with yourself, and with others. For those of us prone to being hard on ourselves, it’s therapeutic to keep in the back of your mind the thought that you might be doing precisely as well as you were always going to be doing – that in the profoundest sense, you couldn’t have done any more. And for those of us prone to raging at others for their minor misdeeds, it’s calming to consider how easily their faults might have been yours. (Sure enough, some research has linked disbelief in free will to increased kindness.)

Harris argues that if we fully grasped the case against free will, it would be difficult to hate other people: how can you hate someone you don’t blame for their actions? Yet love would survive largely unscathed, since love is “the condition of our wanting those we love to be happy, and being made happy ourselves by that ethical and emotional connection”, neither of which would be undermined. And countless other positive aspects of life would be similarly untouched. As Strawson puts it, in a world without a belief in free will, “strawberries would still taste just as good”.

Those early-morning moments aside, I personally can’t claim to find the case against free will ultimately persuasive; it’s just at odds with too much else that seems obviously true about life. Yet even if only entertained as a hypothetical possibility, free will scepticism is an antidote to that bleak individualist philosophy which holds that a person’s accomplishments truly belong to them alone – and that you’ve therefore only yourself to blame if you fail. It’s a reminder that accidents of birth might affect the trajectories of our lives far more comprehensively than we realise, dictating not only the socioeconomic position into which we’re born, but also our personalities and experiences as a whole: our talents and our weaknesses, our capacity for joy, and our ability to overcome tendencies toward violence, laziness or despair, and the paths we end up travelling. There is a deep sense of human fellowship in this picture of reality – in the idea that, in our utter exposure to forces beyond our control, we might all be in the same boat, clinging on for our lives, adrift on the storm-tossed ocean of luck.

  • The long read
  • Philosophy (World news)
  • Neuroscience
  • Stephen Hawking
  • Philosophy (Education)

Most viewed

An illustration of someone standing behind an arrow made up of numbered dots that mirror each other, then converge

There’s No Such Thing as Free Will

But we’re better off believing in it anyway.

F or centuries , philosophers and theologians have almost unanimously held that civilization as we know it depends on a widespread belief in free will—and that losing this belief could be calamitous. Our codes of ethics, for example, assume that we can freely choose between right and wrong. In the Christian tradition, this is known as “moral liberty”—the capacity to discern and pursue the good, instead of merely being compelled by appetites and desires. The great Enlightenment philosopher Immanuel Kant reaffirmed this link between freedom and goodness. If we are not free to choose, he argued, then it would make no sense to say we ought to choose the path of righteousness.

Today, the assumption of free will runs through every aspect of American politics, from welfare provision to criminal law. It permeates the popular culture and underpins the American dream—the belief that anyone can make something of themselves no matter what their start in life. As Barack Obama wrote in The Audacity of Hope , American “values are rooted in a basic optimism about life and a faith in free will.”

So what happens if this faith erodes?

The sciences have grown steadily bolder in their claim that all human behavior can be explained through the clockwork laws of cause and effect. This shift in perception is the continuation of an intellectual revolution that began about 150 years ago, when Charles Darwin first published On the Origin of Species . Shortly after Darwin put forth his theory of evolution, his cousin Sir Francis Galton began to draw out the implications: If we have evolved, then mental faculties like intelligence must be hereditary. But we use those faculties—which some people have to a greater degree than others—to make decisions. So our ability to choose our fate is not free, but depends on our biological inheritance.

Magazine Cover image

Explore the June 2016 Issue

Check out more from this issue and find your next story to read.

Galton launched a debate that raged throughout the 20th century over nature versus nurture. Are our actions the unfolding effect of our genetics? Or the outcome of what has been imprinted on us by the environment? Impressive evidence accumulated for the importance of each factor. Whether scientists supported one, the other, or a mix of both, they increasingly assumed that our deeds must be determined by something .

In recent decades, research on the inner workings of the brain has helped to resolve the nature-nurture debate—and has dealt a further blow to the idea of free will. Brain scanners have enabled us to peer inside a living person’s skull, revealing intricate networks of neurons and allowing scientists to reach broad agreement that these networks are shaped by both genes and environment. But there is also agreement in the scientific community that the firing of neurons determines not just some or most but all of our thoughts, hopes, memories, and dreams.

We know that changes to brain chemistry can alter behavior—otherwise neither alcohol nor antipsychotics would have their desired effects. The same holds true for brain structure: Cases of ordinary adults becoming murderers or pedophiles after developing a brain tumor demonstrate how dependent we are on the physical properties of our gray stuff.

Many scientists say that the American physiologist Benjamin Libet demonstrated in the 1980s that we have no free will. It was already known that electrical activity builds up in a person’s brain before she, for example, moves her hand; Libet showed that this buildup occurs before the person consciously makes a decision to move. The conscious experience of deciding to act, which we usually associate with free will, appears to be an add-on, a post hoc reconstruction of events that occurs after the brain has already set the act in motion.

The 20th-century nature-nurture debate prepared us to think of ourselves as shaped by influences beyond our control. But it left some room, at least in the popular imagination, for the possibility that we could overcome our circumstances or our genes to become the author of our own destiny. The challenge posed by neuroscience is more radical: It describes the brain as a physical system like any other, and suggests that we no more will it to operate in a particular way than we will our heart to beat. The contemporary scientific image of human behavior is one of neurons firing, causing other neurons to fire, causing our thoughts and deeds, in an unbroken chain that stretches back to our birth and beyond. In principle, we are therefore completely predictable. If we could understand any individual’s brain architecture and chemistry well enough, we could, in theory, predict that individual’s response to any given stimulus with 100 percent accuracy.

This research and its implications are not new. What is new, though, is the spread of free-will skepticism beyond the laboratories and into the mainstream. The number of court cases, for example, that use evidence from neuroscience has more than doubled in the past decade—mostly in the context of defendants arguing that their brain made them do it. And many people are absorbing this message in other contexts, too, at least judging by the number of books and articles purporting to explain “your brain on” everything from music to magic. Determinism, to one degree or another, is gaining popular currency. The skeptics are in ascendance.

This development raises uncomfortable—and increasingly nontheoretical—questions: If moral responsibility depends on faith in our own agency, then as belief in determinism spreads, will we become morally irresponsible? And if we increasingly see belief in free will as a delusion, what will happen to all those institutions that are based on it?

In 2002, two psychologists had a simple but brilliant idea: Instead of speculating about what might happen if people lost belief in their capacity to choose, they could run an experiment to find out. Kathleen Vohs, then at the University of Utah, and Jonathan Schooler, of the University of Pittsburgh, asked one group of participants to read a passage arguing that free will was an illusion, and another group to read a passage that was neutral on the topic. Then they subjected the members of each group to a variety of temptations and observed their behavior. Would differences in abstract philosophical beliefs influence people’s decisions?

Yes, indeed. When asked to take a math test, with cheating made easy, the group primed to see free will as illusory proved more likely to take an illicit peek at the answers. When given an opportunity to steal—to take more money than they were due from an envelope of $1 coins—those whose belief in free will had been undermined pilfered more. On a range of measures, Vohs told me, she and Schooler found that “people who are induced to believe less in free will are more likely to behave immorally.”

It seems that when people stop believing they are free agents, they stop seeing themselves as blameworthy for their actions. Consequently, they act less responsibly and give in to their baser instincts. Vohs emphasized that this result is not limited to the contrived conditions of a lab experiment. “You see the same effects with people who naturally believe more or less in free will,” she said.

is free will an illusion essay

In another study, for instance, Vohs and colleagues measured the extent to which a group of day laborers believed in free will, then examined their performance on the job by looking at their supervisor’s ratings. Those who believed more strongly that they were in control of their own actions showed up on time for work more frequently and were rated by supervisors as more capable. In fact, belief in free will turned out to be a better predictor of job performance than established measures such as self-professed work ethic.

Another pioneer of research into the psychology of free will, Roy Baumeister of Florida State University, has extended these findings. For example, he and colleagues found that students with a weaker belief in free will were less likely to volunteer their time to help a classmate than were those whose belief in free will was stronger. Likewise, those primed to hold a deterministic view by reading statements like “Science has demonstrated that free will is an illusion” were less likely to give money to a homeless person or lend someone a cellphone.

Further studies by Baumeister and colleagues have linked a diminished belief in free will to stress, unhappiness, and a lesser commitment to relationships. They found that when subjects were induced to believe that “all human actions follow from prior events and ultimately can be understood in terms of the movement of molecules,” those subjects came away with a lower sense of life’s meaningfulness. Early this year, other researchers published a study showing that a weaker belief in free will correlates with poor academic performance.

The list goes on: Believing that free will is an illusion has been shown to make people less creative, more likely to conform, less willing to learn from their mistakes, and less grateful toward one another. In every regard, it seems, when we embrace determinism, we indulge our dark side.

Few scholars are comfortable suggesting that people ought to believe an outright lie. Advocating the perpetuation of untruths would breach their integrity and violate a principle that philosophers have long held dear: the Platonic hope that the true and the good go hand in hand. Saul Smilansky, a philosophy professor at the University of Haifa, in Israel, has wrestled with this dilemma throughout his career and come to a painful conclusion: “We cannot afford for people to internalize the truth” about free will.

Smilansky is convinced that free will does not exist in the traditional sense—and that it would be very bad if most people realized this. “Imagine,” he told me, “that I’m deliberating whether to do my duty, such as to parachute into enemy territory, or something more mundane like to risk my job by reporting on some wrongdoing. If everyone accepts that there is no free will, then I’ll know that people will say, ‘Whatever he did, he had no choice—we can’t blame him.’ So I know I’m not going to be condemned for taking the selfish option.” This, he believes, is very dangerous for society, and “the more people accept the determinist picture, the worse things will get.”

Determinism not only undermines blame, Smilansky argues; it also undermines praise. Imagine I do risk my life by jumping into enemy territory to perform a daring mission. Afterward, people will say that I had no choice, that my feats were merely, in Smilansky’s phrase, “an unfolding of the given,” and therefore hardly praiseworthy. And just as undermining blame would remove an obstacle to acting wickedly, so undermining praise would remove an incentive to do good. Our heroes would seem less inspiring, he argues, our achievements less noteworthy, and soon we would sink into decadence and despondency.

Smilansky advocates a view he calls illusionism—the belief that free will is indeed an illusion, but one that society must defend. The idea of determinism, and the facts supporting it, must be kept confined within the ivory tower. Only the initiated, behind those walls, should dare to, as he put it to me, “look the dark truth in the face.” Smilansky says he realizes that there is something drastic, even terrible, about this idea—but if the choice is between the true and the good, then for the sake of society, the true must go.

Smilansky’s arguments may sound odd at first, given his contention that the world is devoid of free will: If we are not really deciding anything, who cares what information is let loose? But new information, of course, is a sensory input like any other; it can change our behavior, even if we are not the conscious agents of that change. In the language of cause and effect, a belief in free will may not inspire us to make the best of ourselves, but it does stimulate us to do so.

Illusionism is a minority position among academic philosophers, most of whom still hope that the good and the true can be reconciled. But it represents an ancient strand of thought among intellectual elites. Nietzsche called free will “a theologians’ artifice” that permits us to “judge and punish.” And many thinkers have believed, as Smilansky does, that institutions of judgment and punishment are necessary if we are to avoid a fall into barbarism.

Smilansky is not advocating policies of Orwellian thought control . Luckily, he argues, we don’t need them. Belief in free will comes naturally to us. Scientists and commentators merely need to exercise some self-restraint, instead of gleefully disabusing people of the illusions that undergird all they hold dear. Most scientists “don’t realize what effect these ideas can have,” Smilansky told me. “Promoting determinism is complacent and dangerous.”

Yet not all scholars who argue publicly against free will are blind to the social and psychological consequences. Some simply don’t agree that these consequences might include the collapse of civilization. One of the most prominent is the neuroscientist and writer Sam Harris, who, in his 2012 book, Free Will , set out to bring down the fantasy of conscious choice. Like Smilansky, he believes that there is no such thing as free will. But Harris thinks we are better off without the whole notion of it.

“We need our beliefs to track what is true,” Harris told me. Illusions, no matter how well intentioned, will always hold us back. For example, we currently use the threat of imprisonment as a crude tool to persuade people not to do bad things. But if we instead accept that “human behavior arises from neurophysiology,” he argued, then we can better understand what is really causing people to do bad things despite this threat of punishment—and how to stop them. “We need,” Harris told me, “to know what are the levers we can pull as a society to encourage people to be the best version of themselves they can be.”

According to Harris, we should acknowledge that even the worst criminals—murderous psychopaths, for example—are in a sense unlucky. “They didn’t pick their genes. They didn’t pick their parents. They didn’t make their brains, yet their brains are the source of their intentions and actions.” In a deep sense, their crimes are not their fault. Recognizing this, we can dispassionately consider how to manage offenders in order to rehabilitate them, protect society, and reduce future offending. Harris thinks that, in time, “it might be possible to cure something like psychopathy,” but only if we accept that the brain, and not some airy-fairy free will, is the source of the deviancy.

Accepting this would also free us from hatred. Holding people responsible for their actions might sound like a keystone of civilized life, but we pay a high price for it: Blaming people makes us angry and vengeful, and that clouds our judgment.

“Compare the response to Hurricane Katrina,” Harris suggested, with “the response to the 9/11 act of terrorism.” For many Americans, the men who hijacked those planes are the embodiment of criminals who freely choose to do evil. But if we give up our notion of free will, then their behavior must be viewed like any other natural phenomenon—and this, Harris believes, would make us much more rational in our response.

Although the scale of the two catastrophes was similar, the reactions were wildly different. Nobody was striving to exact revenge on tropical storms or declare a War on Weather, so responses to Katrina could simply focus on rebuilding and preventing future disasters. The response to 9/11 , Harris argues, was clouded by outrage and the desire for vengeance, and has led to the unnecessary loss of countless more lives. Harris is not saying that we shouldn’t have reacted at all to 9/11, only that a coolheaded response would have looked very different and likely been much less wasteful. “Hatred is toxic,” he told me, “and can destabilize individual lives and whole societies. Losing belief in free will undercuts the rationale for ever hating anyone.”

Whereas the evidence from Kathleen Vohs and her colleagues suggests that social problems may arise from seeing our own actions as determined by forces beyond our control—weakening our morals, our motivation, and our sense of the meaningfulness of life—Harris thinks that social benefits will result from seeing other people’s behavior in the very same light. From that vantage point, the moral implications of determinism look very different, and quite a lot better.

What’s more, Harris argues, as ordinary people come to better understand how their brains work, many of the problems documented by Vohs and others will dissipate. Determinism, he writes in his book, does not mean “that conscious awareness and deliberative thinking serve no purpose.” Certain kinds of action require us to become conscious of a choice—to weigh arguments and appraise evidence. True, if we were put in exactly the same situation again, then 100 times out of 100 we would make the same decision, “just like rewinding a movie and playing it again.” But the act of deliberation—the wrestling with facts and emotions that we feel is essential to our nature—is nonetheless real.

The big problem, in Harris’s view, is that people often confuse determinism with fatalism. Determinism is the belief that our decisions are part of an unbreakable chain of cause and effect. Fatalism, on the other hand, is the belief that our decisions don’t really matter, because whatever is destined to happen will happen—like Oedipus’s marriage to his mother, despite his efforts to avoid that fate.

When people hear there is no free will, they wrongly become fatalistic; they think their efforts will make no difference. But this is a mistake. People are not moving toward an inevitable destiny; given a different stimulus (like a different idea about free will), they will behave differently and so have different lives. If people better understood these fine distinctions, Harris believes, the consequences of losing faith in free will would be much less negative than Vohs’s and Baumeister’s experiments suggest.

Can one go further still? Is there a way forward that preserves both the inspiring power of belief in free will and the compassionate understanding that comes with determinism?

Philosophers and theologians are used to talking about free will as if it is either on or off; as if our consciousness floats, like a ghost, entirely above the causal chain, or as if we roll through life like a rock down a hill. But there might be another way of looking at human agency.

Some scholars argue that we should think about freedom of choice in terms of our very real and sophisticated abilities to map out multiple potential responses to a particular situation. One of these is Bruce Waller, a philosophy professor at Youngstown State University. In his new book, Restorative Free Will , he writes that we should focus on our ability, in any given setting, to generate a wide range of options for ourselves, and to decide among them without external constraint.

For Waller, it simply doesn’t matter that these processes are underpinned by a causal chain of firing neurons. In his view, free will and determinism are not the opposites they are often taken to be; they simply describe our behavior at different levels.

Waller believes his account fits with a scientific understanding of how we evolved: Foraging animals—humans, but also mice, or bears, or crows—need to be able to generate options for themselves and make decisions in a complex and changing environment. Humans, with our massive brains, are much better at thinking up and weighing options than other animals are. Our range of options is much wider, and we are, in a meaningful way, freer as a result.

Waller’s definition of free will is in keeping with how a lot of ordinary people see it. One 2010 study found that people mostly thought of free will in terms of following their desires, free of coercion (such as someone holding a gun to your head). As long as we continue to believe in this kind of practical free will, that should be enough to preserve the sorts of ideals and ethical standards examined by Vohs and Baumeister.

Yet Waller’s account of free will still leads to a very different view of justice and responsibility than most people hold today. No one has caused himself: No one chose his genes or the environment into which he was born. Therefore no one bears ultimate responsibility for who he is and what he does. Waller told me he supported the sentiment of Barack Obama’s 2012 “You didn’t build that” speech, in which the president called attention to the external factors that help bring about success. He was also not surprised that it drew such a sharp reaction from those who want to believe that they were the sole architects of their achievements. But he argues that we must accept that life outcomes are determined by disparities in nature and nurture, “so we can take practical measures to remedy misfortune and help everyone to fulfill their potential.”

Understanding how will be the work of decades, as we slowly unravel the nature of our own minds. In many areas, that work will likely yield more compassion: offering more (and more precise) help to those who find themselves in a bad place. And when the threat of punishment is necessary as a deterrent, it will in many cases be balanced with efforts to strengthen, rather than undermine, the capacities for autonomy that are essential for anyone to lead a decent life. The kind of will that leads to success—seeing positive options for oneself, making good decisions and sticking to them—can be cultivated, and those at the bottom of society are most in need of that cultivation.

To some people, this may sound like a gratuitous attempt to have one’s cake and eat it too. And in a way it is. It is an attempt to retain the best parts of the free-will belief system while ditching the worst. President Obama—who has both defended “a faith in free will” and argued that we are not the sole architects of our fortune—has had to learn what a fine line this is to tread. Yet it might be what we need to rescue the American dream—and indeed, many of our ideas about civilization, the world over—in the scientific age.

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 13 May 2009

Is free will an illusion?

  • Martin Heisenberg 1  

Nature volume  459 ,  pages 164–165 ( 2009 ) Cite this article

14k Accesses

58 Citations

66 Altmetric

Metrics details

Scientists and philosophers are using new discoveries in neuroscience to question the idea of free will. They are misguided, says Martin Heisenberg. Examining animal behaviour shows how our actions can be free.

You have full access to this article via your institution.

Our influence on the future is something we take for granted as much as breathing. We accept that what will be is not yet determined, and that we can steer the course of events in one direction or another. This idea of freedom, and the sense of responsibility it bestows, seems essential to day-to-day existence.

is free will an illusion essay

Yet it is under attack as never before. Some scientists and philosophers argue that recent findings in neuroscience — such as data published last year suggesting that our brain makes decisions up to seven seconds before we become aware of them — along with the philosophical principle that any action must be dependent on preceding causes, imply that our behaviour is never self-generated and that freedom is an illusion 1 , 2 , 3 .

This debate has focused on humans and 'conscious free will'. Yet when it comes to understanding how we initiate behaviour, we can learn a lot by looking at animals. Although we do not credit animals with anything like the consciousness in humans, researchers have found that animal behaviour is not as involuntary as it may appear. The idea that animals act only in response to external stimuli has long been abandoned, and it is well established that they initiate behaviour on the basis of their internal states, as we do.

Before going into behaviour, I would like to take a step back and look at the nature of freedom and determinism at a more fundamental level. Almost 100 years ago, quantum physics eliminated a major obstacle to our understanding of this issue when it disposed of the idea of a Universe determined in every detail from the outset. It uncovered an inherent unpredictability in nature, in that we can never know precisely at a given moment all properties of a particle — such as both its position and its momentum.

is free will an illusion essay

How is this reflected at the level of everyday experience? At the scale of planets, quantum effects give way to the deterministic laws of classical mechanics. At an intermediate scale, however, they are occasionally amplified to become observable, for example when we measure radioactive decay. In general, life is an interplay between the deterministic and the random. There is plenty of evidence of chance at work in the brain: take the random opening and closing of ion channels in the neuronal membrane, or the miniature potentials of randomly discharging synaptic vesicles. Behaviour that is triggered by random events in the brain can be said to be truly 'active' — in other words, it has the quality of a beginning.

Evidence of randomly generated action — action that is distinct from reaction because it does not depend upon external stimuli — can be found in unicellular organisms. Take the way the bacterium Escherichia coli moves. It has a flagellum that can rotate around its longitudinal axis in either direction: one way drives the bacterium forward, the other causes it to tumble at random so that it ends up facing in a new direction ready for the next phase of forward motion. This 'random walk' can be modulated by sensory receptors, enabling the bacterium to find food and the right temperature.

What this tells us is that behavioural output can be independent of sensory input. This is in line with the fact that in the early development of individual organisms the motor system slightly precedes the sensory system. The same may have been true in evolution, as merely being dispersed in space should have been advantageous and should have favoured mobility.

What of more complex behaviour? With the emergence of multicellularity, individual cells lost their behavioural autonomy and organisms had to reinvent locomotion. Behaviours in complex organisms typically come in modules: the grasp reflex of the newborn, the syllables of birdsong, the rhythmic motion of the legs during walking. Some modules, such as the heartbeat, last from embryonic development until death; others, such as the snapping of a crocodile's jaw, last just fractions of a second. Some can take place in parallel, like walking and singing; others are mutually exclusive, such as sleeping and playing the piano. Some necessarily follow one another, like flight and landing. From beginning to end, the lives of animals and humans are an ongoing interweaving of these behavioural modules.

There is plenty of evidence that an animal's behaviour cannot be reduced to responses.

As with a bacterium's locomotion, the activation of behavioural modules is based on the interplay between chance and lawfulness in the brain. Insufficiently equipped, insufficiently informed and short of time, animals have to find a module that is adaptive. Their brains, in a kind of random walk, continuously pre-activate, discard and reconfigure their options, and evaluate their possible short-term and long-term consequences.

The physiology of how this happens has been little investigated. But there is plenty of evidence that an animal's behaviour cannot be reduced to responses. For example, my lab has demonstrated that fruit flies, in situations they have never encountered, can modify their expectations about the consequences of their actions. They can solve problems that no individual fly in the evolutionary history of the species has solved before. Our experiments show that they actively initiate behaviour 4 . Like humans who can paint with their toes, we have found that flies can be made to use several different motor outputs to escape a life-threatening danger or to visually stabilize their orientation in space 5 .

Does this tell us anything about freedom in human behaviour? Before I answer that, let's establish what I mean by freedom. One acknowledged definition comes from Immanuel Kant, who resolved that a person acts freely if he does of his own accord what must be done. Thus, my actions are not free if they are determined by something or someone else. As stated above, self-initiated action is not in conflict with physics and can be demonstrated in animals. So, humans can be considered free in their behaviour, in as much as their behaviour is self-initiated and adaptive.

Some define freedom as the ability to consciously decide how to act. I maintain that we need not be conscious of our decision-making to be free. What matters is that our actions are self-generated. Conscious awareness may help improve our behaviour, but it does not necessarily do so and is not essential. Why should an action become free from one moment to the next simply because we reflect upon it?

Kant's famous 'Third Antinomy' in his Critique of Pure Reason (1781) sees us on the one hand determined by natural law and on the other free because of our capacity to obey moral law. He would have been delighted to see this dilemma solved by quantum physics and behavioural biology.

Soon, C. S. Nature Neurosci. 11 , 543–545 (2008).

Article   CAS   Google Scholar  

Libet, B. Behav. Brain Sci. 8 , 529–566 (1985).

Article   Google Scholar  

Wegner, D. M. The Illusion of Conscious Will (MIT Press, 2002).

Book   Google Scholar  

Heisenberg, M. Naturwissenschaften 70 , 70–78 (1983).

Article   ADS   CAS   Google Scholar  

Heisenberg, M. & Wolf, R. Vision in Drosophila in Studies of Brain Function Vol. XII , (ed. Braitenberg, V.) (Springer, 1984).

Google Scholar  

Download references

Author information

Authors and affiliations.

Martin Heisenberg is professor emeritus in the department of biology at the University of Würzburg, Germany. [email protected],

Martin Heisenberg

You can also search for this author in PubMed   Google Scholar

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Heisenberg, M. Is free will an illusion?. Nature 459 , 164–165 (2009). https://doi.org/10.1038/459164a

Download citation

Published : 13 May 2009

Issue Date : 14 May 2009

DOI : https://doi.org/10.1038/459164a

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Evolutionary perspective on human cognitive architecture in cognitive load theory: a dynamic, emerging principle approach.

  • Slava Kalyuga

Educational Psychology Review (2023)

Austrian Economics and Compatibilist Freedom

  • Igor Wysocki
  • Łukasz Dominiak

Journal for General Philosophy of Science (2023)

Consciousness, decision making, and volition: freedom beyond chance and necessity

  • Hans Liljenström

Theory in Biosciences (2022)

Can Reasons and Values Influence Action: How Might Intentional Agency Work Physiologically?

  • Raymond Noble
  • Denis Noble

Journal for General Philosophy of Science (2021)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

is free will an illusion essay

January 16, 2023

Free Will Is Only an Illusion if You Are, Too

New research findings, combined with philosophy, suggest free will is real but may not operate in the ways people expect

By Alessandra Buccella & Tomáš Dominik

Surreal optical illusion, man change his point of view

francescoch/Getty Images

Imagine you are shopping online for a new pair of headphones. There is an array of colors, brands and features to look at. You feel that you can pick any model that you like and are in complete control of your decision. When you finally click the “add to shopping cart” button, you believe that you are doing so out of your own free will.

But what if we told you that while you thought that you were still browsing, your brain activity had already highlighted the headphones you would pick? That idea may not be so far-fetched. Though neuroscientists likely could not predict your choice with 100 percent accuracy, research has demonstrated that some information about your upcoming action is present in brain activity several seconds before you even become conscious of your decision.

As early as the 1960s, studies found that when people perform a simple, spontaneous movement, their brain exhibits a buildup in neural activity —what neuroscientists call a “readiness potential”—before they move. In the 1980s, neuroscientist Benjamin Libet reported this readiness potential even preceded a person’s reported intention to move, not just their movement. In 2008 a group of researchers found that some information about an upcoming decision is present in the brain up to 10 seconds in advance , long before people reported making the decision of when or how to act.

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.

These studies have sparked questions and debates . To many observers, these findings debunked the intuitive concept of free will. After all, if neuroscientists can infer the timing or choice of your movements long before you are consciously aware of your decision, perhaps people are merely puppets, pushed around by neural processes unfolding below the threshold of consciousness.

But as researchers who study volition from both a neuroscientific and philosophical perspective, we believe that there’s still much more to this story. We work with a collaboration of philosophers and scientists to provide more nuanced interpretations—including a better understanding of the readiness potential —and a more fruitful theoretical framework in which to place them. The conclusions suggest “free will” remains a useful concept, although people may need to reexamine how they define it.

Let’s start from a commonsense observation: much of what people do each day is arbitrary. We put one foot in front of the other when we start walking. Most of the time, we do not actively deliberate about which leg to put forward first. It doesn’t matter. The same is true for many other actions and choices. They are largely meaningless and irreflective.

Most empirical studies of free will—including Libet’s—have focused on these kinds of arbitrary actions. In such actions, researchers can indeed “read out” our brain activity and trace information about our movements and choices before we even realize we are about to make them. But if these actions don’t matter to us, is it all that notable that they are initiated unconsciously? More significant decisions—such as whether to take a job, get married or move to a different country—are infinitely more interesting and complex and are quite consciously made.

If we start working with a more philosophically grounded understanding of free will, we realize that only a small subset of our everyday actions is important enough to worry about. We want to feel in control of those decisions, the ones whose outcomes make a difference in our life and whose responsibility we feel on our shoulders. It is in this context—decisions that matter —that the question of free will most naturally applies.

In 2019 neuroscientists Uri Maoz, Liad Mudrik and their colleagues investigated that idea. They presented participants with a choice of two nonprofit organizations to which they could donate $1,000. People could indicate their preferred organization by pressing the left or right button. In some cases, participants knew that their choice mattered because the button would determine which organization would receive the full $1,000. In other cases, people knowingly made meaningless choices because they were told that both organizations would receive $500 regardless of their selection. The results were somewhat surprising. Meaningless choices were preceded by a readiness potential, just as in previous experiments. Meaningful choices were not , however. When we care about a decision and its outcome, our brain appears to behave differently than when a decision is arbitrary.

Even more interesting is the fact that ordinary people’s intuitions about free will and decision-making do not seem consistent with these findings. Some of our colleagues, including Maoz and neuroscientist Jake Gavenas, recently published the results of a large survey, with more than 600 respondents, in which they asked people to rate how “free” various choices made by others seemed. Their ratings suggested that people do not recognize that the brain may handle meaningful choices in a different way from more arbitrary or meaningless ones. People tend, in other words, to imagine all their choices—from which sock to put on first to where to spend a vacation—as equally “free,” even though neuroscience suggests otherwise.

What this tells us is that free will may exist, but it may not operate in the way we intuitively imagine. In the same vein, there is a second intuition that must be addressed to understand studies of volition. When experiments have found that brain activity, such as the readiness potential, precedes the conscious intention to act, some people have jumped to the conclusion that they are “not in charge.” They do not have free will, they reason, because they are somehow subject to their brain activity.

But that assumption misses a broader lesson from neuroscience. “We” are our brain. The combined research makes clear that human beings do have the power to make conscious choices. But that agency and accompanying sense of personal responsibility are not supernatural. They happen in the brain, regardless of whether scientists observe them as clearly as they do a readiness potential.

So there is no “ghost” inside the cerebral machine. But as researchers, we argue that this machinery is so complex, inscrutable and mysterious that popular concepts of “free will” or the “self” remain incredibly useful. They help us think through and imagine—albeit imperfectly—the workings of the mind and brain. As such, they can guide and inspire our investigations in profound ways—provided we continue to question and test these assumptions along the way.

Are you a scientist who specializes in neuroscience, cognitive science or psychology? And have you read a recent peer-reviewed paper that you would like to write about for Mind Matters? Please send suggestions to  Scientific American ’s Mind Matters editor Daisy Yuhas   at  [email protected] .

SEP thinker apres Rodin

“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. (And what a fuss it has been: philosophers have debated this question for over two millenia, and just about every major philosopher has had something to say about it.) Most philosophers suppose that the concept of free will is very closely connected to the concept of moral responsibility. Acting with free will, on such views, is just to satisfy the metaphysical requirement on being responsible for one's action. (Clearly, there will also be epistemic conditions on responsibility as well, such as being aware—or failing that, being culpably unaware—of relevant alternatives to one's action and of the alternatives' moral significance.) But the significance of free will is not exhausted by its connection to moral responsibility. Free will also appears to be a condition on desert for one's accomplishments (why sustained effort and creative work are praiseworthy); on the autonomy and dignity of persons; and on the value we accord to love and friendship. (See Kane 1996, 81ff. and Clarke 2003, Ch.1.)

Philosophers who distinguish freedom of action and freedom of will do so because our success in carrying out our ends depends in part on factors wholly beyond our control. Furthermore, there are always external constraints on the range of options we can meaningfully try to undertake. As the presence or absence of these conditions and constraints are not (usually) our responsibility, it is plausible that the central loci of our responsibility are our choices, or “willings.”

I have implied that free willings are but a subset of willings, at least as a conceptual matter. But not every philosopher accepts this. René Descartes, for example, identifies the faculty of will with freedom of choice, “the ability to do or not do something” (Meditation IV), and even goes so far as to declare that “the will is by its nature so free that it can never be constrained” (Passions of the Soul, I, art. 41). In taking this strong polar position on the nature of will, Descartes is reflecting a tradition running through certain late Scholastics (most prominently, Suarez) back to John Duns Scotus.

The majority view, however, is that we can readily conceive willings that are not free. Indeed, much of the debate about free will centers around whether we human beings have it, yet virtually no one doubts that we will to do this and that. The main perceived threats to our freedom of will are various alleged determinisms: physical/causal; psychological; biological; theological. For each variety of determinism, there are philosophers who (i) deny its reality, either because of the existence of free will or on independent grounds; (ii) accept its reality but argue for its compatibility with free will; or (iii) accept its reality and deny its compatibility with free will. (See the entries on compatibilism ; causal determinism ; fatalism ; arguments for incompatibilism ; and divine foreknowedge and free will .) There are also a few who say the truth of any variety of determinism is irrelevant because free will is simply impossible.

If there is such a thing as free will, it has many dimensions. In what follows, I will sketch the freedom-conferring characteristics that have attracted most of the attention. The reader is warned, however, that while many philosophers emphasize a single such characteristic, perhaps in response to the views of their immediate audience, it is probable that most would recognize the significance of many of the other features discussed here.

1.1 Free Will as Choosing on the Basis of One's Desires

  • 1.2 Free Will as Deliberative Choosing on the Basis of Desires and Values

1.3 Self-mastery, Rightly-Ordered Appetite

2. ownership, 3.1 free will as guidance control, 3.2 free will as ultimate origination (ability to do otherwise), 3.3 do we have free will, 4. theological wrinkles, bibliography, other internet resources, related entries, 1. rational deliberation.

On a minimalist account, free will is the ability to select a course of action as a means of fulfilling some desire. David Hume, for example, defines liberty as “a power of acting or of not acting, according to the determination of the will.” (1748, sect.viii, part 1). And we find in Jonathan Edwards (1957) a similar account of free willings as those which proceed from one's own desires.

One reason to deem this insufficient is that it is consistent with the goal-directed behavior of some animals whom we do not suppose to be morally responsible agents. Such animals lack not only an awareness of the moral implications of their actions but also any capacity to reflect on their alternatives and their long-term consequences. Indeed, it is plausible that they have little by way of a self-conception as an agent with a past and with projects and purposes for the future. (See Baker 2000 on the ‘first-person perspective.’)

1.2 Free Will as deliberative choosing on the basis of desires and values

A natural suggestion, then, is to modify the minimalist thesis by taking account of (what may be) distinctively human capacities and self-conception. And indeed, philosophers since Plato have commonly distinguished the ‘animal’ and ‘rational’ parts of our nature, with the latter implying a great deal more psychological complexity. Our rational nature includes our ability to judge some ends as ‘good’ or worth pursuing and value them even though satisfying them may result in considerable unpleasantness for ourselves. (Note that such judgments need not be based in moral value.) We might say that we act with free will when we act upon our considered judgments/valuings about what is good for us, whether or not our doing so conflicts with an ‘animal’ desire. (See Watson 2003a for a subtle development of this sort of view.) But this would seem unduly restrictive, since we clearly hold many people responsible for actions proceeding from ‘animal’ desires that conflict with their own assessment of what would be best in the circumstances. More plausible is the suggestion that one acts with free will when one's deliberation is sensitive to one's own judgments concerning what is best in the circumstances, whether or not one acts upon such a judgment.

Here we are clearly in the neighborhood of the ‘rational appetite’ accounts of will one finds in the medieval Aristotelians. The most elaborate medieval treatment is Thomas Aquinas's. [ 1 ] His account involves identifying several distinct varieties of willings. Here I note only a few of his basic claims. Aquinas thinks our nature determines us to will certain general ends ordered to the most general goal of goodness. These we will of necessity, not freely. Freedom enters the picture when we consider various means to these ends, none of which appear to us either as unqualifiedly good or as uniquely satisfying the end we wish to fulfill. There is, then, free choice of means to our ends, along with a more basic freedom not to consider something, thereby perhaps avoiding willing it unavoidably once we recognized its value. Free choice is an activity that involves both our intellectual and volitional capacities, as it consists in both judgment and active commitment. A thorny question for this view is whether will or intellect is the ultimate determinant of free choices. How we understand Aquinas on this point will go a long ways towards determining whether or not he is a sort of compatibilist about freedom and determinism. (See below. Good expositions of Aquinas' account are Donagan 1985, Stump 1997, and MacDonald 1998.)

There are two general worries about theories of free will that principally rely on the capacity to deliberate about possible actions in the light of one's conception of the good. First, there are agents who deliberately choose to act as they do but who are motivated to do so by a compulsive, controlling sort of desire. (And there seems to be no principled bar to a compulsive desire's informing a considered judgment of the agent about what the good is for him.) Such agents are not willing freely. (Wallace 2003 offers an account of the way addiction impairs the will.) Secondly, we can imagine a person's psychology being externally manipulated by another agent (via neurophysiological implant, say), such that the agent is caused to deliberate and come to desire strongly a particular action which he previously was not disposed to choose. The deliberative process could be perfectly normal, reflective, and rational, but seemingly not freely made. The agent's freedom seems undermined or at least greatly diminished by such psychological tampering.

Some theorists are much impressed by cases of inner, psychological compulsion and define freedom of will in contrast to this phenomenon. For such thinkers, true freedom of the will involves liberation from the tyranny of base desires and acquisition of desires for the Good. Plato, for example, posits rational, spirited, and appetitive aspects to the soul and holds that willings issue from the higher, rational part alone. In other cases, one is dominated by the irrational desires of the two lower parts. [ 2 ] This is particularly characteristic of those working in a theological context—for example, the New Testament writer St. Paul, speaking of Christian freedom (Romans vi-viii; Galatians v), and those influenced by him on this point, such as Augustine. (The latter, in both early and later writings, allows for a freedom of will that is not ordered to the good, but maintains that it is of less value than the rightly-ordered freedom. See, for example, the discussion in Books II-III of On Free Choice .) More recently, Susan Wolf (1990) defends an asymmetry thesis concerning freedom and responsibility. On her view, an agent acts freely only if he had the ability to choose the True and the Good. For an agent who does so choose, the requisite ability is automatically implied. But those who reject the Good choose freely only if they could have acted differently. This is a further substantive condition on freedom, making freedom of will a more demanding condition in cases of bad choices.

In considering such rightly-ordered-appetites views of freedom, I again focus on abstract features common to all. It explicitly handles the inner-compulsion worry facing the simple deliberation-based accounts. The other, external manipulation problem could perhaps be handled through the addition of an historical requirement: agents will freely only if their willings are not in part explicable by episodes of external manipulation which bypass their critical and deliberative faculties (Mele 1995). But another problem suggests itself: an agent who was a ‘natural saint,’ always and effortlessly choosing the good with no contrary inclination, would not have freedom of will among his virtues. Doubtless we would greatly admire such a person, but would it be an admiration suffused with moral praise of the person or would it, rather, be restricted to the goodness of the person's qualities? (Cf. Kant, 1788.) The appropriate response to such a person, it seems, is on an analogy with aesthetic appreciation of natural beauty, in contrast to the admiration of the person who chooses the good in the face of real temptation to act selfishly. Since this view of freedom of will as orientation to the good sometimes results from theological reflections, it is worth noting that other theologian-philosophers emphasize the importance for human beings of being able to reject divine love: love of God that is not freely given—given in the face of a significant possibility of one's having not done so—would be a sham, all the more so since, were it inevitable, it would find its ultimate and complete explanation in God Himself.

Harry Frankfurt (1982) presents an insightful and original way of thinking about free will. He suggests that a central difference between human and merely animal activity is our capacity to reflect on our desires and beliefs and form desires and judgments concerning them. I may want to eat a candy bar (first-order desire), but I also may want not to want this (second-order desire) because of the connection between habitual candy eating and poor health. This difference, he argued, provides the key to understanding both free action and free will. (These are quite different, in Frankfurt's view, with free will being the more demanding notion. Moreover, moral responsibility for an action requires only that the agent acted freely, not that the action proceeded from a free will.)

On Frankfurt's analysis, I act freely when the desire on which I act is one that I desire to be effective. This second-order desire is one with which I identify : it reflects my true self. (Compare the addict: typically, the addict acts out of a desire which he does not want to act upon. His will is divided, and his actions proceed from desires with which he does not reflectively identify. Hence, he is not acting freely.) My will is free when I am able to make any of my first-order desires the one upon which I act. As it happens, I will to eat the candy bar, but I could have willed to refrain from doing so.

With Frankfurt's account of free will, much hangs on what being able to will otherwise comes to, and on this Frankfurt is officially neutral. (See the related discussion below on ability to do otherwise.) But as he connects moral responsibility only to his weaker notion of free action, it is fitting to consider its adequacy here. The central objection that commentators have raised is this: what's so special about higher-order willings or desires? (See in particular Watson 2003a.) Why suppose that they inevitably reflect my true self, as against first-order desires? Frankfurt is explicit that higher-order desires need not be rooted in a person's moral or even settled outlook (1982, 89, n.6). So it seems that, in some cases, a first-order desire may be much more reflective of my true self (more “internal to me,” in Frankfurt's terminology) than a weak, faint desire to be the sort of person who wills differently.

In later writings, Frankfurt responds to this worry first by appealing to “decisions made without reservations” (“Identification and Externality” and “Identification and Wholeheartedness” in Frankfurt, 1988) and then by appealing to higher-order desires with which one is “satisfied,” such that one has no inclination to make changes to them (1992). But the absence of an inclination to change the desire does not obviously amount to the condition of freedom-conferring identification. It seems that such a negative state of satisfaction can be one that I just find myself with, one that I neither approve nor disapprove (Pettit, 2001, 56).

Furthermore, we can again imagine external manipulation consistent with Frankfurt's account of freedom but inconsistent with freedom itself. Armed with the wireless neurophysiology-tampering technology of the late 21st century, one might discreetly induce a second-order desire in me to be moved by a first-order desire—a higher-order desire with which I am satisfied—and then let me deliberate as normal. Clearly, this desire should be deemed “external” to me, and the action that flows from it unfree.

3. Causation and Control

Our survey of several themes in philosophical accounts of free will suggests that a—perhaps the —root issue is that of control . Clearly, our capacity for deliberation and the potential sophistication of some of our our practical reflections are important conditions on freedom of will. But any proposed analysis of free will must also ensure that the process it describes is one that was up to, or controlled by, the agent.

Fantastic scenarios of external manipulation and less fantastic cases of hypnosis are not the only, or even primary, ones to give philosophers pause. It is consistent with my deliberating and choosing ‘in the normal way’ that my developing psychology and choices over time are part of an ineluctable system of causes necessitating effects. It might be, that is, that underlying the phenomena of purpose and will in human persons is an all-encompassing, mechanistic world-system of ‘blind’ cause and effect. Many accounts of free will are constructed against the backdrop possibility (whether accepted as actual or not) that each stage of the world is determined by what preceded it by impersonal natural law. As always, there are optimists and pessimists.

John Martin Fischer (1994) distinguishes two sorts of control over one's actions: guidance and regulative. A person exerts guidance control over his own actions insofar as they proceed from a ‘weakly’ reasons-responsive (deliberative) mechanism. This obtains just in case there is some possible scenario where the agent is presented with a sufficient reason to do otherwise and the mechanism that led to the actual choice is operative and it issues in a different choice, one appropriate to the imagined reason. In Fischer and Ravizza (1998), the account is elaborated and refined. They require, more strongly, that the mechanism be the person's own mechanism (ruling out external manipulation) and that it be ‘moderately’ responsive to reasons: one that is “regularly receptive to reasons, some of which are moral reasons, and at least weakly reactive to reason” (82, emphasis added). Receptivity is evinced through an understandable pattern of reasons recognition—beliefs of the agent about what would constitute a sufficient reason for undertaking various actions. (See 69-73 for details.)

None of this, importantly, requires ‘regulative’ control: a control involving the ability of the agent to choose and act differently in the actual circumstances. Regulative control requires alternative possibilities open to the agent, whereas guidance control is determined by characteristics of the actual sequence issuing in one's choice. Fischer allows that there is a notion of freedom that requires regulative control but does not believe that this kind of freedom is required for moral responsibility. (In this, he is persuaded by a form of argument originated by Harry Frankfurt. See Frankfurt 1969 and Fischer 1994, Ch.7 for an important development of the argument. The argument has been debated extensively in recent years, and Widerker and McKenna 2002 offers a representative sampling.)

Many do not follow Fischer here, however, and maintain the traditional view that the sort of freedom required for moral responsibility does indeed require that the agent could have acted differently. As Aristotle put it, “…when the origin of the actions is in him, it is also up to him to do them or not to do them” (1985, Book III). [ 3 ]

A flood of ink has been spilled, especially in the modern era, on how to understand the concept of being able to do otherwise. On one side are those who give it a deflationary reading, on which it is consistent with my being able to do otherwise that the past (including my character and present beliefs and desires) and the basic laws of nature logically entail that I do what I actually do. These are the ‘compatibilists,’ holding that freedom and causal determinism are compatible. (For discussion, see O'Connor, 2000, Ch.1; compatibilism ; and incompatibilism: arguments for .) Conditional analyses of ability to do otherwise have been popular among compatibilists. The general idea here is that to say that I am able to do otherwise is to say that I would do otherwise if it were the case that … , where the ellipsis is filled by some elaboration of “I had an appropriately strong desire to do so, or I had different beliefs about the best available means to satisfy my goal, or … .” In short: something about my prevailing character or present psychological states would have differed, and so would have brought about a different outcome in my deliberation.

Incompatibilists think that something stronger is required: for me to act with free will requires that there are a plurality of futures open to me consistent with the past (and laws of nature) being just as they were . I could have chosen differently even without some further, non-actual consideration's occurring to me and ‘tipping the scales of the balance’ in another direction. Indeed, from their point of view, the whole scale-of-weights analogy is wrongheaded: free agents are not mechanisms that respond invariably to specified ‘motive forces.’ They are capable of acting upon any of a plurality of motives making attractive more than one course of action. Ultimately, the agent must determine himself this way or that.

We may distinguish two broad families of ‘incompatibilist’ or ‘indeterminist’ self-determination accounts. The more radical group holds that the agent who determines his own will is not causally influenced by external causal factors, including his own character. Descartes, in the midst of exploring the scope and influence of ‘the passions,’ declares that “the will is by its nature so free that it can never be constrained” (1984, v.I, 343). And as we've seen, he believed that such freedom is present on every occasion when we make a conscious choice—even, he writes, “when a very evident reason moves us in one direction….” (1984, v.III, 245). More recently, John Paul Sartre notoriously held that human beings have ‘absolute freedom’: “No limits to my freedom can be found except freedom itself, or, if you prefer, we are not free to cease being free.” (567) His views on freedom flowed from his radical conception of human beings as lacking any kind of positive nature. Instead, we are ‘non-beings’ whose being, moment to moment, is simply to choose:

For human reality, to be is to choose oneself; nothing comes to it either from the outside or from within which it can receive or accept….it is entirely abandoned to the intolerable necessity of making itself be, down to the slightest details. Thus freedom…is the being of man, i.e., his nothingness of being. (568-9)

Scotus and, more recently, C.A. Campbell, appear to agree with Descartes and Sartre on the lack of direct causal influence on the activity of free choice while allowing that the scope of possibilities for what I might thus will may be more or less constricted. So while Scotus holds that “nothing other than the will is the total cause” of its activity, he grants (with Aquinas and other medieval Aristotelians) that we are not capable of willing something in which we see no good, nor of positively repudiating something which appears to us as unqualifiedly good. Contrary to Sartre, we come with a ‘nature’ that circumscribes what we might conceivably choose, and our past choices and environmental influences also shape the possibilities for us at any particular time. But if we are presented with what we recognize as an unqualified good, we still can choose to refrain from willing it. And while Campbell holds that character cannot explain a free choice, he supposes that “[t]here is one experiential situation, and one only , … in which there is any possibility of the act of will not being in accordance with character; viz. the situation in which the course which formed character prescribes is a course in conflict with the agent's moral ideal: in other words, the situation of moral temptation” (1967, 46). (Van Inwagen 1994 and 1995 is another proponent of the idea that free will is exercised in but a small subset of our choices, although his position is less extreme on this point than Campbell's. Fischer and Ravizza 1992, O'Connor 2000, Ch.5, and Clarke 2003, Ch.7 all criticize van Inwagen's argument for this position.)

A more moderate grouping within the self-determination approach to free will allows that beliefs, desires, and external factors all can causally influence the act of free choice itself. But theorists within this camp differ sharply on the metaphysical nature of those choices and of the causal role of reasons. We may distinguish three varieties. I will discuss them only briefly, as they are explored at length in incompatibilist (nondeterministic) theories of free will .

First is a noncausal (or ownership) account (Ginet 1990, 2002 and McCann 1998). According to this view, I control my volition or choice simply in virtue of its being mine—its occurring in me. I do not exert a special kind of causality in bringing it about; instead, it is an intrinsically active event, intrinsically something I do . While there may be causal influences upon my choice, there need not be, and any such causal influence is wholly irrelevant to understanding why it occurs. Reasons provide an autonomous, non-causal form of explanation. Provided my choice is not wholly determined by prior factors, it is free and under my control simply in virtue of being mine.

Proponents of the event-causal account (e.g. Nozick 1995 and Ekstrom 2001) would say that uncaused events of any kind would be random and uncontrolled by anyone, and so could hardly count as choices that an agent made . They hold that reasons influence choices precisely by causing them. Choices are free insofar as they are not deterministically caused, and so might not have occurred in just the circumstances in which they did occur. (See nondeterministic theories of free will and probabilistic causation .) A special case of the event-causal account of self-determination is Kane (1996). Kane believes that the free choices of greatest significance to an agent's autonomy are ones that are preceded by efforts of will within the process of deliberation. These are cases where one's will is conflicted, as when one's duty or long-term self-interest compete with a strong desire for a short-term good. As one struggles to sort out and prioritize one's own values, the possible outcomes are not merely undetermined, but also indeterminate : at each stage of the struggle, the possible outcomes have no specific objective probability of occurring. This indeterminacy, Kane believes, is essential to freedom of will.

Finally, there are those who believe freedom of will consists in a distinctively personal form of causality, commonly referred to as “agent causation.” The agent himself causes his choice or action, and this is not to be reductively analyzed as an event within the agent causing the choice. (Compare our ready restatement of “the rock broke the window” into the more precise “the rock's being in momentum M at the point of contact with the window caused the window's subsequent shattering.”) This view is given clear articulation by Thomas Reid:

I grant, then, that an effect uncaused is a contradiction, and that an event uncaused is an absurdity. The question that remains is whether a volition, undetermined by motives, is an event uncaused. This I deny. The cause of the volition is the man that willed it. (Letter to James Gregory, in 1967, 88)

Roderick Chisholm advocated this view of free will in numerous writings (e.g., 1982 and 1976). And recently it has been developed in different forms by Randolph Clarke (1993, 1996, 2003) and O'Connor (2000). Nowadays, many philosophers view this account as of doubtful coherence (e.g., Dennett 1984). For some, this very idea of causation by a substance just as such is perplexing (Ginet 1997 and Clarke 2003, Ch.10). Others see it as difficult to reconcile with the causal role of reasons in explaining choices. (Clarke and O'Connor devote considerable effort to addressing this concern.) And yet others hold that, coherent or not, it is inconsistent with seeing human beings as part of the natural world of cause and effect (Pereboom 2001).

A recent trend is to suppose that agent causation accounts capture, as well as possible, our prereflective idea of responsible, free action. But the failure of philosophers to work the account out in a fully satisfactory and intelligible form reveals that the very idea of free will (and so of responsibility) is incoherent (Strawson 1986) or at least inconsistent with a world very much like our own (Pereboom 2001). Smilansky (2000) takes a more complicated position, on which there are two ‘levels’ on which we may assess freedom, ‘compatibilist’ and ‘ultimate’. On the ultimate level of evaluation, free will is indeed incoherent. (Strawson, Pereboom, and Smilansky all provide concise defenses of their positions in Kane 2002.)

The will has also recently become a target of empirical study in neuroscience and cognitive psychology. Benjamin Libet (2002) conducted experiments designed to determine the timing of conscious willings or decisions to act in relation to brain activity associated with the physical initiation of behavior. Interpretation of the results is highly controversial. Libet himself concludes that the studies provide strong evidence that actions are already underway shortly before the agent wills to do it. As a result, we do not consciously initiate our actions, though he suggests that we might nonetheless retain the ability to veto actions that are initiated by unconscious psychological structures. Wegner (2002) masses a much range of studies (including those of Libet) to argue that the notion that human actions are ever initiated by their own conscious willings is simply a deeply-entrenched illusion and proceeds to offer an hypothesis concerning the reason this illusion is generated within our cognitive systems. O'Connor (forthcoming) argues that the data adduced by Libet and Wegner wholly fail to support their revisionary conclusions.

A large portion of Western philosophical writing on free will was and is written within an overarching theological framework, according to which God is the ultimate source and sustainer of all else. Some of these thinkers draw the conclusion that God must be a sufficient, wholly determining cause for everything that happens; all suppose that every creaturely act necessarily depends on the explanatorily prior, cooperative activity of God. It is also presumed that human beings are free and responsible (on pain of attributing evil in the world to God alone, and so impugning His perfect goodness). Hence, those who believe that God is omni-determining typically are compatibilists with respect to freedom and (in this case) theological determinism. Edwards (1957) is a good example. But those who suppose that God's sustaining activity (and special activity of conferring grace) is only a necessary condition on the outcome of human free choices need to tell a more subtle story, on which omnipotent God's cooperative activity can be (explanatorily) prior to a human choice and yet the outcome of that choice be settled only by the choice itself. (For important medieval discussions—the period of the apex of treatments of philosophical/theological matters—see the relevant portions of Aquinas 1945 and Scotus 1994. For an example of a more recent discussion, see Quinn 1983.)

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

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

One suggested solution to this puzzle begins by reconsidering the relationship of two strands in (much) thinking about freedom of will: being able to do otherwise and being the ultimate source of one's will. Contemporary discussions of free will often emphasize the importance of being able to do otherwise. Yet it is plausible (Kane 1996) that the core metaphysical feature of freedom is being the ultimate source, or originator, of one's choices, and that being able to do otherwise is closely connected to this feature. For human beings or any created persons who owe their existence to factors outside themselves, the only way their acts of will could find their ultimate origin in themselves is for such acts not to be determined by their character and circumstances. For if all my willings were wholly determined, then if we were to trace my causal history back far enough, we would ultimately arrive at external factors that gave rise to me, with my particular genetic dispositions. My motives at the time would not be the ultimate source of my willings, only the most proximate ones. Only by there being less than deterministic connections between external influences and choices, then, is it be possible for me to be an ultimate source of my activity, concerning which I may truly say, “the buck stops here.”

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

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

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

However, William Rowe (1993) has countered that the thesis that there is no upper limit on goodness of worlds has a very different consequence: it shows that there could not be a morally perfect Creator! For suppose our world has an on-balance moral value of n and that God chose to create it despite being aware of possibilities having values higher than n that He was able to create. It seems we can now imagine a morally better Creator: one having the same options who chooses to create a better world. For a critical reply to Rowe, see the Howard-Snyders (1994) and Wainwright (1996). Rowe (2004) continues the discussion in response to a variety of views, both historical and contemporary.

Finally, Norman Kretzmann (1997, 220-25) has argued in the context of Aquinas's theological system that there is strong pressure to say that God must have created something or other, though it may well have been open to Him to create any of a number of contingent orders. The reason is that there is no plausible account of how an absolutely perfect God might have a resistible motivation—one consideration among other, competing considerations—for creating something rather than nothing. (It obviously cannot have to do with any sort of utility, for example.) The best general understanding of God's being motivated to create at all—one which in places Aquinas himself comes very close to endorsing—is to see it as reflecting the fact that God's very being, which is goodness, necessarily diffuses itself. Perfect goodness will naturally communicate itself outwardly; God who is perfect goodness will naturally create, generating a dependent reality that imperfectly reflects that goodness. (Wainwright (1996) is a careful discussion of a somewhat similar line of thought in Jonathan Edwards. See also Rowe 2004.)

Further Reading . Pink (2004) is an excellent, concise introduction. Pereboom (1997) provides good selections from a number of important historical writers on free will. Bourke (1964) and Dilman (1999) provide critical overviews of many such writers. For thematic treatments, see Fischer (1994); Kane (1996), esp. Ch.1-2; 5-6; Ekstrom (2001); Watson (2003b); and the outstanding collection of survey articles in Kane (2002).

  • Adams, Robert (1987). “Must God Create the Best?,” in The Virtue of Faith and Other Essays in Philosophical Theology . New York: Oxford University Press, 51-64.
  • Aquinas, Thomas (1945). Basic Writings of Saint Thomas Aquinas (2 vol.). New York: Random House.
  • ----- (1993). Selected Philosophical Writings , ed. T. McDermott. Oxford: Oxford University Press.
  • Aristotle (1985). Nichomachean Ethics , translated by Terence Irwin. Indianapolis: Hackett Publishing.
  • Augustine (1993). On the Free Choice of the Will , tr. Thomas Williams. Indianapolis: Hackett Publishing.
  • Ayer, A.J. (1982). “Freedom and Necessity,” in Watson (1982b), ed., 15-23.
  • Baker, Lynne (2000). Persons and Bodies: A Constitution View . Cambridge: Cambridge University Press.
  • Bourke, Vernon (1964). Will in Western Thought . New York: Sheed and Ward.
  • Campbell, C.A. (1967). In Defence of Free Will & other essays . London: Allen & Unwin Ltd.
  • Chisholm, Roderick (1982). “Human Freedom and the Self,” in Watson (1982b), 24-35.
  • ----- (1976). Person and Object . LaSalle: Open Court.
  • Clarke, Randolph (1993). “Toward a Credible Agent-Causal Account of Free Will,” in O'Connor (1995), ed., 201-15.
  • ----- (1995). “Indeterminism and Control,” American Philosophical Quarterly 32, 125-138.
  • ----- (1996). “Agent Causation and Event Causation in the Production of Free Action,” Philosophical Topics 24 (Fall), 19-48.
  • ----- (2003). Libertarian Accounts of Free Will . Oxford: Oxford University Press.
  • Dennett, Daniel (1984). Elbow Room: The Varieties of Free Will Worth Having . Cambridge. MA: MIT Press.
  • Descartes, René (1984). Meditations on First Philosophy [1641] and Passions of the Soul [1649], in The Philosophical Writings of Descartes , vol. I-III, translated by Cottingham, J., Stoothoff, R., & Murdoch, D.. Cambridge: Cambridge University Press.
  • Donagan, Alan (1985). Human Ends and Human Actions: An Exploration in St. Thomas's Treatment . Milwaukee: Marquette University Press.
  • Dilman, Ilham (1999). Free Will: An Historical and Philosophical Introduction . London: Routledge.
  • Double, Richard (1991). The Non-Reality of Free Will . New York: Oxford University Press.
  • Edwards, Jonathan (1957) [1754]. Freedom of Will , ed. P. Ramsey. New Haven: Yale University Press.
  • Ekstrom, Laura (2000). Free Will: A Philosophical Study . Boulder, CO: Westview Press.
  • Farrer, Austin (1958). The Freedom of the Will . London: Adam & Charles Black.
  • Fischer, John Martin (1994). The Metaphysics of Free Will . Oxford: Blackwell.
  • ----- (1999). “Recent Work on Moral Responsibility,” Ethics 110, 93-139.
  • Fischer, John Martin and Ravizza, Mark. (1992). “When the Will is Free,” in O'Connor (1995), ed., 239-269.
  • Fischer, John Martin (1998) Responsibility and Control . Cambridge: Cambridge University Press.
  • Frankfurt, Harry (1969). “Alternate Possibilities and Moral Responsibility,” Journal of Philosophy 66, 829-39.
  • ----- (1982). “Freedom of the Will and the Concept of a Person,” in Watson (1982), ed., 81-95.
  • ----- (1988). The Importance of What We Care About . Cambridge: Cambridge University Press.
  • ----- (1992). “The Faintest Passion,” Proceedings and Addresses of the American Philosophical Association 66, 5-16.
  • Ginet, Carl (1990). On Action . Cambridge: Cambridge University Press.
  • ----- (1997). “Freedom, Responsibility, and Agency,” The Journal of Ethics 1, 85-98.
  • ----- (2002) “Reasons Explanations of Action: Causalist versus Noncausalist Accounts,” in Kane, ed., (2002), 386-405.
  • Hobbes, Thomas and Bramhall, John (1999) [1655-1658]. Hobbes and Bramhall on Liberty and Necessity , ed. V. Chappell. Cambridge: Cambridge University Press.
  • Honderich, Ted (1988). A Theory of Determinism . Oxford: Oxford University Press.
  • Howard-Snyder, Daniel and Frances (1994). “How an Unsurpassable Being Can Create a Surpassable World,” Faith and Philosophy 11, 260-8.
  • Howard-Snyder, Daniel and Moser, Paul, eds. (2002). Divine Hiddenness: New Essays . Cambridge: Cambridge University Press.
  • Hume, David (1977) [1748]. An Enquiry Concerning Human Understanding . Indianapolis: Hackett Publishing.
  • Kane, Robert (1995). “Two Kinds of Incompatibilism,” in O'Connor (1995), ed., 115-150.
  • ----- (1996). The Significance of Free Will . New York: Oxford University Press.
  • Kane, Robert, ed., (2002). Oxford Handbook on Free Will . New York: Oxford University Press.
  • Kant, Immanuel (1993) [1788]. Critique of Practical Reason , tr. by Lewis White Beck. Upper Saddle River, NJ: Prentice-Hall Inc.
  • Kretzmann, Norman (1997). The Metaphysics of Theism: Aquinas's Natural Theology in Summa Contra Gentiles I. Oxford: Clarendon Press.
  • Leibniz, Gottfried (1985) [1710]. Theodicy . LaSalle, IL: Open Court.
  • Libet, Benjamin (2002). “Do We Have Free Will?” in Kane, ed., (2002), 551-564.
  • MacDonald, Scott (1998). “Aquinas's Libertarian Account of Free Will,” Revue Internationale de Philosophie , 2, 309-328.
  • Magill, Kevin (1997). Freedom and Experience . London: MacMillan.
  • McCann, Hugh (1998). The Works of Agency: On Human Action, Will, and Freedom . Ithaca: Cornell University Press.
  • Mele, Alfred (1995). Autonomous Agents (New York: Oxford University Press).
  • Morris, Thomas (1993). “Perfection and Creation,” in E. Stump. (1993), ed., 234-47
  • Murray, Michael (1993). “Coercion and the Hiddenness of God,” American Philosophical Quarterly 30, 27-38.
  • ----- (2002). “Deus Absconditus,” in Howard-Snyder amd Moser (2002), 62-82.
  • Nozick, Robert (1995). “Choice and Indeterminism,” in O'Connor (1995), ed., 101-14.
  • O'Connor, Timothy (1993). “Indeterminism and Free Agency: Three Recent Views,” Philosophy and Phenomenological Research , 53, 499-526.
  • -----, ed., (1995). Agents, Causes, and Events: Essays on Indeterminism and Free Will . New York: Oxford University Press.
  • ----- (2000). Persons and Causes: The Metaphysics of Free Will . New York: Oxford University Press.
  • ----- (forthcoming). “Freedom With a Human Face,” Midwest Studies in Philosophy , Fall 2005.
  • Pettit, Philip (2001). A Theory of Freedom . Oxford: Oxford University Press.
  • Pereboom, Derk (2001). Living Without Free Will . Cambridge: Cambridge University Press.
  • -----, ed., (1997). Free Will . Indianapolis: Hackett Publishing.
  • Pink, Thomas (2004). Free Will: A Very Short Introduction . Oxford: Oxford University Press.
  • Plato (1997). Complete Works , ed. J. Cooper. Indianapolis: Hackett Publishing.
  • Quinn, Phillip (1983). “Divine Conservation, Continuous Creation, and Human Action,” in A. Freddoso, ed. The Existence and Nature of God . Notre Dame: Notre Dame University Press.
  • Reid, Thomas (1969). Essays on the Active Powers of the Human Mind , ed. B. Brody. Cambridge: MIT Press.
  • Rowe, William (1993). “The Problem of Divine Perfection and Freedom,” in E. Stump (1993), ed., 223-33.
  • ----- (1995). “Two Concepts of Freedom,” in O'Connor (1995), ed. 151-71.
  • ----- (2004). Can God Be Free? . Oxford: Oxford University Press.
  • Sartre, John Paul (1956). Being and Nothingness . New York: Washington Square Press.
  • Schopenhauer, Arthur (1999) [1839]. Prize Essay on the Freedom of the Will , ed. G. Zoller. Cambridge: Cambridge University Press.
  • Scotus, John Duns (1986). “Questions on Aristotle's Metaphysics IX, Q.15” in Duns Scotus on the Will and Morality [selected and translated by Allan B. Wolter, O.F.M.]. Washington: Catholic University of America Press.
  • ----- (1994) [1297-99]. Contingency and Freedom: Lectura I 39 , tr. Vos Jaczn et al . Dordrecht: Kluwer Academic Publishers.
  • Shatz, David (1986). “Free Will and the Structure of Motivation,” Midwest Studies in Philosophy 10, 451-482.
  • Smilansky, Saul (2000). Free Will and Illusion . Oxford: Oxford University Press.
  • Strawson, Galen (1986). Freedom and Belief . Oxford: Clarendon Press.
  • Strawson, Peter (1982). “Freedom and Resentment,” in Watson (1982), ed., 59-80.
  • Stump, Eleonore (1996). “Persons: Identification and Freedom,” Philosophical Topics 24, 183-214.
  • -----(1997). “Aquinas's Account of Freedom: Intellect and Will,” The Monist 80, 576-597.
  • -----, ed., (1993). Reasoned Faith . Ithaca: Cornell University Press.
  • van Inwagen, Peter (1983). An Essay on Free Will . Oxford: Oxford University Press.
  • ----- (1994). “When the Will is Not Free,” Philosophical Studies , 75, 95-113.
  • ----- (1995). “When Is the Will Free?” in O'Connor (1995), ed., 219-238.
  • Wainwright, William (1996). “Jonathan Edwards, William Rowe, and the Necessity of Creation,” in J. Jordan and D. Howard-Snyder, eds., Faith Freedom, and Rationality . Lanham: Rowman and Littlefield, 119-133.
  • Wallace, R. Jay (2003). “Addiction as Defect of the Will: Some Philosophical Reflections,” in Watson, ed., (2003b), 424-452.
  • Watson, Gary (1987). “Free Action and Free Will,” Mind 96, 145-72.
  • ----- (2003a). “Free Agency,” in Watson, ed., 1982b.
  • -----, ed., (2003b). Free Will . 2nd ed. Oxford: Oxford University Press.
  • Wegner, Daniel (2002). The Illusion of Conscious Will . Cambridge, MA: MIT Press.
  • Widerker, David and McKenna, Michael, eds., (2002). Moral Responsibility and Alternative Possibilities . Aldershot: Ashgate Publishing.
  • Wolf, Susan (1990). Freedom Within Reason . Oxford: Oxford University Press.
  • The Garden of Forking Paths: A Free Will/Moral Responsibility Blog , multi-contributor (maintained at the University of California, Riverside)
  • The Determinism and Freedom Philosophy Website , edited by Ted Honderich (University College London)
  • Bibliography on Free Will , maintained by David Chalmers (Australian National University)

action | compatibilism | determinism: causal | fatalism | free will: divine foreknowledge and | incompatibilism: (nondeterministic) theories of free will | incompatibilism: arguments for | moral responsibility

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Humanities LibreTexts

4.3: The Illusion of Free Will

  • Last updated
  • Save as PDF
  • Page ID 29975

  • Golden West College via NGE Far Press

The Illusion of Free Will 35

80. Theologians repeatedly tell us, that man is free, while all their principles conspire to destroy his liberty. By endeavouring to justify the Divinity, they in reality accuse him of the blackest injustice. They suppose, that without grace, man is necessitated to do evil. They affirm, that God will punish him, because God has not given him grace to do good!

Little reflection will suffice to convince us, that man is necessitated in all his actions, that his free will is a chimera, even in the system of theologians. Does it depend upon man to be born of such or such parents? Does it depend upon man to imbibe or not to imbibe the opinions of his parents or instructors? If I had been born of idolatrous or Mahometan parents, would it have depended upon me to become a Christian? Yet, divines gravely assure us, that a just God will damn without pity all those, to whom he has not given grace to know the Christian religion!

Man's birth is wholly independent of his choice. He is not asked whether he is willing, or not, to come into the world. Nature does not consult him upon the country and parents she gives him. His acquired ideas, his opinions, his notions true or false, are necessary fruits of the education which he has received, and of which he has not been the director. His passions and desires are necessary consequences of the temperament given him by nature. During his whole life, his volitions and actions are determined by his connections, habits, occupations, pleasures, and conversations; by the thoughts, that are involuntarily presented to his mind; in a word, by a multitude of events and accidents, which it is out of his power to foresee or prevent. Incapable of looking into futurity, he knows not what he will do. From the instant of his birth to that of his death, he is never free. You will say, that he wills, deliberates, chooses, determines; and you will hence conclude, that his actions are free. It is true, that man wills, but he is not master of his will or his desires; he can desire and will only what he judges advantageous to himself; he can neither love pain, nor detest pleasure. It will be said, that he sometimes prefers pain to pleasure; but then he prefers a momentary pain with a view of procuring a greater and more durable pleasure. In this case, the prospect of a greater good necessarily determines him to forego a less considerable good.

The lover does not give his mistress the features which captivate him; he is not then master of loving, or not loving the object of his tenderness; he is not master of his imagination or temperament. Whence it evidently follows, that man is not master of his volitions and desires. "But man," you will say, "can resist his desires; therefore he is free." Man resists his desires, when the motives, which divert him from an object, are stronger than those, which incline him towards it; but then his resistance is necessary. A man, whose fear of dishonour or punishment is greater than his love of money, necessarily resists the desire of stealing.

"Are we not free, when we deliberate?" But, are we masters of knowing or not knowing, of being in doubt or certainty? Deliberation is a necessary effect of our uncertainty respecting the consequences of our actions. When we are sure, or think we are sure, of these consequences, we necessarily decide, and we then act necessarily according to our true or false judgment. Our judgments, true or false, are not free; they are necessarily determined by the ideas, we have received, or which our minds have formed.

Man is not free in his choice; he is evidently necessitated to choose what he judges most useful and agreeable. Neither is he free, when he suspends his choice; he is forced to suspend it until he knows, or thinks he knows, the qualities of the objects presented to him, or, until he has weighed the consequences of his actions. "Man," you will say, "often decides in favour of actions, which he knows must be detrimental to himself; man sometimes kills himself; therefore he is free." I deny it. Is man master of reasoning well or ill? Do not his reason and wisdom depend upon the opinions he has formed, or upon the conformation of his machine? As neither one nor the other depends upon his will, they are no proof of liberty. "If I lay a wager, that I shall do, or not do a thing, am I not free? Does it not depend upon me to do it or not?" No, I answer; the desire of winning the wager will necessarily determine you to do, or not to do the thing in question. "But, supposing I consent to lose the wager?" Then the desire of proving to me, that you are free, will have become a stronger motive than the desire of winning the wager; and this motive will have necessarily determined you to do, or not to do, the thing in question.

"But," you will say, "I feel free." This is an illusion, that may be compared to that of the fly in the fable, who, lighting upon the pole of a heavy carriage, applauded himself for directing its course. Man, who thinks himself free, is a fly, who imagines he has power to move the universe, while he is himself unknowingly carried along by it.

The inward persuasion that we are free to do, or not to do a thing, is but a mere illusion. If we trace the true principle of our actions, we shall find, that they are always necessary consequences of our volitions and desires, which are never in our power. You think yourself free, because you do what you will; but are you free to will, or not to will; to desire, or not to desire? Are not your volitions and desires necessarily excited by objects or qualities totally independent of you?

81. "If the actions of men are necessary, if men are not free, by what right does society punish criminals? Is it not very unjust to chastise beings, who could not act otherwise than they have done?" If the wicked act necessarily according to the impulses of their evil nature, society, in punishing them, acts necessarily by the desire of self-preservation. Certain objects necessarily produce in us the sensation of pain; our nature then forces us against them, and avert them from us. A tiger, pressed by hunger, springs upon the man, whom he wishes to devour; but this man is not master of his fear, and necessarily seeks means to destroy the tiger.

82. "If every thing be necessary, the errors, opinions, and ideas of men are fatal; and, if so, how or why should we attempt to reform them?" The errors of men are necessary consequences of ignorance. Their ignorance, prejudice, and credulity are necessary consequences of their inexperience, negligence, and want of reflection, in the same manner as delirium or lethargy are necessary effects of certain diseases. Truth, experience, reflection, and reason, are remedies calculated to cure ignorance, fanaticism and follies. But, you will ask, why does not truth produce this effect upon many disordered minds? It is because some diseases resist all remedies; because it is impossible to cure obstinate patients, who refuse the remedies presented to them; because the interest of some men, and the folly of others, necessarily oppose the admission of truth.

A cause produces its effect only when its action is not interrupted by stronger causes, which then weakens or render useless, the action of the former. It is impossible that the best arguments should be adopted by men, who are interested in error, prejudiced in its favour, and who decline all reflection; but truth must necessarily undeceive honest minds, who seek her sincerely. Truth is a cause; it necessarily produces its effects, when its impulse is not intercepted by causes, which suspend its effects.

83. "To deprive man of his free will," it is said, "makes him a mere machine, an automaton. Without liberty, he will no longer have either merit or virtue." What is merit in man? It is a manner of acting, which renders him estimable in the eyes of his fellow-beings. What is virtue? It is a disposition, which inclines us to do good to others. What can there be contemptible in machines, or automatons, capable of producing effects so desirable? Marcus Aurelius was useful to the vast Roman Empire. By what right would a machine despise a machine, whose springs facilitate its action? Good men are springs, which second society in its tendency to happiness; the wicked are ill-formed springs, which disturb the order, progress, and harmony of society. If, for its own utility, society cherishes and rewards the good, it also harasses and destroys the wicked, as useless or hurtful.

84. The world is a necessary agent. All the beings, that compose it, are united to each other, and cannot act otherwise than they do, so long as they are moved by the same causes, and endued with the same properties. When they lose properties, they will necessarily act in a different way. God himself, admitting his existence, cannot be considered a free agent. If there existed a God, his manner of acting would necessarily be determined by the properties inherent in his nature; nothing would be capable of arresting or altering his will. This being granted, neither our actions, prayers, nor sacrifices could suspend, or change his invariable conduct and immutable designs; whence we are forced to infer, that all religion would be useless.

85. Were not divines in perpetual contradiction with themselves, they would see, that, according to their hypothesis, man cannot be reputed free an instant. Do they not suppose man continually dependent on his God? Are we free, when we cannot exist and be preserved without God, and when we cease to exist at the pleasure of his supreme will? If God has made man out of nothing; if his preservation is a continued creation; if God cannot, an instant, lose sight of his creature; if whatever happens to him, is an effect of the divine will; if man can do nothing of himself; if all the events, which he experiences, are effects of the divine decrees; if he does no good without grace from on high, how can they maintain, that a man enjoys a moment's liberty? If God did not preserve him in the moment of sin, how could man sin? If God then preserves him, God forces him to exist, that he may sin.

University of Notre Dame

Notre Dame Philosophical Reviews

  • Home ›
  • Reviews ›

Exploring the Illusion of Free Will and Moral Responsibility

Placeholder book cover

Gregg D. Caruso (ed.), Exploring the Illusion of Free Will and Moral Responsibility , Lexington Books, 2013, 324pp., $85.00 (hbk), ISBN 9780739177310.

Reviewed by Brian Robinson, Grand Valley State University

Gregg Caruso has brought together a distinguished collection of contributors to lead the reader through an investigation of the arguments for, implications of, and potential responses to the rise of skepticism about free will. Free will skepticism, which Caruso tells us has become increasingly popular in recent decades, questions the existence of free will without necessarily adopting the truth of causal determinism. There are, as the authors of the sixteen chapters point out, a wide range of philosophical and scientific reasons for regarding free will as an illusion. As the title suggests though, free will is only half of the illusion. To doubt the existence of free will is typically to doubt the existence of moral responsibility. Hence, they are tied together as one collective illusion.

The book is divided into two sections. The first eleven chapters that form Part I are philosophical arguments for skepticism about free will and moral responsibility. The remaining five chapters focus on recent scientific research concerning free will and moral responsibility, including the related topics of consciousness, intentionality, and agency. While this division is natural and generally helpful, it could obscure the nature of the chapters in each section. Some of the chapters in Part I discuss empirical findings. Part II, meanwhile, is not by any means bereft of philosophical substance. Within Part I, several of the chapters are grouped with others that focus on related topics. Manuel Vargas (chapter 10), for instance, re-iterates his call for the possibility and necessity of a revised account of free will that remains empirically plausible. Shaun Nichols (chapter 11) then situates Vargas's position in the larger context of a dispute about whether erroneous folk notions should be eliminated or revised. Likewise, Saul Smilansky (chapter 6), Thomas Nadelhoffer and Daniela Goya Tocchetto (chapter 7), Benjamin Vilhauer (chapter 8), and Susan Blackmore (chapter 9) all deal with whether humans should and can live without their illusion of free will and moral responsibility. One almost wonders why there were not subsections to highlight these topical groupings. It is also worth pointing out that while this collection is not aimed at the philosophical novice, the review of the behavioral, cognitive, and neuroscientific material in Part II is not overly technical, making it readily accessible to a non-scientific audience.

The free will debate admittedly is one that, even among some professional philosophers, elicits eye-rolling or not-so-subtle attempts to change the topic of conversation. Some feel that this debate stagnated long ago and so have stopped paying attention. For those who have tuned out the free will literature, there are a few chapters that are novel and well worth your attention. Then for the free will aficionados, at least most of the chapters will be of interest. And for anyone considering assigning this book or portions of it for an upper-level or graduate course on free will, there are several chapters that will be informative and catch students up on much of the state of the debate. This is, in fact, one of the strengths of the collection; there is something here for everyone.

I'll begin by reviewing the chapters that should be of interest to the widest audience. Nichols's outstanding contribution (chapter 11) begins with the assumption that the folk notion of free will is in error because it presupposes the truth of indeterminism. The question is then whether the concept of free will should be revised or eliminated. Since free will is obviously not the first natural kind term to face this problem, Nichols notes that all such debates boil down to whether or not the erroneous folk term in question successfully refers or not. Eliminativists employ a descriptivist convention for reference, according to which a natural kind term refers by means of an associated description. Revisionists, however, maintain a causal-historical convention for reference, according to which an initial baptism set the referent of a term, even if the nature of the object referred to was misunderstood at the time of baptism. Nichols then offers recent experimental evidence that speakers regularly use different referential conventions for different kinds of sentences in which natural kind terms appear. So, when it comes to free will, it is correct to say, "Free will does not exist," since people employ the descriptivist convention in sentences like this one. Nevertheless, it is also correct to say, "Free will isn't what we thought," since a causal-historical convention of reference can be employed for sentences of this type. With this distinction between conventions of reference, both sides can agree that libertarian free will does not exist without precluding the possibility of revision. What is especially appealing about Nichols's "geography of error" and his pluralistic approach is that it can readily be applied to similar debates about other natural kind terms that philosophers care about.

In another intriguing contribution, Bruce Waller argues for a diagnosis of the psychological causes of the persistence of the belief in moral responsibility. Philosophers and the folk, Waller claims, are highly reluctant to abandon their deep-seated commitment to moral responsibility, despite ever growing evidence to the contrary. Thus, Waller takes as an assumption that moral responsibility is an illusion, and instead of arguing why we should give it up, tries to determine why we cling to it so desperately. The main reason, he argues, is a correlated, nonconscious belief in a just world. Philosophers may consciously reject the belief in a just world, but according to Waller it still regularly exerts itself nonconsciously. An example of this nonconscious belief in a just world at work is when people blame rape victims as having invited the attacks. Instances of rape threaten one's belief in a just world, so blaming the victim is a way to preserve that belief. Waller references a number of psychological studies establishing the presence and effectiveness of the belief in a just world. The argument, however, linking it with the belief in moral responsibility is a conceptual one. It would seem that Waller's thesis is empirically testable, and so it will be interesting to see how his claim bears up under future investigation.

The final chapter that is likely to be of interest to those not already heavily engaged in the contemporary free will debate is Mark Hallett's brief but effective discussion of what and when our brains know about what we are doing (chapter 14). Many are familiar with the seminal work of Libet et al. (1983), who found brain activity that caused a bodily movement one second prior to participants reporting that they willed the bodily movement. Hallett surveys this research and the directions in which it has progressed in the subsequent three decades. He then rightly concludes that moral responsibility cannot require consciousness.

The next contributions I'd like to highlight are those that will be of particular interest to free will aficionados. Nadelhoffer and Tocchetto (chapter 7) enter the discussion of the usefulness of disabusing the folk of their erroneous notion of free will. As they note, a number of studies have found that inducing free will skepticism in people correlates with anti-social behavior (such as cheating and aggressive behavior). The implication of these studies is that there are good reasons for letting people continue to believe in nonexistent free will. Nadelhoffer and Tocchetto worry, however, that not enough attention has been paid to the negative effects of believing in free will. They report their results from two studies that find a dark side of free will, namely correlations between free will beliefs and both right wing authoritarianism and just world beliefs. These results are without doubt a significant contribution. Nevertheless, one is left slightly unsure how to fully interpret their findings. Their results only partially matched their predictions, as Nadelhoffer and Tocchetto admit. It is also unclear whether appropriate methodologies were used to limit the Type I error rate (the likelihood of false positives), since they were testing multiple predictions simultaneously using five, multi-question scales in each study. It is also unclear from their present research if inducing free will skepticism in those endorsing right wing authoritarianism would lead to a corresponding drop in right wing authoritarianism. This research program looks to be a promising one, and I look forward to their future work.

John-Dylan Haynes and Michael Pauen (chapter 12) offer an engaging exploration of how and why the concept of intentionality must be revised. For instance, recent neuroscientific evidence indicates that we perform a number of intentional actions without conscious awareness, contrary to the commonsense notion of intentionality that requires conscious awareness. After reviewing the extensive empirical challenges to this commonsense view of intentions, Haynes and Pauen offer a "modest" revision of the concept that comports with psychology and neuroscience. Despite the fact that this modest revision will initially strike most as deeply counter-intuitive, Haynes and Pauen make a compelling case, though the full implications and details of their revision remain to be worked out. One must also appreciate that the scientific material is well presented to a lay readers.

Neil Levy (chapter 5) delivers a noteworthy contribution that continues a specific exchange he and Tamler Sommers have recently begun. Sommers (2012), responding to Levy's (2011) Hard Luck , asserts that by appealing to intuitions about moral responsibility, we do not discover any necessary or sufficient conditions for moral responsibility. Instead, we merely systematize our conceptions, since intuitions about moral responsibility are necessarily culturally relative. Levy here replies that Sommers's meta-skepticism collapses into Levy's own regular skepticism about moral responsibility. Furthermore, Sommers's meta-skepticism undercuts itself, Levy maintains, by relying on culturally relative intuitions as well. Levy has marshaled some intriguing arguments, though Sommers has a number of responses left open to him. So this debate is far from resolved, and it will be interesting to see how it progresses.

Derk Pereboom (chapter 1) has for some time argued for hard indeterminism and continues that argument here. Whether or not causal determinism is true, Pereboom claims, makes no difference to the impossibility of moral responsibility (in the sense of basic deserts). The chapter at times appears to be covering a lot of ground at once (perhaps too much), though it is understandable given that Pereboom is condensing a great deal of his previous work into one chapter. Still, it can serve as a superb text to quickly acquaint students with his view.

Despite the overall quality of the collection, there are a few chapters that all but the most devoted free will scholars can skip without missing much. Galen Strawson (chapter 2), for instance, essentially re-iterates his Basic Argument for the impossibility of moral responsibility. I do not mean to impugn the quality of Strawson's work. Rather, as he has made this argument elsewhere and for sometime, there is not enough new here to attract most readers' attention. Susan Blackmore (chapter 9) argues from personal experience that it is possible (and preferable) to live without the illusion of free will. While I do not doubt the truth of her experiences, her argument apparently extends to the claim that we all should live without the illusion, and this larger argument seems problematic. As Blackmore notes, a number of prominent free will skeptics have told her that they still live as if they have free will because they have to. The trouble with this apparent further claim is that she denies that we have contra-causal free will. So if others live as if they have contra-causal free will while intellectually denying its existence, then by Blackmore's logic, they cannot do otherwise. There seems little that would be capable of causing them to change. If instead Blackmore's contribution is intended only as a personal mediation on living with free will, it can be an interesting read to those wondering if it possible for them as well.

Returning to the collection as a whole, a few concluding remarks are warranted. Given that all but one work is original to this collection and that several chapters focus on various related topics within the general exploration of the illusion of free will and moral responsibility, the reader may sometimes long for more explicit conversation between the authors. At times, it would have been informative and engaging if the author(s) for one chapter were explicitly responding to or developing a thought from another chapter. Instead the book is a collection of independent, though related essays. This is not to criticize any of the chapters on their own merit. Yet, for the free will devotees, due to this editorial decision the collection does give the slight feeling of a missed opportunity for a conversation between so many experts. On the whole, this collection is an excellent contribution to the ongoing free will debate. Those interested or engaged in that debate will find much to appreciate (or argue vociferously against), including those chapters that could not be fit into this review. And even for those readers who only occasionally re-acquaint themselves with this debate, there are certain chapters well worth their time.

Levy, N. (2011). Hard Luck: How Luck Undermines Free Will and Moral Responsibility . Oxford University Press.

Libet, B., Gleason, C. A., Wright, E. W., and Pearl, D. K. (1983). "Time of Conscious Intention to Act in Relation to Onset of Cerebral Activity (Readiness-Potential): The Unconscious Initiation of a Freely Voluntary Act." Brain 106.3: 623-42.

Sommers, T. (2012). Relative Justice: Cultural Diversity, Free Will, and Moral Responsibility . Princeton University Press.

Advertisement

It's not an illusion, you have free will. It's just not what you think

The idea that free will doesn't exist is based on misguided intuitions of what it means to be a biological machine, as a famous insect, the digger wasp, reveals

By Tom Stafford

3 April 2019

arrows

plainpicture

A SIMPLE insect can help us understand free will, and the lack of it. When a female digger wasp is ready to lay her eggs, she hunts down a cricket or similar prey, paralyses it with a sting, drags it back to the lip of her burrow, and then enters to check for blockages. If you move the cricket a few centimetres away before she re-emerges, she will again drag it to the threshold and again leave it to check for blockages. She will do this over and over. The wasp has no choice. This mindlessly inflexible behaviour has led to the wasp,  Sphex ichneumoneus , becoming a byword among biologists for determinism, the idea that what we think of as a “choice” is in fact a path dictated by pre-existing factors.

It is tempting to think that we aren’t like the wasp – that what we do is the result of choices that are freely made. Yet the more we learn about the neuroscience of decision-making, the more “sphexish” we seem to be. You hear people arguing that humans are mere biological machines trapped in cycles of behaviour that are ultimately beyond our control – that free will is just an illusion.

As a cognitive scientist who studies decision-making, I disagree. Of course, humans are animals. The problem, I believe, is our misguided intuitions of what it means to be a biological machine. In an attempt to dispel some of these misconceptions, I have created an interactive essay on Twitter called The Choice Engine.

Waspish behaviour

How Sphex came to be linked with free will is a long story. Charles Darwin…

Article amended on 9 April 2019

Sign up to our weekly newsletter.

Receive a weekly dose of discovery in your inbox! We'll also keep you up to date with New Scientist events and special offers.

To continue reading, subscribe today with our introductory offers

No commitment, cancel anytime*

Offer ends 2nd of July 2024.

*Cancel anytime within 14 days of payment to receive a refund on unserved issues.

Inclusive of applicable taxes (VAT)

Existing subscribers

More from New Scientist

Explore the latest news, articles and features

Has Neuralink made a breakthrough in brain implant technology?

Human brains have been mysteriously preserved for thousands of years, genetics may protect against disease linked to eating human brains, how manners can be a weapon to divide and disempower.

Subscriber-only

Popular articles

Trending New Scientist articles

Why the Classical Argument Against Free Will Is a Failure

is free will an illusion essay

In the last several years, a number of prominent scientists have claimed that we have good scientific reason to believe that there’s no such thing as free will — that free will is an illusion. If this were true, it would be less than splendid. And it would be surprising, too, because it really seems like we have free will. It seems that what we do from moment to moment is determined by conscious decisions that we freely make.

We need to look very closely at the arguments that these scientists are putting forward to determine whether they really give us good reason to abandon our belief in free will. But before we do that, it would behoove us to have a look at a much older argument against free will — an argument that’s been around for centuries.

is free will an illusion essay

The older argument against free will is based on the assumption that determinism is true. Determinism is the view that every physical event is completely caused by prior events together with the laws of nature. Or, to put the point differently, it’s the view that every event has a cause that makes it happen in the one and only way that it could have happened.

If determinism is true, then as soon as the Big Bang took place 13 billion years ago, the entire history of the universe was already settled. Every event that’s ever occurred was already predetermined before it occurred. And this includes human decisions. If determinism is true, then everything you’ve ever done — every choice you’ve ever made — was already predetermined before our solar system even existed. And if this is true, then it has obvious implications for free will.

Suppose that you’re in an ice cream parlor, waiting in line, trying to decide whether to order chocolate or vanilla ice cream. And suppose that when you get to the front of the line, you decide to order chocolate. Was this choice a product of your free will? Well, if determinism is true, then your choice was completely caused by prior events. The immediate causes of the decision were neural events that occurred in your brain just prior to your choice. But, of course, if determinism is true, then those neural events that caused your decision had physical causes as well; they were caused by even earlier events — events that occurred just before they did. And so on, stretching back into the past. We can follow this back to when you were a baby, to the very first events of your life. In fact, we can keep going back before that, because if determinism is true, then those first events were also caused by prior events. We can keep going back to events that occurred before you were even conceived, to events involving your mother and father and a bottle of Chianti.

If determinism is true, then as soon as the Big Bang took place 13 billion years ago, the entire history of the universe was already settled.

So if determinism is true, then it was already settled before you were born that you were going to order chocolate ice cream when you got to the front of the line. And, of course, the same can be said about all of our decisions, and it seems to follow from this that human beings do not have free will.

Let’s call this the classical argument against free will . It proceeds by assuming that determinism is true and arguing from there that we don’t have free will.

There’s a big problem with the classical argument against free will. It just assumes that determinism is true. The idea behind the argument seems to be that determinism is just a commonsense truism. But it’s actually not a commonsense truism. One of the main lessons of 20th-century physics is that we can’t know by common sense, or by intuition, that determinism is true. Determinism is a controversial hypothesis about the workings of the physical world. We could only know that it’s true by doing some high-level physics. Moreover — and this is another lesson of 20th-century physics — as of right now, we don’t have any good evidence for determinism. In other words, our best physical theories don’t answer the question of whether determinism is true.

During the reign of classical physics (or Newtonian physics), it was widely believed that determinism was true. But in the late 19th and early 20th centuries, physicists started to discover some problems with Newton’s theory, and it was eventually replaced with a new theory — quantum mechanics . (Actually, it was replaced by two new theories, namely, quantum mechanics and relativity theory. But relativity theory isn’t relevant to the topic of free will.) Quantum mechanics has several strange and interesting features, but the one that’s relevant to free will is that this new theory contains laws that are probabilistic rather than deterministic. We can understand what this means very easily. Roughly speaking, deterministic laws of nature look like this:

If you have a physical system in state S, and if you perform experiment E on that system, then you will get outcome O.

But quantum physics contains probabilistic laws that look like this:

If you have a physical system in state S, and if you perform experiment E on that system, then there are two different possible outcomes, namely, O1 and O2; moreover, there’s a 50 percent chance that you’ll get outcome O1 and a 50 percent chance that you’ll get outcome O2.

It’s important to notice what follows from this. Suppose that we take a physical system, put it into state S, and perform experiment E on it. Now suppose that when we perform this experiment, we get outcome O1. Finally, suppose we ask the following question: “Why did we get outcome O1 instead of O2?” The important point to notice is that quantum mechanics doesn’t answer this question . It doesn’t give us any explanation at all for why we got outcome O1 instead of O2. In other words, as far as quantum mechanics is concerned, it could be that nothing caused us to get result O1 ; it could be that this just happened .

Now, Einstein famously thought that this couldn’t be the whole story. You’ve probably heard that he once said that “God doesn’t play dice with the universe.” What he meant when he said this was that the fundamental laws of nature can’t be probabilistic. The fundamental laws, Einstein thought, have to tell us what will happen next, not what will probably happen, or what might happen. So Einstein thought that there had to be a hidden layer of reality , below the quantum level, and that if we could find this hidden layer, we could get rid of the probabilistic laws of quantum mechanics and replace them with deterministic laws, laws that tell us what will happen next, not just what will probably happen next. And, of course, if we could do this — if we could find this hidden layer of reality and these deterministic laws of nature — then we would be able to explain why we got outcome O1 instead of O2.

But a lot of other physicists — most notably, Werner Heisenberg and Niels Bohr — disagreed with Einstein. They thought that the quantum layer of reality was the bottom layer. And they thought that the fundamental laws of nature — or at any rate, some of these laws — were probabilistic laws. But if this is right, then it means that at least some physical events aren’t deterministically caused by prior events. It means that some physical events just happen . For instance, if Heisenberg and Bohr are right, then nothing caused us to get outcome O1 instead of O2; there was no reason why this happened; it just did .

The debate between determinists like Einstein and indeterminists like Heisenberg and Bohr has never been settled.

The debate between Einstein on the one hand and Heisenberg and Bohr on the other is crucially important to our discussion. Einstein is a determinist. If he’s right, then every physical event is predetermined — or in other words, completely caused by prior events. But if Heisenberg and Bohr are right, then determinism is false . On their view, not every event is predetermined by the past and the laws of nature; some things just happen , for no reason at all. In other words, if Heisenberg and Bohr are right, then indeterminism is true.

And here’s the really important point for us. The debate between determinists like Einstein and indeterminists like Heisenberg and Bohr has never been settled. We don’t have any good evidence for either view. Quantum mechanics is still our best theory of the subatomic world, but we just don’t know whether there’s another layer of reality, beneath the quantum layer. And so we don’t know whether all physical events are completely caused by prior events. In other words, we don’t know whether determinism or indeterminism is true. Future physicists might be able to settle this question, but as of right now, we don’t know the answer.

But now notice that if we don’t know whether determinism is true or false, then this completely undermines the classical argument against free will. That argument just assumed that determinism is true. But we now know that there is no good reason to believe this. The question of whether determinism is true is an open question for physicists. So the classical argument against free will is a failure — it doesn’t give us any good reason to conclude that we don’t have free will.

Despite the failure of the classical argument, the enemies of free will are undeterred. They still think there’s a powerful argument to be made against free will. In fact, they think there are two such arguments. Both of these arguments can be thought of as attempts to fix the classical argument, but they do this in completely different ways.

The first new-and-improved argument against free will — which is a scientific argument — starts with the observation that it doesn’t matter whether the full-blown hypothesis of determinism is true because it doesn’t matter whether all events are predetermined by prior events. All that matters is whether our decisions are predetermined by prior events. And the central claim of the first new-and-improved argument against free will is that we have good evidence (from studies performed by psychologists and neuroscientists) for thinking that, in fact, our decisions are predetermined by prior events.

The second new-and-improved argument against free will — which is a philosophical argument, not a scientific argument — relies on the claim that it doesn’t matter whether determinism is true because in determinism is just as incompatible with free will as determinism is. The argument for this is based on the claim that if our decisions aren’t determined, then they aren’t caused by anything, which means that they occur randomly . And the central claim of the second new-and-improved argument against free will is that if our decisions occur randomly, then they just happen to us , and so they’re not the products of our free will.

My own view is that neither of these new-and-improved arguments succeeds in showing that we don’t have free will. But it takes a lot of work to undermine these two arguments. In order to undermine the scientific argument, we need to explain why the relevant psychological and neuroscientific studies don’t in fact show that we don’t have free will. And in order to undermine the philosophical argument, we need to explain how a decision could be the product of someone’s free will — how the outcome of the decision could be under the given person’s control — even if the decision wasn’t caused by anything.

So, yes, this would all take a lot of work. Maybe I should write a book about it.

Mark Balaguer is Professor in the Department of Philosophy at California State University, Los Angeles. He is the author of several books, including “ Free Will ,” from which this article is adapted.

  • Reference Manager
  • Simple TEXT file

People also looked at

Hypothesis and theory article, free will and neuroscience: from explaining freedom away to new ways of operationalizing and measuring it.

is free will an illusion essay

  • Neuroethics, Centro Universitario Internazionale, Arezzo, Italy

The concept of free will is hard to define, but crucial to both individual and social life. For centuries people have wondered how freedom is possible in a world ruled by physical determinism; however, reflections on free will have been confined to philosophy until half a century ago, when the topic was also addressed by neuroscience. The first relevant, and now well-known, strand of research on the brain correlates of free will was that pioneered by Libet et al. (1983) , which focused on the allegedly unconscious intentions taking place in decisions regarded as free and voluntary. Libet’s interpretation of the so-called readiness potential (RP) seems to favor a sort of deflation of freedom ( Soon et al., 2008 ). However, recent studies seem to point to a different interpretation of the RP, namely that the apparent build-up of the brain activity preceding subjectively spontaneous voluntary movements (SVM) may reflect the ebb and flow of the background neuronal noise, which is triggered by many factors ( Schurger et al., 2016 ). This interpretation seems to bridge the gap between the neuroscientific perspective on free will and the intuitive, commonsensical view of it ( Roskies, 2010b ), but many problems remain to be solved and other theoretical paths can be hypothesized. The article therefore, proposes to start from an operationalizable concept of free will ( Lavazza and Inglese, 2015 ) to find a connection between higher order descriptions (useful for practical life) and neural bases. This new way to conceptualize free will should be linked to the idea of “capacity”: that is, the availability of a repertoire of general skills that can be manifested and used without moment by moment conscious control. The capacity index, which is also able to take into account the differences of time scales in decisions, includes reasons-responsiveness and is related to internal control, understood as the agent’s ownership of the mechanisms that trigger the relevant behavior. Cognitive abilities, needed for one to have capacity, might be firstly operationalized as a set of neuropsychological tests, which can be used to operationalize and measure specific executive functions, as they are strongly linked to the concept of control. Subsequently, a free will index would allow for the search of the underlying neural correlates of the capacity exhibited by people and the limits in capacity exhibited by each individual.

Introduction—Free Will as a Problem (Not Only) for Science

The concept of free will is hard to define, but crucial to both individual and social life ( Kane, 2005 ). Free will can be the reason why someone is not sent to jail during a trial upon appealing to insanity: the subject was not “free” when they committed the crime, not because someone was pointing a gun to their head, but because a psychiatric illness prevented them from controlling their actions. According to a long-standing philosophical tradition, if someone was not “free” when they did something, they cannot be held responsible for their deed ( Glannon, 2015 ). And the freedom in question is both “social” freedom (linked to constraints imposed by our peers or by external factors), and the one indicated by the term free will .

Free will can be defined by three conditions ( Walter, 2001 ). The first one is the “ability to do otherwise.” This is an intuitive concept: to be free, one has to have at least two alternatives or courses of action between which to choose. If one has an involuntary spasm of the mouth, for example, one is not in the position to choose whether to twist one’s mouth or not. The second condition is the “control over one’s choices.” The person who acts must be the same who decides what to do. To be granted free will, one must be the author of one’s choices, without the interference of people and of mechanisms outside of one’s reach. This is what we call agency, that is, being and feeling like the “owner” of one’s decisions and actions. The third condition is the “responsiveness to reasons”: a decision can’t be free if it is the effect of a random choice, but it must be rationally motivated. If I roll a dice to decide whom to marry, my choice cannot be said to be free, even though I will freely choose to say “I do”. On the contrary, if I choose to marry a specific person for their ideas and my deep love for them, then my decision will be free.

Thus defined, free will is a kind of freedom that we are willing to attribute to all human beings as a default condition. Of course there are exceptions: people suffering from mental illness and people under psychotropic substances ( Levy, 2013 ). Nevertheless, the attribution of free will as a general trend does not imply that all decisions are always taken in full freedom, as outlined by the three conditions illustrated above: “We often act on impulse, against our interests, without being fully aware of what we are doing. But this does not imply that we are not potentially able to act freely. Ethics and law have incorporated these notions, adopting the belief that usually people are free to act or not to act in a certain way and that, as a result, they are responsible for what they do, with the exceptions mentioned above” ( Lavazza and Inglese, 2015 ).

It is commonly experienced that the conditions of “ability to do otherwise”, “control” and “responsiveness to reasons” are very rarely at work all at once. Moreover, they would require further discussion, because there is wide disagreement on those conditions as regards their definition and scope ( Kane, 2016 ). But for the purposes of this article, this introductory treatment should suffice. In fact, the description of free will that I have sketched here is the one that dominated the theoretical discourse on, and practical applications of, the evaluation of human actions. From a philosophical point of view, however, starting with Plato, the main problem has been that of the actual existence of freedom, beyond the appearances and the insights that guide our daily life. The main challenge to free will has been determinism: the view that everything that happens (human decisions and actions included) is the consequence of sufficient conditions for its occurrence ( Berofsky, 2011 ). More specifically, “It is the argument that all mental phenomena and actions are also, directly or indirectly, causally produced—according to the laws of nature (such as those of physics and neurobiology)—by previous events that lie beyond the control of the agents” ( Lavazza and Inglese, 2015 ). Determinism was first a philosophical position; then, the birth of Galilean science—founded on the existence of immutable laws that are empirically verifiable—has increased its strength, giving rise to the concept of incompatibilism, namely the idea that free will and natural determinism cannot coexist. Only one of them can be true.

Throughout the centuries, despite its conceptual progress, philosophy hasn’t been able to solve this dilemma. As a result, today there are different irreconcilable positions about human free will: determinism is not absolute and free will exists; free will does not exist for a number of reasons, first of all (but not only) determinism; free will can exist even if determinism is true ( Kane, 2011 ). A little more than 30 years ago, neuroscience and empirical psychology came into play. Although biological processes cannot be considered strictly deterministic on the observable level of brain functioning (nerve signal transmission), new methods of investigation of the brain, more and more precise, have established that the cerebral base is a necessary condition of behavior and even of mental phenomena. On the basis of these acquisitions, neuroscience has begun to provide experimental contributions to the debate on free will.

In order to better understand the neural bases of free will, provided that there are any, in this article I’ll review and integrate findings from studies in different fields (philosophy, cognitive neuroscience, experimental and clinical psychology, neuropsychology). Unlike previous reviews on free will and neuroscience ( Haggard, 2008 , 2009 ; Passingham et al., 2010 ; Roskies, 2010a ; Brass et al., 2013 ), I have no claim of being exhaustive. My goal is to highlight a paradigm shift in the analysis and interpretation of the brain determinants preceding and/or causing free or voluntary action ( Haggard, 2008 takes voluntary decision to be non-stimulus driven, as much as possible). Firstly, following Libet’s experiments, a widespread interpretation of the so-called readiness potential (RP) went in the direction of a deflation of freedom ( Crick, 1994 ; Greene and Cohen, 2004 ; Cashmore, 2010 ; Harris, 2012 ). Indeed, the discovery of the role of the RP has been taken as evidence of the fact that free will is an illusion, since it seems that specific brain areas activate before we are aware of the onset of the movement. However, recent studies seem to point to a different interpretation of the RP, namely that the apparent build-up of the brain activity preceding subjectively spontaneous voluntary movements (SVM) may reflect the ebb and flow of the background neuronal noise, which is triggered by many factors ( Schurger et al., 2016 ). This interpretation seems to bridge, at least partially, the gap between the neuroscientific perspective on free will and the intuitive, commonsensical view of it ( Roskies, 2010b ), but many problems remain to be solved and other theoretical paths can be hypothesized. After analyzing the change of paradigm of these perspectives, I’ll propose to start from an operationalizable concept of free will ( Lavazza and Inglese, 2015 ) to find a connection between higher order descriptions (useful for practical life) and neural bases.

Neuroscience: Purporting to Explain Free Will

The discovery of the readiness potential.

As a preliminary consideration, it is important to underline that the idea of using an experiment (or a series of experiments) to establish whether the human being can be said to have free will implies accepting a direct link between a measurement of brain functioning and a pre-existing theoretical construct. This direct connection, as it is known, presents several problems and as we shall see, needs conceptual refinement to avoid simplifications and unfounded claims. What one can see and measure in brain activity may in fact only grasp a part of the idea of free will that we would like to test. This was one of the main criticisms to the experiments conducted so far ( Mele, 2009 ; Nachev and Hacker, 2014 ). What is measured at the level of brain functioning in the laboratory does not match the concept of free will we refer to, for example, to determine whether someone who engaged in violent behavior could have done otherwise in that specific circumstance.

The first relevant, and now well-known, strand of research on the brain correlates of free will was that pioneered by Libet et al. (1983) , which focused on the allegedly unconscious intentions affecting decisions regarded as free and voluntary. It should be noted that the concepts involved—“conscious intentions”, “voluntary decisions”, “free decisions”—have no clear and shared definition ( Nachev and Hacker, 2014 ), and the experiments themselves have been differently interpreted and often criticized ( Lavazza and De Caro, 2010 ). In any case, Libet’s experiments and their variants have been repeated several times until very recently, confirming their findings with a sufficient degree of reliability.

Libet based his work on Kornhuber and Deecke’s (1965) discovery of the bereitschaftpotential : the RP, a slow build-up of a scalp electrical potential (of a few microvolts), mainly measured through electroencephalography (EEG), that precedes the onset of subjectively SVM ( Kornhuber and Deecke, 1965 ). According to its discoverers, the RP is “the electro-physiological sign of planning, preparation, and initiation of volitional acts” ( Kornhuber and Deecke, 1990 ). “The neurobiologist John Eccles speculated that the subject must become conscious of the intention to act before the onset of this RP. Libet had the idea that he should test Eccles’s prediction” ( Doyle, 2011 ).

In his experiments, Libet invited the participants to move their right wrist and to report the precise moment when they had the impression that they decided to do so, thanks to a big clock they had in front of them ( Libet et al., 1983 ). In this way, it was possible to estimate the time of awareness with respect to the beginning of the movement, measured using an electromyogram (which records the muscle contraction). During the execution of the task, brain electrical activity was recorded through electrodes placed on the participants’ scalps. The attention was focused on a specific negative brain potential, namely the RP, originated from the supplementary motor area (SMA): a brain area involved in motor preparation, which is visible in the EEG signal as a wave that starts before any voluntary movement, while being absent or reduced before involuntary and automatic movements.

When one compares the subjective “time” of decision and what appeared at a cerebral level, the result appears as a striking blow to the traditional view of free will ( Libet, 1985 , 2004 ). In the experiment, the RP culminating in the execution of the movement starts in the prefrontal motor areas long before the time when the subject seems to have made the decision: participants became aware of their intention to take action about 350 ms after the onset of such potential. The volitional process is detected to start unconsciously 550 ms before the action is made in the case of non-preplanned acts and 1000 ms before in the case of preplanned acts. Thus these findings seem to show that our simple actions (and therefore, potentially, also more complex ones) are triggered by unconscious neural activity and that the awareness of those actions only occurs at a later time, when we think we are willing to act.

In the first phase of its intervention in the debate on free will, therefore, neuroscience seemed to argue for a deflation of freedom. Neuroscientists identified a specific aspect of the notion of freedom (the conscious control of the start of the action) and researched it: the experimental results seemed to indicate that there is no such conscious control, hence the conclusion that free will does not exist. However, it is important to highlight that this interpretation strongly depends on the idea that free choices or actions are fully internally generated, in the sense that they are not externally determined—where “external” means outside the subject’s conscience and the subject is something akin to the self. As we shall see, though, this distinction seems to be neither relevant nor truly informative when considering if and how choices are free.

In fact, Libet left the subject some time to veto: about 150 ms. This is the time needed for the muscles to flex in response to the command of the primary motor cortex (M2) through the spinal motor nerve cells. In the last 50 ms the action is realized with its external manifestations (bending the wrist) without any more possible intervention by the prefrontal brain areas (see “The Veto Power” Section). Libet thought there was a role for conscious will precisely in this situation: conscious will can let the action go to completion or it can block it with the explicit veto of the movement implemented by the prefrontal areas ( Doyle, 2011 ). But the intentional inhibition of an action (a decision itself) is preceded by neural activity as well ( Filevich et al., 2012 , 2013 ). So it cannot be a completely different decision from that to take a positive decision to act.

In their experiments, Haggard and Eimer (1999) used Libet’s method, but asked the participants to perform a different task. They had to move at will either the right index finger or the left in a series of repeated trials. The authors have compared the RP and the lateralized readiness potential (LRP) in trials in which awareness appeared in shorter or longer time, that is, considering the latency of awareness compared to the RP. In their words, “the RP tended to occur later on trials with early awareness of movement initiation than on trials with late awareness, ruling out the RP as a cause of our awareness of movement initiation. However, the LRP occurred significantly earlier on trials with early awareness than on trials with late awareness, suggesting that the processes underlying the LRP may cause our awareness of movement initiation” ( Haggard and Eimer, 1999 ). From this, one can deduce that the awareness of the intention to move one finger or the other comes after the decision was “taken by the brain”, as reflected in the LRP.

Sirigu et al. (2004) and Desmurget et al. (2009) have shown that, repeating Libet’s experiments on patients with parietal lesions, it appears that they become aware of their decision to take action only when the action itself is being carried out. In these subjects the awareness of the decision does not even come before the beginning of the movement, as it tends to coincide with the motor action. It seems that in such cases the brain alteration has reduced, if not cancelled altogether, the interval of consciousness preceding the actual implementation of the action. The authors proposed that when a movement is planned, activity in the parietal cortex, as part of a cortical sensorimotor processing loop, generates a predictive internal model of the upcoming movement. And this model might form the neural correlate of motor awareness.

Fried et al. (2011) recorded the activity of 1019 neurons as 12 subjects performed self-initiated finger movements. They found progressive neuronal recruitment, particularly in the SMA, over 1500 ms before subjects reported making the decision to move. A population of 256 SMA neurons was sufficient to predict in single trials the impending decision to move: 700 ms before the participants became aware of the decision, the accuracy of the prevision was higher than 80%. Fried et al. (2011) were also able to predict, “with a precision of a few 100 ms, the time point of that voluntary decision to move”, and they implemented a computational model thanks to which “volition emerged when a change in the internally generated firing rate of neuronal assemblies crossed a certain threshold”.

Unreliability of the Conscious Intention

A slightly different trend of research compared to Libet’s comprises studies suggesting that the conscious intention of an action is strongly influenced by events that occur after the action itself was performed. In this sense, intentions are therefore partially reconstructed according to a process of inference, based on elements that come after the action. For instance, a study by Lau et al. (2006) has produced results that empirically support this hypothesis. The authors have used transcranic magnetic stimulation (TMS) on the pre-supplementary motor (pre-SM) area, while the subjects were performing Libet’s task. The stimulation of the pre-SM through TMS happened at different time intervals, in relation to a simple voluntary movement. When the stimulation was applied 200 ms after the movement, the judgment W was moved back in time, indicating that the perception of the intention was influenced by the neural activity of the pre-SM after the motor action was made (cf. also Lau et al., 2004 ; Lau and Passingham, 2007 ).

In another experiment, Banks and Isham (2009) have set a slightly different version of Libet’s task: participants were asked to push a button whenever they wanted, and later they had to indicate the precise moment when they had the intention to do so. When they pushed the button, subject received an auditory feedback with a delay from 5 to 60 ms, so as to give them the impression that the response happened after they pushed the button. Even though the subjects weren’t aware of the delay between the action and the auditory feedback, the intention to press the button was reported as happening later in time, according to a linear function with the delay of the auditory signal feedback. The identification of the moment in which the subject had intended to press the button—measured by judgment W—was therefore largely determined by the apparent time of the subject’s response, and not the actual answer. This result indicates that the people evaluate the time when they have had the intention to take an action based on the consequences of their action and not just on the motor action itself.

Kühn and Brass (2009) conducted an experiment combining the paradigm of the stop signal ( Logan et al., 1984 ) with an intentional action paradigm. The subjects had to react in the quickest possible way by pushing a button as soon as a stimulus (e. g., a letter) was displayed at the center of a computer screen. Sometimes, just after the presentation of the stimulus, either a stop signal or a decision signal was shown: in the first case, the subjects had to try to stop responding; in the second case they could decide whether to press the button or stop responding. In the decision trials in which subjects had provided an answer, the subjects were asked if it had actually been the result of a decision, or if it had been inhibited—that is, if they had not been able to stop before the decision signal was presented.

The results have shown that in some instances, the subjects judged as intentional responses—i.e., as the result of a decision—those answers that in reality, on the basis of reaction times, were failed inhibitions. In other words, sometimes the subjects had a subjective experience of having intentionally decided to perform an action that they had actually not decided to take. These studies have empirically supported the hypothesis that the intentions to take voluntary actions are strongly influenced by events occurring after the execution of the action. In addition, they seem to confirm that the brain motor system produces a movement as the final result of its inputs and outputs; consciousness would be “informed” of the fact that a movement is going to occur and this would produce the subjective perception that the movement was decided voluntarily ( Hallett, 2007 ).

Predicting Choices

More recently, studying the activity of the frontal and parietal cortex, other neuroscientists of the group coordinated by Soon et al. (2008 , 2013) have managed to detect the “rise” of a behavioral or abstract choice/decision (to move either the right finger or the left one; to perform a mathematical operation or another with two numbers) a few seconds before the subject becomes aware of it. An unconscious brain process has already “decided” what to do when the subject still does not know what she would choose and thinks she still has the power to decide. More precisely, Soon et al. (2008) studied “free decisions” between many behavioral options using the multivariate pattern classification analysis (MVPA) which, combined with fMRI, allows one to identify specific contents of cognitive processes. “A pattern classifier, usually adopted from machine learning, can be trained on exemplars of neural patterns acquired when participants make different decisions and can learn to distinguish between these. If the activation patterns contain information about the decisions, the trained classifier can then successfully predict decision outcomes from independent data” ( Bode et al., 2014 ).

In Soon et al.’s (2008) experiment, subjects carried out a freely paced motor-decision task (choosing to press a button with either the left or the right index finger) while their brain activity was being measured using fMRI. The subjects then had to report the moment of the decision, not by using a clock as in Libet’s experiment, but by selecting a letter in a stream that was being presented during the task. Soon et al. (2008) used fMRI signals to find local neural patterns and draw from such patterns all possible information decoded second by second thanks to the statistical techniques of pattern recognition. The brain areas that were mostly involved in the performance of the actions are the primary M2 and the SMA, while two other brain regions encoded the subject’s motor decision ahead of time and with high accuracy. Indeed, the frontopolar cortex (BA10) and a portion of the cingulate cortex can be monitored to understand what kind of choice will be made by the person before they are conscious of having taken a specific decision in the task they were given. The prediction can be made, with a relevant approximation (60% mean accuracy), up to 7 s before the conscious choice is experienced by the subject, thanks to the fMRI signals detected in the BA10 (one should take into account that the subjects are asked to think hard about the choice before making it, whereas usually simple choices do not require long subjective reflection). “The temporal ordering of information suggests a tentative causal model of information flow, where the earliest unconscious precursors of the motor decision originated in frontopolar cortex, from where they influenced the buildup of decision-related information in the precuneus and later in SMA, where it remained unconscious for up to 10 s” ( Soon et al., 2008 ).

This seems to revive the old issue of God’s foreknowledge that forced theologians to wonder if man can be considered free, if someone already knows his future choices. Indeed, the authors speak of “free” decisions determined by brain activity ahead of time by placing “free” between inverted commas, as freedom is taken to be a commonsensical hypothesis. In this regard, the authors claim: “we found that the outcome of a decision can be encoded in brain activity of prefrontal and parietal cortex up to 10 s before it enters awareness. This delay presumably reflects the operation of a network of high-level control areas that begin to prepare an upcoming decision long before it enters awareness” ( Soon et al., 2008 ).

Another interesting study is that conducted by Alexander et al. (2016) : using a new experimental design, it found that the RP also occurs in the absence of movement. It suggests that “the RP measured here is unlikely to reflect preconscious motor planning or preparation of an ensuing movement, and instead may reflect decision-related or anticipatory processes that are non-motoric in nature” ( Alexander et al., 2016 ). The experimental design used a modified version of Libet’s task. Subjects had to choose between four letters whenever they wanted, by taking note of the exact moment of their choice. Later, in half the trials, the subjects had to push a button as soon as they made the decision, whereas in the other half subjects had to do nothing to mark their choice. At the end of the task, all subjects had to report when they had made their decision. In this way, by EEG, electrooculography (EOG) and electromyography (EMG), it was possible to see the RP of the decision-making both in motor and non-motor contexts.

The authors did not find any strong differences between the two RPs, thereby affirming that there is a pure cognitive contribution to RP that does not reflect processes related to movement. They thus suggest that cognitive RP might reflect action preparation, general anticipation and spontaneous neural fluctuations. Interestingly, they exclude that the RP reflects action preparation since it is a non-motor processing. And as to anticipation they cannot exclude that RP may be specifically associated with free choice. So the RP could merely reflect the average of spontaneous fluctuations (see “Other Neuroscientific Hypotheses on Free Will” Section).

Free Will as an Illusion

All these experiments seem to indicate that free will is an illusion. Yet, these relevant experiments can be interpreted in many ways. A possible view is that, in some way, determinism can be observed directly within ourselves. This interpretation might lead to the conclusion that free will is just an illusion. In fact, if one considers as a condition of free will the fact that it should be causa sui (i.e., it should be able to consciously start new causal chains), such a condition is incompatible with determinism as it is usually defined. For it, in fact, all events are linked by casual relations in the form of natural laws, which started long before we were born and which we cannot escape.

However, determinism has generally been regarded as a metaphysical claim, not refutable by empirical findings. One could properly talk of automatism in the brain, not of determinism, based on the evidence available. (In any case, endorsing indeterminism might lead to consider our behavior as the causal product of choices that every time produce different results, as if we rolled a dice. This doesn’t seem to make us any freer than if determinism were overturned; cf. Levy, 2011 ). Most importantly, another feature of freedom seems to be a pure illusion, namely the role of consciousness. The experiments considered thus far heavily question the claim that consciousness actually causes voluntary behavior. Neural activation starts the decisional process culminating in the movement, while consciousness “comes after”, when “things are done”. Therefore, consciousness cannot trigger our voluntary decisions. But the role of consciousness in voluntary choices is part of the definition of free will (but the very definition of consciousness is a matter of debate, cf. Chalmers, 1996 ).

Empirical research in psychology also shows that our mind works and makes choices without our conscious control. As proposed by psychologist Wegner (2002 , 2003 , 2004) and Aarts et al. (2004) , we are “built” to have the impression to consciously control our actions or to have the power to freely choose, even though all that is only a cognitive illusion. Many priming experiments show that people act “mechanically” (even when their behavior might appear suited to the environment and even refined). Automatic cognitive processes, of which we aren’t always aware, originate our decisions, and they were only discovered thanks to the most advanced scientific research. Ultimately, consciousness, which should exercise control and assess the reasons for a choice, is thus allegedly causally ineffective: a mere epiphenomenon, to use the terminology of the philosophy of mind. This is what has been called Zombie Challenge , “based on an amazing wealth of findings in recent cognitive science that demonstrate the surprising ways in which our everyday behavior is controlled by automatic processes that unfold in the complete absence of consciousness” ( Vierkant et al., 2013 ).

These experiments have triggered a huge debate and led scientists, philosophers and intellectuals to claim (or insist even more, if they already denied free will) that free will doesn’t exist ( Greene and Cohen, 2004 ; Cashmore, 2010 ; Harris, 2012 ). It seemed as though neuroscience had produced empirical evidence against free will, so that the century-long debate on it could be considered solved. However, Libet’s experiments have been also criticized. Much criticism was directed to the philosophical interpretation of these studies ( Mele, 2014 ) or to their theoretical assumptions ( Nachev and Hacker, 2014 ), which are important but not relevant here. Among the forms of criticism, one has to mention the theories of action that separate the deciding from the initiating ( Gollwitzer, 1999 ; Gollwitzer and Sheeran, 2006 ). In that case, free and conscious deliberating could still have a relevant casual role, long before the actual performance of the action.

Other objections, more markedly neuroscientific, were made for instance by Trevena and Miller (2010) . They argued that the RP is not an intention to move, but only indicates that an attentional process is in place in the brain, since when subjects “attended to their intention rather than their movement, there was an enhancement of activity in the pre-SMA” ( Lau et al., 2004 ). In any case, “there was no evidence of stronger electrophysiological signs before a decision to move than before a decision not to move, so these signs clearly are not specific to movement preparation”, ( Trevena and Miller, 2010 ). Others have noted that the introspective estimates of event timing are disputable or inaccurate, and measures in general are not sufficiently exact ( Dennett, 1984a , b , 2003 ).

More than Explaining Away

Other studies using multivariate pattern analysis with EEG confirmed that the subjectively free decisions might be made in the brain in the same way as evidence-based perceptual decisions ( Bode et al., 2012a , b , 2013 ). Indeed, Bode et al. (2012b) wrote,

we directly decoded choice-predictive information from neural activity before stimulus presentation on pure noise trials on which no discriminative information was present. Choice behavior on these trials was shown to be primed by the recent choice history. Modelling of sequential effects in RT and accuracy confirmed that such choice priming biased the starting point of a diffusion process toward a decision boundary, as conceptualized in evidence accumulation models of perceptual decision making ( Bode et al., 2012b ).

In other words, the authors found that internally (and maybe stochastically) generated neural activity can bias decisions that are expected to be stimulus-responsive or (possibly) reason-responsive. In this case, as in others that I will consider below, the understanding that we begin to have of the neuronal processes in play shows us that there is a complexity of factors at work. Some of these factors seem to be genuinely random, due to the pure noise produced by the default brain activity, while other factors can be traced back to the previous history of decisions taken in similar situations or related to the present one. Therefore, there is no “mysterious” start of the action as a linear process that, from the initial command, is then executed, as in Libet’s simplified model. Rather, this outcome is the result of a multiplicity of causal elements, which are homogeneous from the viewpoint of proximal mechanisms but of different relevance from the viewpoint of interpretation in terms of intentional psychology.

Another study has shown that attempts to account for (make sense of) insufficient perceptive clues use the same neural networks as those involved in “free” decision-making ( Bode et al., 2013 ). An fMRI-based pattern classifier can be trained to differentiate between different perceptual guesses and try to predict the outcome of non-perceptual decisions, like those made by the participants in the experiments considered so far. Specific activation patterns detected in the medial posterior parietal cortex have allowed the authors to make correct predictions on the participants’ free choices based on the previously decoded perceptual guesses decoded, and the other way round.

The task was the following: the participants were given a masked stimulus and had to say what category the stimulus belonged to. They had to freely choose among many categories. Thanks to the multivariate pattern analysis it was possible to identify the model of “free decisions” to make correct predictions in the context of perceptual judgments and identify the model of the “guess decisions”, to make correct predictions in the context of “free decisions”. It thus seems that a similar neural code for both types of decision is present. In those cases one could say that guessing is similar to making a free decision, since the brain, in the absence of sufficient external cues, has to decide internally. So perceptual decisions can be predicted from specific preceding neural activity when the brain doesn’t have enough internal elements to reach the threshold of perceptive decision.

Studies and commentaries have nevertheless drawn attention to possible confounds and bias in those experiments, namely they might be affected by previous choices with a form of auto-correlation in spontaneous decisions. In particular, Lages and Jaworska (2012) “trained a linear classifier to predict “spontaneous decisions” and “hidden intentions” from responses in preceding trials and achieved comparable prediction accuracies as reported for multivariate pattern classification based on voxel activities in frontopolar cortex”. Lages et al. (2013) have stressed a possible sequential information processing between trials that can introduce a confound, and recommended that “rather postulating a 50% chance level, prediction should be tested with a permutation test and/or separate multivariate classification analyses conditional on the previous response”.

The prediction of perceptual decisions from specific preceding neural activity is linked to what is defined “evidence accumulator model for free choice” ( Bode et al., 2014 ). The explanation starts with the fact that predictive activation patterns preceding decisions become increasingly similar to the patterns detected when the decision is consciously experienced by the subject. This could mean that a slow build-up of decision-related activity occurs, as it happens in accumulation of decision-related evidence to a decision threshold ( Ratcliff, 1978 ; Ratcliff and McKoon, 2008 ). Also, as already noted, when no external feedback is available, the previous choice is used as external feedback ( Akaishi et al., 2014 ). The history of previous decisions has a systematic effect on subsequent choices, related to the activity in medial posterior parietal cortex/posterior and posterior cingulate cortex ( Bode et al., 2011 , 2013 ). And the systematic effect can go in the direction of repetition or of avoidance of repetition depending on the task ( Mochizuki and Funahashi, 2014 ).

Here is an important point that deserves study from the neuroscientific point of view but also from that of a philosophical interpretation of free will. It consists in the fact that the internally generated brain activity has to do both with the stochastic noise and with the history of the subject’s choices. On the one hand, the stochastic noise comes both from the configuration that the brain has on average as a result of evolution (adaptive significance) and from individual development, resulting from random processes and environmental influences. On the other hand, the history of the choices is derived from the same process (in part stochastic) that I have just described.

In any case, if (at least some) very short-term decisions have a genesis similar to that described here, these decisions contribute to shaping the brain activity, and then, presumably, also to influencing decisions on a longer time scale that it is not yet possible to investigate experimentally. Ultimately, this could mean that there is a confluence of causal factors at the level of microdecisions. These factors add up in a way that it is hardly possible to tackle for current science. Then also the reasons motivating an action, typical of free actions, such as “I punched the stalker because it is right to punish those who behave in this way and because I wanted to set an example for all”, encoded in neural activity, can be part of the sum of neural causes.

In fact, experimental psychology has been trying to take into account long-term influences. In the so-called marshmallow experiment, researchers focused on delayed gratification ( Mischel et al., 1972 , 1989 ). A child was given a choice between one small, immediate reward and two small rewards (i. e. a larger reward) if they were able to wait some minutes while the psychologist left the room and then came back. Children who waited longer for the their rewards tended to have better life outcomes and accomplishments. Such experiments are relevant in terms of explanations and predictions, but it seems hard to trace behavioral profiles back to specific profiles of cerebral activation, once we are aware of the complexity of causal chains in the evidence accumulation model.

As Bode et al. (2014) write, in the hypothesis of an evidence accumulator that collects sensory evidence until a decision threshold is reached,

task instructions, participants’ internal motivation, and previous choices all have a strong influence on how decision tasks are performed when external information is either unavailable (as in free decisions), or unhelpful (as in perceptual guessing). In the case of free decision tasks, fluctuating intention for one or the other option may result from active competition between neural representations of both options in decision networks (or rather although not consciously monitored by the participants, the previous choice history, embodied in dynamic states of decision networks, can become the primary determinant of behavior, simply because nothing else is available ( Bode et al., 2014 ).

However, in this way things get more complicated and at a macroscopic level of behavioral observation, this blurs but doesn’t do away with the idea of free behaviors and behaviors that could be taken as unconscious decisions, of which we become aware only when the action has been performed. What remains to be solved is the problem of the distinction between external stimuli that trigger a stimulus-response circuit, and internal self-paced intentions and decisions that trigger voluntary circuit ( Haggard, 2008 ).

Other Neuroscientific Hypotheses on Free Will

Beyond determinism and consciousness.

The concept of free will relevant to our moral and legal, personal and social practices is much more complex than that captured by the experiments considered up to now. But here what matters are not so much theoretical considerations or those derived from experimental psychology (such as the role played in decisions by implementation intentions, which then re-evaluate the active role of consciousness; Gollwitzer, 1999 ), but those that originate from the neuroscientific research itself. In what might be called a new phase of empirical investigation on free will, the problem of determinism and the role of consciousness is left in the background, and the focus goes to other factors that enter the brain mechanisms of decision-making, without asking first if those processes (necessarily the most simple, at least for now) are deterministic or stochastic. On the other side, neuroscientists are trying to confine the concept of free will to operationalizable situations, so as to measure it and be able to identify, at least as a goal, its neural correlates.

There is a line of research on non-human primates, but more recently also on humans, which studies fine decision-making at the neuronal level, bringing it back to a mechanistic process that might be the neuronal interface of our common sense descriptions. This trend has been well described by Roskies (2010a , b , 2013) , who is one of the major supporters of this approach. For example, in Shadlen and Newsome’s (2001) experiments, monkeys are trained to look at stimuli consisting of points that move randomly to the right and to the left and to “indicate” the overall direction of the points. The monkeys give this indication moving their eyes (with a saccade) to the right or to the left. What emerges is that the activity of the neurons of the lateral interparietal (LIP) area increases with the information in the sensory cells of the middle temporal (MT) area and upper middle temporal (MST) area. The discharge rates rise up to reaching a given level, at which the monkey performs the saccade and the neurons stop discharging. This is the threshold for a decision to take place. The time taken to reach the threshold level depends on the perceptual characteristics of the stimulus (the strength of the movement over time) and the discharges stop after the answer was given.

The discharges also depend on whether the monkey is asked to answer when he wishes, or rather to hold back the response until the signal is given for the saccade. If the monkey is asked to wait until the signal is given to respond, LIP neurons continue to discharge even in the absence of the visual stimulus ( Gold and Shadlen, 2007 ). According to Roskies (2010b) , this is the discharge scheme of a neuron involved in the decision-making process; the levels of discharge can be maintained in the absence of the stimulus, signifying the independence of the decision from the inputs on which it operates, and the activity continues until it reaches the critical level at which the response is generated, or until the neurons that represent the elements accumulated in favor of a different choice lead to eye movement. In addition, electrical stimulation of LIP neurons can influence the monkey’s decision, indicating that LIP cells causally contribute to the process that triggers decision and action ( Hanks et al., 2006 ). It remains, however, to be established whether this role is that of deliberation that leads to a decision or that of the decision itself.

The reaction times and the accuracy in the evaluation are very similar between monkeys and humans, with the probability of choice and the response time connected in a similar way to the difficulty of discriminating the stimulus, so that it can be assumed that also in humans these neural processes are similar. A mathematical description of the dynamics of this system allows one to talk about the race towards the critical threshold ( Gold and Shadlen, 2007 ; Wong et al., 2007 ). According to this model, the neuronal populations with specific response properties represent different “hypotheses”. The discharge rates represent the strength of the evidence in favor of those hypotheses based on evidence gathered from the environment. When the evidence for and against each hypothesis is integrated, the discharge rates reach or move away from the critical level, which represents the decision point. This is the point at which the animal “made a choice” about the overall direction of movement. The first group that reaches this threshold “wins”, leading the motor response.

Schurger et al. (2012) proposed a different interpretation of the premovement buildup of neuronal activity preceding voluntary self-initiated movements in humans as well. They used “a leaky stochastic accumulator to model the neural decision of “when” to move in a task where there is no specific temporal cue, but only a general imperative to produce a movement after an unspecified delay on the order of several seconds”. According to their model, “when the imperative to produce a movement is weak, the precise moment at which the decision threshold is crossed leading to movement is largely determined by spontaneous subthreshold fluctuations in neuronal activity. Time locking to movement onset ensures that these fluctuations appear in the average as a gradual exponential-looking increase in neuronal activity” ( Schurger et al., 2012 ).

The model proposed by Schurger et al. (2012) accounts for the behavioral and EEG data recorded from human subjects performing the task and also makes a specific prediction that was confirmed in a second electroencephalography experiment: fast responses to temporally unpredictable interruptions should be preceded by a slow negative-going voltage deflection beginning well before the interruption itself, even when the subject was not preparing to move at that particular moment. The task was to repeatedly push a button, sometimes at will, sometimes in response to a sound produced by the experimenters according to a causal sequence. The speed of response (pressing the button) when the sound is produced is related to the proximity to the peak of the background brain activity, which appears to be random, an ebb and flow that has its highest point in the threshold at which it produces the decision to push the button.

According to this explanation, “the RP does not reflect processing within a specific action domain. Our finding that movement does not significantly modulate RP amplitude supports this aspect of their claim by extending the RP to the domain of covert decisions” ( Alexander et al., 2016 ). Another consequence is the fact that the neural decision to move at a specific time happens much later compared to Libet’s hypothesis, and the RP is only a by-product of a drift diffusion process. But the RP would still be predictive in that it precedes action and conscious awareness of both motor and cognitive action. However, the RP is predictive with regards the whether and the when, if a known task is performed, but not with regards to the what of the action ( Brass and Haggard, 2008 ).

Jo et al. (2013) seems to go in the same direction with their work: they considered both the positive and the negative potential shifts in a “self-initiated movement condition” as well as in a no-movement condition. The comparison of the potential shifts in different conditions showed that the onset of the RP appeared to be unchanged. “This reveals that the apparently negative RP emerges through an unequal ratio of negative and positive potential shifts. These results suggest that ongoing negative shifts of the SCPs facilitate self-initiated movement but are not related to processes underlying preparation or decision to act” ( Jo et al., 2013 ).

Murakami et al. (2014) confirmed those findings. They used rats, who had to perform a specific task: wait for a tone (which was purposely delayed) and decide when to stop waiting for it. The rats’ neuronal activity of the secondary M2 was recorded and resulted consistent with the model of integration-to-bound decision. “A first population of M2 neurons ramped to a constant threshold at rates proportional to waiting time, strongly resembling integrator output. A second population, which they propose provide input to the integrator, fired in sequences and showed trial-to-trial rate fluctuations correlated with waiting times” ( Murakami et al., 2014 ). Also, an integration model based on the recorded neuronal activity in the considered brain areas has allowed the researchers to quantitatively foresee the inter-neuronal correlations manifested during the task performance. “Together, these results reinforce the generality of the integration-to-bound model of decision-making. These models identify the initial intention to act as the moment of threshold crossing while explaining how antecedent subthreshold neural activity can influence an action without implying a decision” ( Murakami et al., 2014 ).

Schurger et al. (2016) stress that the main new finding about the brain activity preceding SVM “is that the apparent build-up of this activity, up until about 200 ms pre-movement, may reflect the ebb and flow of background neuronal noise, rather than the outcome of a specific neural event corresponding to a “decision” to initiate movement”. The model used is the bounded-integration process, “a computational model of decision making wherein sensory evidence and internal noise (both in the form of neural activity) are integrated over time by one or more decision neurons until a fixed threshold-level firing rate us reached, at which the animal issues a motor response. In the case of spontaneous self-initiated movement there is no sensory evidence, so the process is dominated by internal noise” ( Schurger et al., 2016 ). The stochastic decision model (SDM) used by Schurger et al. (2012) allowed them to claim that bounded integration seems to explain stimulus-response decision as relying on the same neural decision mechanism used for perceptual decisions and internal self-paced intention and decision as “dominated by ongoing stochastic fluctuations in neural activity that influence the precise moment at which the decision threshold is reached” ( Schurger et al., 2016 ). And this mechanism seems to be shared with all animals including crayfish ( Kagaya and Takahata, 2010 ).

The philosophical implications could be that “when one forms an intention to act, one is significantly disposed to act but not yet fully committed. The commitment comes when one finally decides to act. The SDM reveals a remarkably similar picture on the neuronal level, with the decision to act being a threshold crossing neural event that is preceded by a neural tendency toward this event” ( Schurger et al., 2016 ).

The Veto Power

Another recent study has brought back to the center of neuroscientific research the space of autonomy that the subject seems to have compared to the idea of free will as an illusion supported by the experiments based on the alleged unconscious onset of the action. Schultze-Kraft et al. (2016) showed that people are able to cancel movements after elicitation of RP if stop signals occur earlier than 200 ms before movement onset. In the real-time experiment, “subjects played a game where they tried to press a button to earn points in a challenge with a brain–computer interface (BCI) that had been trained to detect their RPs in real time and to emit stop signals” ( Schultze-Kraft et al., 2016 ).

The subjects had to press with their foot a button on the floor after a green light flashed: they could so whenever they wanted after about 2 s. Participants earned points if they pressed the button before the red light to come back (the stop signal). The experiment was composed of three phases. In the first phase, the stop signals were lit at random and the movements of the subjects were not predicted. In the second phase, the authors used data taken from the EEG on the participants in the first phase. In this way a classifier was trained to predict (with imperfect accuracy) the movements (the When and the Whether, not the What). In this phase, the BCI could foresee the fact that the subject would press the button thanks to the detection of the RP and therefore turned on the red light to earn points against the subject if it could not stop the movement. In the third phase, the subjects were informed that the BCI could “see their preparation of the movement” and they had to try to beat the computer by moving in an unforeseeable way.

In all phases of the experiment, there was no difference between RPs. While in the first phase, in 66.5% of the cases, subjects were winning by pressing the button with the green light on, in stages two and three trials in which subjects were able to beat the computer, by not pushing the button with the red light on, decreased to 31%, and warning participants of the prediction of the BCI would not help them do any better. The authors could thus claim that “despite the stereotypical shape of the RP and its early onset at around 1000 ms before EMG activity, several aspects of our data suggest that subjects were able to cancel an upcoming movement until a point of no return was reached around 200 ms before movement onset. If the stop signal occurs later than 200 ms before EMG onset, the subject cannot avoid moving” ( Schultze-Kraft et al., 2016 ). The explanation of the minimum threshold of 200 ms could reflect the time necessary for the stop signal to light up, the subject to perceive it and cancel the movement that was already being prepared.

As to which cortical areas are involved in vetoing an already initiated movement, some studies have tried to identify them. Brass and Haggard (2007) examined the voluntary inhibition using an experimental paradigm that was based on the Libet task. The subjects were asked to press a button while watching a cursor moving along the face of a clock. Every time, after pressing the button, the subjects had to signal the precise moment when they thought they decided to press the button. In addition, the instructions specified that the participants had to inhibit the execution of the response in some tests of their choice. Comparing this voluntary inhibition condition with the condition in which the action had not been inhibited, the authors observed an activation of the dorsal fronto-medial cortex (DFM). This area is different from the brain regions involved in the stop signal tasks, in which the inhibition is controlled by external signals. Furthermore, the DFM cortex is also distinct from the brain regions controlling the activity linked to the when and what components of voluntary action. Brass and Haggard (2007) have interpreted this finding as evidence that there is a mechanism of voluntary inhibition that can be dissociated, in neuroanatomo-functional terms, from an “environmental” inhibiting mechanism, which involves the lateral prefrontal cortex.

This finding was replicated in a subsequent study of Kühn et al. (2009) , in which the subjects had to avoid dropping a ball sliding down a ramp, by pressing a button before the ball came down and broke. In some tests of their choice, they could choose to voluntarily inhibit the response. The comparison of the condition of voluntary inhibition with the condition of voluntary action still showed activation of the DFM cortex, supporting the idea that this area is involved in the inhibition of voluntary action ( Schel et al., 2014 ).

Finally, Schultze-Kraft et al. (2016) declared to be agnostic about the interpretation of their data in regards of RP. As the RP is predictive of the subsequent movement, it could be read as “the leaky integration of spontaneous fluctuations in autocorrelated neural signals”. Theoretically, the question remains about the departure of the intention to block the action while the movement is being prepared, along with the possible coexistence of two intentions suggested by the commands of the experimenters. The participants in the experiment, in fact, want to win against the computer, therefore they want to push the button, and also have the intention, partly contrasting, not to push the button when the computer turns the red light on.

A More Realistic Model

This novel perspective offered by the line of research by Schurger et al. (2012) here described works on very simple decision-making processes and could be exposed to the same criticism in this regard have been made to Libet’s research line. But Roskies (2010b) has suggested some tracks along which to develop research on more complex decision-making processes, close to those relevant to social life. First, one must introduce the value of the decision, seen as a subjective or moral feature that drives action. By manipulating the expected rewards for correct action or for a particular type of decisions, or by manipulating the probabilities of the outcomes, both the decision and the activity levels of LIP neurons are altered ( Platt and Glimcher, 1999 ; Glimcher, 2002 ; Dorris and Glimcher, 2004 ; Sugrue et al., 2004 ). In this way it is possible to change the monkey’s choice about the objective of the saccade by offering her favorite reward. Although it is not known how the figures are represented, it seems that the Lip neurons can integrate the information on the value or on the reward in the decision-making process, and that information has a causal role.

As for the reasons, and the responsiveness to them, Roskies (2010b) suggests that also the reasons, albeit discursive and propositional, may be encoded as information at the neuronal level. Simplifying, in her view one might think that in a situation where, say, there is little food and many people, different populations of neurons represent the content “I am hungry”, while others represent “others need this more than I do”, others “the weak come first” and so forth, weighing reasons in terms of activation and modulation of the response of the populations of neurons delegated to the choice and the final decision. However, such a model ( Dorris and Glimcher, 2004 ) should be considered as purely hypothetical because first we do not know what are the specific populations of neurons, we don’t yet have the instruments to identify them, and we do not know their interactions (also considering the recent failures of naturalized semantics).

Secondly—and perhaps most importantly—it is unclear how what we externally call “reasons” could be activated and weighed by the decision-maker understood as a unitary subject or self, according to the description for which we truly act based on reasons. In this case, I believe one cannot seek a simple neural interface for commonsensical concepts and notions. In fact, the idea of a deep and unitary self—the idea of a conscious subject controlling her behavior instant by instant—has been strongly challenged by evidence coming from empirical psychology and cognitive neuroscience ( Dennett, 1991 ; Metzinger, 2004 , 2009 ). Therefore, one should avoid the temptation to reproduce such a description in neural levels. But if we trace back the reasons to populations of neurons in a mechanistic model—if we trace them back to thresholds—it is not easy to figure out who makes the decisions and why. If it is true that some people seem to be more sensitive to specific reasons, other than those to which other people are sensitive, and if people can change over time the reasons by which they are usually motivated, and in certain situations the same people may not to respond to the reasons to which they are usually sensitive, one has to wonder if what prevails are processes that we would call random or that, in any case, are beyond our control.

Here the role of consciousness seems again to be relevant. If experiments à la Libet seemed to have ruled it out from a causal standpoint, the experiment by Schultze-Kraft et al. (2016) on movement vetoing seems to reassess its role in blocking the preparation process triggered in the brain. In this sense, this seems to be a more realistic line of neuroscientific investigation on free will, one that contemplates, even in broad terms, stochastic brain processes, for the most part triggered by environmental stimuli, which often are not aware of (the same as our train of thoughts arising spontaneously without us being able to orientate it from the beginning), but also by spontaneous activity of the brain ( Changeux, 2004 ; Brembs, 2011 ) that creates models of reality. “Learning mechanisms evolved to permit flexible behavior as a modification of reflexive behavioral strategies. In order to do so, not one, but multiple representations and action patterns should be generated by the brain” ( De Ridder et al., 2013 ). And this repertoire is not infinite. Indeed, “our evolutionary-evolved brain potential to generate multiple action plans is constrained by what is stored in memory and by what is present in the environment” ( De Ridder et al., 2013 ). Schurger and Uithol (2015) also argue that the “actions emerge from a causal web in the brain” and that the “proprioceptive feedback might play a counterintuitive role in the decision process”. They, thus recommend the use of dynamical systems approach for the study of the origins of voluntary action.

On these spontaneous processes we can exercise control, which can be considered automatic and unconscious when evaluated with the classical theoretical criteria of conscious control. First, there is an innate behavioral repertoire of provisions linked to survival in the environments within which we evolved. Secondly, there is a repertoire of behavioral provisions that is stratified in terms of conscious repetitions due to environmental stimuli or to internal choices (with all the limitations that this expression has in reference to the brain mechanisms analyzed so far) and then becomes automatic. The control can, however, also be explicit, with obvious limitations and cases of complete control failure. Based on this complex self-construction (which has a neural correlates), we are creatures with a higher or lower degree of free will. This free will may then be better understood and circumscribed, so as to be more objectively operationalized and also measured.

Operazionalizing, Measuring and Verifying: From the Action to the Brain

My view is that a richer conceptualization of free will—one that is able to overcome the stall of the metaphysical debate as well as the current difficulties of neuroscience ( Nachev and Hacker, 2014 ) and empirical psychology ( Nahmias, 2014 )–has to be linked to the idea of “capacity”. In fact, as claimed by Mecacci and Haselager (2015) , the kind of free will investigated by neuroscientific experiments, which is self-generated and defined according to the absence of cues, “does little justice to the common sense practice of holding people responsible for their freely willed actions that consists in asking explanations and justifications from the actor” ( Mecacci and Haselager, 2015 ).

Another important point is that there are differences in time scales between laboratory tasks (the milliseconds to seconds time range) and real life or, better, life as we measure it temporally (seconds, minutes, hours, weeks, years) regarding decisions that really concern us. Even if the underlying mechanism might be the same, the experiments described so far cannot investigate whether decisions with a longer maturation process are free and to what extent they are such. It might be possible to distinguish between proximal and distal mechanisms, but this doesn’t seem feasible lacking the tools to address decisions involving longer time scales. For this reason it might be useful to introduce other and different ways to conceptualize and operationalize (supposedly) free actions.

“By capacity, in the context of free will, we mean the availability of a repertoire of general skills that can be manifested and used without the moment by moment conscious control that is required by the second condition of free will we have previously seen” ( Lavazza and Inglese, 2015 ). The concept of capacity is related to that of internal control, understood as the agent’s “ownership” of the mechanism that triggers the relevant behavior and the reasons-responsiveness of that mechanism ( Fischer and Ravizza, 2000 ). And reasons-responsiveness must involve a coherent pattern of reasons-recognition. “More specifically, it must involve a pattern of actual and hypothetical recognition of reasons that is understandable by some appropriate external observer. And the pattern must be at least minimally grounded in reality” ( Lavazza and Inglese, 2015 ). The concept of capacity used in this sense, and combined with the idea of reasons-responsiveness, also avoids the objection of determinism that has always weighed on the debate on free will. From a philosophical point of view, the approach related to capacity may fall indeed in the strand of so-called compatibilism, which defends the fact that human freedom can exist even if determinism is true of the physical world.

Cognitive abilities might be firstly operationalized as a set of neuropsychological tests, which can be used to operationalize and measure specific executive functions, as they are strongly linked to the concept of control. Executive functions, also known as control functions, are essential to organize and plan everyday behavior—which is not the instant behavior found in Libet’s experiments. Those skills are necessary to perform the greater part of our goal-oriented actions. They allow us to modulate our behavior, control its development and change it according to the environmental stimuli (the environment being both physical and social). Also, executive functions allow us to change our behavior based on it effects, with a sophisticated feedback mechanism; finally, they are also necessary for tasks of abstraction, inventiveness and judgment. Those who, for whatever reason, have a deficit in their executive functions cannot respond to their social environment appropriately, and struggle to plan their behavior or to choose between alternatives based on their judgment or interest. Sufferers of these deficits in executive functions often fail to control their instinctive responses and to modify their regular courses of action, or are unable to concentrate or persist in the pursuit of a goal ( Barkley, 2012 ; Goldstein and Naglieri, 2014 ).

In general terms, the executive functions refer to the set of mental processes necessary for the development of cognitive-behavioral patterns adaptive in response to new and demanding environmental conditions. The domain of executive functions includes ( Lavazza and Inglese, 2015 ):

• the ability of planning and evaluation of effective strategies in relation to a specific purpose related to the skills of problem-solving and cognitive flexibility.

• inhibitory control and decision-making processes that support the selection of functional response and the modification of the response (behavior) in relation to changing environmental contingencies.

• attentional control referred to the ability to inhibit interfering stimuli and to activate the relevant information.

• working memory referring to the cognitive mechanisms that can maintain online and manipulate information necessary to perform complex cognitive tasks.

• (and it can be added with regards to free will) creativity and the ability to cope with environmental changes through novel solutions.

Those of empirical psychology are higher order concepts, which act as a bridge between free will, which is something that is not in the brain but can be observed in behavior (along with its causes), and the underlying brain processes. It has been convincingly suggested that in the construction of a hierarchy of mechanisms and explanations ( Craver, 2007 ), also to guide the exploration, one must go from inside to outside and from outside to inside. One goes from measurable skills to their brain basis, and from the tentative index of free will to the underlying (real) mechanisms.

Based on the evidence presented, I believe that a viable proposal is to construct an index related to compatible tests whose relevance can be uniformly ascertained. It would be a kind of IQ-like profile that would allow for the operationalization and quantification of a person’s cognitive skills. All the tests used (for example, Stroop Test, Wisconsin Card Sorting Test, Weigl’s Color-Form Sorting Test, Go-No Go Test) should be related to the subject’s age and education and then transformed in new standardized scores (Equivalent Scores, ES) on an ordinal scale, e. g. ranging from 0 to 4, with 0 representing scores below cut-off point and 4 representing scores equal to or better than average. Specific standardized scores exist in many countries or linguistic areas. The subjects would get for each test a raw score (or RS), given by the sum of the scores obtained in each item that makes up the test, which would then be standardized.

A synthetic index such as the one here proposed measures a certain range of cognitive and behavioral control skills that configure a certain kind of free will at the psychological-functional level. These are potential capacities measured with standardized instruments in laboratory situations, which do not consider any other factors that may restrict the freedom of a subject in specific situations, such as those that are relevant in moral scenarios and legal contexts. The same goes for moral judgment. However, an index such as the one I’m proposing here could be the first step, albeit certainly imperfect, towards more objective measures to discriminate between people who have more or less “free will” or, in other words, are more or less capable of self-control and rational choice (i.e., a reasons-responsive choice).

This hypothesis would be in line with the proposals of operationalizing free will advanced so far. According to Vohs (2010) , freedom might be conceived of as the sum of executive functions and goal-directed, future-oriented behaviors, which include rational choice, planning, intelligent thought, and self-control. Free will can be then constituted by a limited stock of energy, devoted to guiding executive functioning processes. The free will index I am proposing is also consistent with Baumeister’s contribution:

Psychologists should focus on what we do best: collecting evidence about measurable variance in behaviors and inner processes and identifying consistent patterns in them. With free will, it seems most productive for psychologists to start with the well-documented observation that some acts are freer than others. As already noted, dissonance, reactance, coping with stress, and other behaviors have been shown in the laboratory to depend on variations in freedom and choice. Hence, it is only necessary to assume that there are genuine phenomena behind those subjective and objective differences in freedom. In a nutshell, we should explain what happens differently between free and unfree actions ( Baumeister, 2008 ).

Empirical research on how human beings work has recently focused on self-control as a feature of free will. Self-control can be defined as the exertion of willpower on behavior. Self-control is thus generally regarded as the capacity to override inappropriate impulses and automatic or habitual responses and to suppress or delay immediate gratification so as to reach a long-term goal ( Gailliot and Baumeister, 2007 ). “Being in control” includes the capacity to maintain goals, to balance long- and short-term values, to consider and evaluate the consequences of a planned action, and to resist being “carried away by emotion” ( Churchland, 2006 ). Self-control can also be regarded as the ability of higher-order functions to modulate the activity of lower-level functions, where higher-order functions manifest themselves externally in complex behavior, adjusted according to the environmental needs, while lower-level functions are manifested in simple and stereotyped behaviors, not adjusted according to the demands of the environment ( Roskies, 2010a ). Everyone exhibits a different degree of self-control compared to other individuals, and for each person the degree of self-control varies over time ( Baumeister et al., 2006 ; Casey et al., 2011 ; Dang et al., 2015 ). The variability of self-control that is manifested in behavior and can be measured with the test has its base in neuronal functioning, which in turn depends on education and habits, external circumstances and the internal neuronal noise.

However, two executive functions turn out to be central:

(i) the ability to predict the future outcomes of a given action; and (ii) the ability to suppress inappropriate, i.e., not sufficiently valuable, actions. Importantly, these two executive functions operate not only during the genesis of an action, but also during the planning of an already selected action. In fact, during the temporal gap between the time when an action has been chosen and the moment when the motor output is going to be generated, the context might have changed, altering the computed value of the action and thus requiring a radical change of the planned motor strategy ( Mirabella, 2014 ).

It seems that the peculiarity of our freedom at the cognitive level is the ability to modulate or block courses of action that environmental stimuli automatically or unconsciously arouse in us—a reproposal in different form of Libet’s free won’t and Schultze-Kraft’s vetoing . These psychological-functional indicators must then lead to their cerebral bases. For instance, one can consider a situation in which one’s needs are satisfied (or not) and the consequent motivation to act based on the evaluation process of the need satisfaction.

This is an essential process and one that is continuously performed by our motor system. In fact, in most places where we live, if not all, we are surrounded by tools whose sight automatically activates motor schemas that would normally be employed to interact with those objects. These actions are prompted by the features of the objects, the so-called affordances ( Gibson, 1979 ). It has been shown that even the simple observation of pictures depicting affordable objects (such as graspable objects) activates a sub-region of the medial frontal cortex, the SMA, even when there is no requirement to actually act on those stimuli ( Grèzes and Decety, 2002 ). These stimulus-driven activations are rapid, involuntary, and unconscious ( Mirabella, 2014 ).

Environmental stimuli, in this case, can induce a subject to make specific choices through a priming process that exploits our action tendencies. Typically, individuals are able to control their behavior, but in some cases they fail to do so; for example if suffering from microlesions of the SMA, people have a tendency to invariably implement a certain type of action, even if the environment, both physical and social, does not require it ( Sumner et al., 2007 ). In fact, “the suppression of a triggered action might be seen not as an active process, but rather as an automatic consequence of the evaluative procedure” ( Mirabella, 2014 ). One could then say that those who have the ability to better monitor, control and direct their own behavior are “freerer” than those who do not have this capability. Individuals affected by disorders of the executive functions are not able to grasp and process environmental stimuli to direct their behavior. For example, these people may not be able to stop the utilization behavior, an automatic mechanism that tends to make us interact with all the objects that are in our perceptual sphere.

Churchland (2006) and Suhler and Churchland (2009) proposed a hypothesis concerning the neural basis for control, which can bridge the gap between higher-order concepts and brain mechanisms. As she wrote,

Perhaps we can identify various parameters of the normal profile of being in control, which would include specific connectivity patterns between amygdala, orbitofrontal cortex, and insula, between anterior cingulate gyrus and prefrontal cortex, and so forth. Other parameters would identify, for each of the six non specific systems [identified via the neurotransmitter secreted at the axon terminals: serotonin, dopamine, norepinephrine, epinephrine, histamine and acetylcholine], the normal distribution of axon terminals and the normal patterns of neurotransmitter release, uptake, and co-localization with other neurotransmitters such as glutamate. Levels of various hormones would specify another set of parameters. Yet other parameters contrast the immature with adult pattern of synaptic density and axon myelinization. At the current stage of neuroscience, we can identify the normal range for these parameters only roughly, not precisely ( Churchland, 2006 ).

This hypothesis would allow for specific brain correlates of a free will index based on the executive functions-guided self-control and even, hypothetically, a direct brain measure of being in control For example, a recent study ( Bartelle et al., 2016 ) highlights the possibility of having MRI imaging of dopamine release thanks to a engineered protein that binds to the neurotransmitter and works as a MRI-visible probe. As the authors put it, “one could imagine a future in which molecular fMRI is used to determine brain-wide neurochemicals maps corresponding to a universe of stimuli and behavioral programs”. Even though one should always consider that there isn’t perfect correspondence between higher-order concepts and putative neural correlates.

In particular, one must consider that what matters in interpersonal relations and in law, to give two examples of practical relevance of free will, is freedom as actually performed: that is, freedom as it can be observed and with some approximation, also measured through a series of psychological tests. This does not mean that the same level of freedom manifested in behavior matches the same level of activation of the related brain areas. However, one can investigate the brain causes of “freedom deficit” compared with the average shown by relevant samples of the population, and so come to a progressive refinement of the research on the neural bases of free will.

Another example is given by the investigation of the role of the cholinergic interneurons in behavioral flexibility ( Aoki et al., 2015 ). This class of neurons seem to be connected to survive in an ever-changing world, which requires behaving flexibly. Flexibility can be assessed (and measured) at a behavioral level, but cerebral mechanisms remain largely unknown. Using conventional tests on behavioral flexibility, which require animals to shift their attention from one stimulus property (e.g., color) to another (e.g., shape), researchers probed the effects of an immunotoxin-induced lesion of cholinergic interneurons in the striatum.

A selective cholinergic ablation was made by means of injections of immunotoxin, which targeted neurons containing choline acetyltransferase in the dorsomedial or ventral striatum. A control group was instead injected with saline. “When encountering a change of behavioral rules after the set-shift, either lesion made animals stick to a previously correct but now invalid response strategy. They also showed less exploratory behavior toward finding a new rule. Most interestingly, ablation of cholinergic neurons in the dorsomedial striatum impaired a shift of set when it required attention to a previously irrelevant cue. On the other hand, ventral cholinergic lesions had an effect on a shift in which a novel stimulus was introduced as a new directional cue” ( Aoki et al., 2015 ). Animals thus can be taken to be “less free” when striatal cholinergic interneurons don’t work properly.

This last example serves to indicate how to bridge the gap between overt behaviors (to which we tend to attribute the property of freedom) with neuronal mechanisms that are clearly identifiable and even manipulable. In fact, it is not so important to look at the conscious aspect of a single proximal mechanism, but rather to consider the manifest behavioral effect that the considered mechanism helps to produce. This way there would be a paradigm shift with respect to the neuroscience research on free will, which seems to have long been too closely linked to the falsification of the theoretical assumption that an action is free only if it has a beginning that is fully controlled by a conscious process. The proposal, I am making here has only the ambition to be a potentially helpful contribution to theoretical debate and empirical research, although its limits are very clear. First, it focuses on a specific part of what is intuitively called “free will”, relating it to the idea of “capacity”. Second, it proposes to measure free will at a psychological level by means of a unitary index that inevitably misses many nuances of the notion and the relative capacity. Furthermore, the search for the neural correlated of such capacities implies not only the identification of causal mechanisms, but also the consideration of many cerebral areas. All of this makes things harder compared to approaches à la Libet. Nevertheless, there is manifest advantage: there is a greater degree of realism and adherence to the actual behavioral manifestation of what we call “free will”.

Free will is an elusive but crucial concept. For many years we have known that the functioning of our brain has to do not only with the belief that we have free will but also with the existence of free will itself. Evidence of the unconscious start of movement, highlighted by the RP signal, has led to believe that we had reached an experimental proof of the non-existence of free will—which many already claimed at a theoretical level based on the argument of the incompatibility between determinism and freedom. Along with other evidence provided by experimental psychology, the branch of studies inaugurated by Libet has contributed to seeing free will as an illusion: this view seemed to be reliably supported by science, and in particular by neuroscience. Recent studies, however, seem to question this paradigm, which sees the initiation and conscious control of the action as the first requirement of free will, allegedly proving that there are no such things.

The stochastic models and the models of evidence accumulation consider decision as the crossing of a threshold of activity in specific brain regions. They do not restore the idea of conscious control but turn away from the previous paradigm. These studies cannot yet fully explain how the intention to perform an action arises in the brain, but they better account for the complexity of the process. In particular, they recognize the role of the spontaneous activity of the brain, of external cues and other factors—including those that might be called “will” and “reasons” (which, however, do not currently have precisely identified neural correlates)—in reaching the critical threshold. Studies that show how we can consciously block movements whose preparation has already begun unconsciously, then, indicate how the subject is able to exercise a form of control, whose genesis however is still unclear.

One could state that free “decision-making draws upon a rich history of accumulated information, manifested in preferences, attitudes and motivations, and is related to the current internal and external environment in which we act. Complete absence of context is impossible” ( Bode et al., 2014 ). In this framework, I have here proposed to integrate neuroscientific research on free will by connecting higher-level concepts with their neural correlates through a psychological operationalization in terms of skills and cognitive functions that do not necessarily imply a continuous conscious control over the decision-making and action process. This may also allow one to create a quantitative index, albeit still quite rudimentary, of the degree of freedom of each subject. This freedom would be specifically defined and therefore may not perfectly coincide with the intuitive concept of free will. Starting from these functional indicators, which psychology has well clarified, one could then move on to investigate the precise neural correlates for a different and (possibly) more fundamental level of explanation in terms of brain processes that enable the executive functions.

According to Craver (2007) , a mechanistic explanation is able to lead to an inter-field integration. There are two relevant aspects to this approach. The functional knowledge that can be drawn from psychological research is a tool to identify neural mechanisms; the knowledge of the brain structure can guide the construction of far more sophisticated psychological models ( Bechtel and Mundale, 1999 ). The index of free will that I am proposing ( Lavazza and Inglese, 2015 )—despite surely needing further refinement—might be useful to explore the brain mechanisms that underlie what appears in behavior as “free will”, which no longer seems to be an illusion, not even for neuroscientific research.

Author Contributions

AL confirms being the sole contributor of this work and approved it for publication.

Conflict of Interest Statement

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Aarts, H., Custers, R., and Wegner, D. (2004). On the inference of personal authorship: enhancing experienced agency by priming effect information. Conscious. Cogn. 14, 439–458. doi: 10.1016/j.concog.2004.11.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Akaishi, R., Umeda, K., Nagase, A., and Sakai, K. (2014). Autonomous mechanism of internal choice estimate underlies decision inertia. Neuron 81, 195–206. doi: 10.1016/j.neuron.2013.10.018

Alexander, P., Schlegel, A., Sinnott-Armstrong, W., Roskies, A. L., Wheatley, T., and Tse, P. U. (2016). Readiness potentials driven by non-motoric processes. Conscious. Cogn. 39, 38–47. doi: 10.1016/j.concog.2015.11.011

Aoki, S., Liu, A. W., Zucca, A., Zucca, S., and Wickens, J. R. (2015). Role of striatal cholinergic interneurons in set-shifting in the rat. J. Neurosci. 35, 9424–9431. doi: 10.1523/JNEUROSCI.0490-15.2015

Banks, W. P., and Isham, E. A. (2009). We infer rather than perceive the moment we decided to act. Psychol. Sci. 20, 17–21. doi: 10.1111/j.1467-9280.2008.02254.x

Barkley, R. A. (2012). Executive Functions: What They Are, How They Work and Why They Evolved. New York, NY: The Guilford Press.

Google Scholar

Bartelle, B. B., Barandov, A., and Jasanoff, A. (2016). Molecular fMRI. J. Neurosci. 36, 4139–4148. doi: 10.1523/JNEUROSCI.4050-15.2016

Baumeister, R. F. (2008). Free will in scientific psychology. Perspect. Psychol. Sci. 3, 14–19. doi: 10.1111/j.1745-6916.2008.00057.x

CrossRef Full Text | Google Scholar

Baumeister, R. F., Gailliot, M., DeWall, C. N., and Oaten, M. (2006). Self-regulation and personality: how interventions increase regulatory success and how depletion moderates the effects of traits on behavior. J. Pers. 74, 1773–1802. doi: 10.1111/j.1467-6494.2006.00428.x

Bechtel, W., and Mundale, J. (1999). Multiple realizability revisited: linking cognitive and neural states. Philos. Sci. 66, 175–207. doi: 10.1086/392683

Berofsky, B. (2011). Nature’s Challenge to Free Will. New York, NY: Oxford University Press.

Bode, S., Bogler, C., and Haynes, J. D. (2013). Similar neural mechanisms for guesses and free decisions. Neuroimage 65, 456–465. doi: 10.1016/j.neuroimage.2012.09.064

Bode, S., Bogler, C., Soon, C. S., and Haynes, J. D. (2012a). The neural encoding of guesses in the human brain. Neuroimage 59, 1924–1931. doi: 10.1016/j.neuroimage.2011.08.106

Bode, S., Sewell, D., Lilburn, S., Forte, J., Smith, P. L., and Stahl, J. (2012b). Predicting perceptual decision biases from early brain activity. J. Neurosci. 32, 12488–12498. doi: 10.1523/JNEUROSCI.1708-12.2012

Bode, S., He, A. H., Soon, C. S., Trampel, R., Turner, R., and Haynes, J. D. (2011). Tracking the unconscious generation of free decisions using ultra-high field fMRI. PLoS One 6:e21612. doi: 10.1371/journal.pone.0021612

Bode, S., Murawski, C., Soon, C. S., Bode, P., Stahl, J., and Smith, P. L. (2014). Demystifying “free will”: the role of contextual information and evidence accumulation for predictive brain activity. Neurosci. Biobehav. Rev. 47, 636–645. doi: 10.1016/j.neubiorev.2014.10.017

Brass, M., and Haggard, P. (2007). To do or not to do: the neural signature of self-control. J. Neurosci. 27, 9141–9145. doi: 10.1523/JNEUROSCI.0924-07.2007

Brass, M., and Haggard, P. (2008). The what, when, whether model of intentional action. Neuroscientist 14, 319–325. doi: 10.1177/1073858408317417

Brass, M., Lynn, M. T., Demanet, J., and Rigoni, D. (2013). Imaging volition: what the braincan tell us about the will. Exp. Brain Res. 299, 301–312. doi: 10.1007/s00221-013-3472-x

Brembs, B. (2011). Towards a scientific concept of free will as a biological trait: spontaneous actions and decision-making in invertebrates. Proc. Biol. Sci. 278, 930–939. doi: 10.1098/rspb.2010.2325

Casey, B. J., Somerville, L. H., Gotlib, I. H., Ayduk, O., Franklin, N. T., Askren, M. K., et al. (2011). Behavioral and neural correlates of delay of gratification 40 years later. Proc. Natl. Acad. Sci. U S A 108, 14998–15003. doi: 10.1073/pnas.1108561108

Chalmers, D. J. (1996). The Conscious Mind: In Search of a Fundamental Theory. New York, NY: Oxford University Press.

Changeux, J. P. (2004). The Physiology of Truth: Neuroscience and Human Knowledge. Cambridge, MA: Harvard University Press.

Cashmore, A. R. (2010). The lucretian swerve: the biological basis of human behavior and the criminal justice system. Proc. Natl. Acad. Sci. U S A 107, 4499–4504. doi: 10.1073/pnas.0915161107

Churchland, P. S. (2006). “Moral decision-making and the brain,” in Neuroethics. Defining the Issues in Theory, Practice, and Policy , ed. J. Illes (New York, NY: Oxford University Press), 3–16.

Craver, C. F. (2007). Explaining the Brain. Mechanisms and the Mosaic Unity of Neuroscience. New York, NY: Oxford University Press.

Crick, F. (1994). The Astonishing Hypothesis: The Science Search for the Soul. New York, NY: Scribner’s.

Dang, J., Xiao, S., Shi, Y., and Mao, L. (2015). Action orientation overcomes the ego depletion effect. Scand. J. Psychol. 56, 223–227. doi: 10.1111/sjop.12184

Dennett, D. C. (1984a). Elbow Room: The Varieties of Free Will Worth Wanting. Cambridge, MA: The Mit Press.

Dennett, D. C. (1984b). I could not have done otherwise—so what? J. Philos. 81, 553–565. doi: 10.5840/jphil1984811022

Dennett, D. C. (1991). Consciousness Explained. Boston, MA: Little, Brown and Co.

Dennett, D. C. (2003). Freedom Evolves. New York, NY: Viking.

De Ridder, D., Verplaetse, J., and Vanneste, S. (2013). The predictive brain and the “free will” illusion. Front. Psychol. 4:131. doi: 10.3389/fpsyg.2013.00131

Desmurget, M., Reilly, K. T., Richard, N., Szathmari, A., Mottolese, C., and Sirigu, A. (2009). Movement intention after parietal cortex stimulation in humans. Science 324, 811–813. doi: 10.1126/science.1169896

Dorris, M. C., and Glimcher, P. W. (2004). Activity in posterior parietal cortex is correlated with the relative subjective desirability of action. Neuron 44, 365–378. doi: 10.1016/j.neuron.2004.09.009

Doyle, B. (2011). Free Will: The Scandal in Philosophy. Vambridge, MA: I-Phi Press.

Filevich, E., Kühn, S., and Haggard, P. (2012). Intentional inhibition in human action: thepower of ‘no’. Neurosci. Biobehav. Rev. 36, 1107–1118. doi: 10.1016/j.neubiorev.2012.01.006

Filevich, E., Kühn, S., and Haggard, P. (2013). There is no free won’t: antecedent brainactivity predicts decisions to inhibit. PLoS One 8:e53053. doi: 10.1371/journal.pone.0053053

Fischer, J. M., and Ravizza, M. (2000). Responsibility and Control: A Theory of Moral Responsibility. Cambridge, UK: Cambridge University Press.

Fried, I., Mukamel, R., and Kreiman, G. (2011). Internally generated preactivation ofsingle neurons in human medial frontal cortex predicts volition. Neuron 69, 548–562. doi: 10.1016/j.neuron.2010.11.045

Gailliot, M. T., and Baumeister, R. F. (2007). The physiology of willpower: linking blood glucose to self-control. Pers. Soc. Psychol. Rev. 11, 303–327. doi: 10.1177/1088868307303030

Gibson, J. J. (1979). The Ecological Approach to Visual Perception. Boston, MA: Houghton-Mifflin.

Glannon, W. (Ed.). (2015). Free Will and the Brain: Neuroscientific, Philosophical and Legal Perspectives. Cambridge, UK: Cambridge University Press.

Glimcher, P. W. (2002). Decisions, decisions, decisions: choosing a biological science of choice. Neuron 36, 323–332. doi: 10.1016/S0896-6273(02)00962-5

Gold, J. I., and Shadlen, M. L. (2007). The neural basis of decision making. Annu. Rev. Neurosci. 30, 535–574. doi: 10.1146/annurev.neuro.29.051605.113038

Goldstein, S., and Naglieri, J. A. (Eds). (2014). Handbook of Executive Functioning. Berlin-New York: Springer.

Gollwitzer, P. M. (1999). Implementation intentions: strong effects of simple plans. Am. Psychol. 54:493. doi: 10.1037/0003-066x.54.7.493

Gollwitzer, P. M., and Sheeran, P. (2006). Implementation intentions and goal achievement: a meta-analysis of effects and processes. Adv. Exp. Soc. Psychol. 38, 69–119. doi: 10.1016/s0065-2601(06)38002-1

Greene, J., and Cohen, J. (2004). For the law, neuroscience changes nothing and everything. Philos. Trans. R. Soc. Lond. B Biol. Sci. 359, 1775–1785. doi: 10.1098/rstb.2004.1546

Grèzes, J., and Decety, J. (2002). Does visual perception of object afford action? Evidence from a neuroimaging study. Neuropsychologia 40, 212–222. doi: 10.1016/s0028-3932(01)00089-6

Haggard, P. (2008). Human volition: towards a neuroscience of will. Nat. Rev. Neurosci. 9, 934–946. doi: 10.1038/nrn2497

Haggard, P. (2009). Neuroscience. The sources of human volition. Science 324, 731–733. doi: 10.1126/science.1173827

Haggard, P., and Eimer, M. (1999). On the relation between brain potentials and the awareness of voluntary movements. Exp. Brain Res. 126, 128–133. doi: 10.1007/s002210050722

Hallett, M. (2007). Volitional control of movement: the physiology of free will. Clin. Neurophysiol. 118, 1179–1192. doi: 10.1016/j.clinph.2007.03.019

Hanks, T. D., Ditterich, J., and Shadlen, M. N. (2006). Microstimulation of macaque area LIP affects decision-making in a motion discrimination task. Nat. Neurosci. 9, 682–689. doi: 10.1038/nn1683

Harris, S. (2012). Free Will. New York, NY: Free Press.

Jo, H. G., Hinterberger, T., Wittmann, M., Borghardt, T. L., and Schmidt, S. (2013). Spontaneous EEG fluctuations determine the readiness potential: is preconscious brain activation a preparation process to move? Exp. Brain Res. 231, 495–500. doi: 10.1007/s00221-013-3713-z

Kagaya, K., and Takahata, M. (2010). Readiness discharge for spontaneous initiation of walking in crayfish. J. Neurosci. 30, 1348–1362. doi: 10.1523/JNEUROSCI.4885-09.2010

Kane, R. (2005). A Contemporary Introduction to Free Will. New York, NY: Oxford University Press.

Kane, R. (Ed.). (2011). The Oxford Handbook of Free Will. New York, NY: Oxford University Press.

Kane, R. (2016). The complex tapestry of free will: striving will, indeterminism and volitional streams. Synthese 1–16.

Kornhuber, H. H., and Deecke, L. (1965). Hirnpotentialänderungen bei Willkürbewe-gungen und passiven bewegungen des menschen: bereitschaftspotential undreafferente Potentiale. Pflügers Arch. EJP 284, 1–17. doi: 10.1007/bf00412364

Kornhuber, H. H., and Deecke, L. (1990). Readiness for movement—the Bereitschafts potential-story. Curr. Contents Life Sci. 33:14.

Kühn, S., and Brass, M. (2009). Retrospective construction of the judgement of free choice. Conscious. Cogn 18, 12–21. doi: 10.1016/j.concog.2008.09.007

Kühn, S., Haggard, P., and Brass, M. (2009). Intentional inhibition: how the “veto-area” exerts control. Hum. Brain Mapp. 30, 2834–2843. doi: 10.1002/hbm.20711

Lages, M., Boyle, S. C., and Jaworska, K. (2013). Flipping a coin in your head without monitoring outcomes? Comments on predicting free choices and a demo program. Front. Psychol. 4:925. doi: 10.3389/fpsyg.2013.00925

Lau, H. C., and Passingham, R. E. (2007). Unconscious activation of the cognitive control system in the human prefrontal cortex. J. Neurosci. 27, 5805–5811. doi: 10.1523/jneurosci.4335-06.2007

Lages, M., and Jaworska, K. (2012). How predictable are “Spontaneous Decisions” and “Hidden Intentions”? Comparing classification results based on previous responses with multivariate pattern analysis of fMRI BOLD signals. Front. Psychol. 3:56. doi: 10.3389/fpsyg.2012.00056

Lau, H. C., Rogers, R. D., Haggard, P., and Passingham, R. E. (2004). Attention to intention. Science 303, 1208–1210. doi: 10.1126/science.1090973

Lau, H. C., Rogers, R. D., and Passingham, R. E. (2006). On measuring the perceived onsets of spontaneous actions. J. Neurosci. 26, 7265–7271. doi: 10.1523/JNEUROSCI.1138-06.2006

Lavazza, A., and De Caro, M. (2010). Not so fast: on some bold neuroscientific claimsconcerning human agency. Neuroethics 3, 23–41. doi: 10.1007/s12152-009-9053-9

Lavazza, A., and Inglese, S. (2015). Operationalizing and measuring (a kind of) free will (and responsibility). Towards a new framework for psychology, ethics and law. Riv. Int. di Filos. e Psicol. 6, 37–55. doi: 10.4453/rifp.2015.0004

Levy, N. (2011). Hard Luck: How Luck Undermines Free Will and Moral Responsibility. New York, NY: Oxford University Press.

Levy, N. (Ed.). (2013). Addiction and Self-Control: Perspectives from Philosophy, Psychology and Neuroscience. New York, NY: Oxford University Press.

Libet, B. (1985). Unconscious cerebral initiative and the role of conscious will involuntary action. Behav. Brain Sci. 8, 529–566. doi: 10.1017/s0140525x00044903

Libet, B. (2004). Mind Time: The Temporal Factor in Consciousness. Cambridge, MA: Harvard University Press.

Libet, B., Gleason, C. A., Wright, E. W., and Pearl, D. K. (1983). Time of conscious inten-tion to act in relation to onset of cerebral activities (readiness-potential): theunconscious initiation of a freely voluntary act. Brain 106, 623–642. doi: 10.1093/brain/106.3.623

Logan, G. D., Cowan, W. B., and Davis, K. A. (1984). On the ability to inhibit simple and choice reaction time responses: a model and a method. J. Exp. Psychol. Hum. Percept. Perform. 10, 276–291. doi: 10.1037/0096-1523.10.2.276

Mecacci, G., and Haselager, P. (2015). A reason to be free. Neuroethics 8, 327–334. doi: 10.1007/s12152-015-9241-8

Mele, A. R. (2009). Effective intentions: The Power of Conscious Will. New York, NY: Oxford University Press.

Mele, A. R. (2014). Free. Why Science Hasn’t Disproved Free Will. New York, NY: Oxford University Press.

Metzinger, T. (2004). Being No One: The Self-Model Theory of Subjectivity. Cambridge, MA: The MIT Press.

Metzinger, T. (2009). The Ego Tunnel. The Science of the Soul and the Myth of the Self. New York, NY: Basic Books.

Mirabella, G. (2014). Should i stay or should i go? conceptual underpinnings of goal-directed actions. Front. Syst. Neurosci. 8:206. doi: 10.3389/fnsys.2014.00206

Mischel, W., Ebbesen, E. B., and Raskoff Zeiss, A. (1972). Cognitive and attentional mechanisms in delay of gratification. J. Pers. Soc. Psychol. 21, 204–218. doi: 10.1037/h0032198

Mischel, W., Shoda, Y., and Rodriguez, M. L. (1989). Delay of gratification in children. Science 244, 933–938. doi: 10.1126/science.2658056

Mochizuki, K., and Funahashi, S. (2014). Opposing history effect of preceding decision andaction in the free choice of saccade direction. J. Neurophysiol. 112, 923–932. doi: 10.1152/jn.00846.2013

Murakami, M., Vicente, M. I., Costa, G. M., and Mainen, Z. F. (2014). Neural antecedents of self-initiated actions in secondary motor cortex. Nat. Neurosci. 17, 1574–1582. doi: 10.1038/nn.3826

Nachev, P., and Hacker, P. (2014). The neural antecedents to voluntary action: a conceptual analysis. Cogn. Neurosci. 5, 193–208. doi: 10.1080/17588928.2014.934215

Nahmias, E. (2014). “Is free will an illusion? Confronting challenges from the modern mind sciences,” in Moral Psychology, Vol. 4, Free Will and Moral Responsibility , ed. W. Sinnott-Armstrong (Cambridge, MA: The MIT Press), 1–26.

Passingham, R. E., Bengtsson, S. L., and Lau, H. C. (2010). Medial frontal cortex: from self-generated action to reflection on one’s own performance. Trends Cogn. Sci. 14, 16–21. doi: 10.1016/j.tics.2009.11.001

Platt, M. L., and Glimcher, P. W. (1999). Neural correlates of decision variables in parietal cortex. Nature 400, 233–238. doi: 10.1038/22268

Ratcliff, R. (1978). A theory of memory retrieval. Psychol. Rev. 85, 59–108. doi: 10.1037/0033-295x.85.2.59

Ratcliff, R., and McKoon, G. (2008). The diffusion decision model: theory and data for two-choice decision tasks. Neural Comput. 20, 873–922. doi: 10.1162/neco.2008.12-06-420

Roskies, A. L. (2010a). How does neuroscience affect our conception of volition? Annu. Rev. Neurosci. 33, 109–130. doi: 10.1146/annurev-neuro-060909-153151

Roskies, A. L. (2010b). “Freedom, mechanism, and consciousness,” in Free Will and Consciousness: How Might They Work , eds R. F. Baumeister, A. R. Mele, and K. D. Vohs (New York, NY: Oxford University Press), 153–171.

Roskies, A. L. (2013). “The neuroscience of volition,” in Decomposing the Will , eds A. Clark, J. Kiverstein and T. Vierkant (New York: Oxford University Press), 33–59.

Schel, M. A., Kühn, S., Brass, M., Haggard, P., Ridderinkhof, K. R., and Crone, E. A. (2014). Neural correlates of intentional and stimulus-driven inhibition: a comparison. Front. Hum. Neurosci. 8:27. doi: 10.3389/fnhum.2014.00027

Schultze-Kraft, M., Birman, D., Rusconi, M., Allefeld, C., Görgen, K., Dähne, S., et al. (2016). The point of no return in vetoing self-initiated movements. Proc. Natl. Acad. Sci. U S A 113, 1080–1085. doi: 10.1073/pnas.1513569112

Schurger, A., Mylopoulos, M., and Rosenthal, D. (2016). Neural antecedents of spontaneous voluntary movement: a new perspective. Trends Cogn. Sci. 20, 77–79. doi: 10.1016/j.tics.2015.11.003

Schurger, A., Sitt, J. D., and Dehaene, S. (2012). An accumulator model for spontaneousneural activity prior to self-initiated movement. Proc. Natl. Acad. Sci. U S A 109, E2904–E2913. doi: 10.1073/pnas.1210467109

Schurger, A., and Uithol, S. (2015). Nowhere and everywhere: the causal origin of voluntary action. Rev. Phil. Psych. 6, 761–778. doi: 10.1007/s13164-014-0223-2

Shadlen, M. N., and Newsome, W. T. (2001). Neural basis of a perceptual decision in the parietal cortex (area LIP) of the rhesus monkey. J. Neurophysiol. 86, 1916–1936.

PubMed Abstract | Google Scholar

Sirigu, A., Daprati, E., Ciancia, S., Giraux, P., Nighoghossian, N., Posada, A., et al. (2004). Altered awareness of voluntary action after damage to the parietal cortex. Nat. Neurosci. 7, 80–84. doi: 10.1038/nn1160

Soon, C. S., Brass, M., Heinze, H. J., and Haynes, J. D. (2008). Unconscious determinants offree decisions in the human brain. Nat. Neurosci. 11, 543–545. doi: 10.1038/nn.2112

Soon, C. S., He, A. H., Bode, S., and Haynes, J. D. (2013). Predicting free choices for abstractintentions. Proc. Natl. Acad. Sci. U S A 110, 5733–5734. doi: 10.1073/pnas.1212218110

Sugrue, L. P., Corrado, G. S., and Newsome, W. T. (2004). Matching behavior and the representation of value in the parietal cortex. Science 304, 1782–1787. doi: 10.1126/science.1094765

Suhler, C. L., and Churchland, P. S. (2009). Control: conscious and otherwise. Trends Cogn. Sci. 13, 341–347. doi: 10.1016/j.tics.2009.04.010

Sumner, P., Nachev, P., Morris, P., Peters, A. M., Jackson, S. R., Kennard, C., et al. (2007). Human medial frontal cortex mediates unconscious inhibition of voluntary action. Neuron 54, 697–711. doi: 10.1016/j.neuron.2007.05.016

Trevena, J., and Miller, J. (2010). Brain preparation before a voluntary action: evidence against unconscious movement initiation. Conscious. Cogn. 19, 447–456. doi: 10.1016/j.concog.2009.08.006

Vierkant, T., Kiverstein, J., and Clark, A. (2013). “Decomposing the will: meeting the zombie challenge,” in Decomposing the Will , eds A. Clark, J. Kiverstein and T. Vierkant (New York: Oxford University Press), 1–30.

Vohs, K. (2010). “Free will is costly: action control, making choices, mental time travel and impression management use precious volitional resources,” in Free Will and Consciousness: How Might They Work? eds R. Baumeister, A. R. Mele, and K. Vohs (New York, NY: Oxford University Press), 66–81.

Walter, H. (2001). Neurophilosophy of Free Will: From Libertarian Illusion to a Concept of Natural Autonomy. Cambridge, MA: The MIT Press.

Wegner, D. M. (2002). The Illusion of Conscious Will. Cambridge, MA: The MIT Press.

Wegner, D. M. (2003). The mind’s best trick: how we experience conscious will. Trends Cogn. Sci. 7, 65–69. doi: 10.1016/s1364-6613(03)00002-0

Wegner, D. M. (2004). Précis of the illusion1 of conscious will. Behav. Brain Sci. 27, 649–659. doi: 10.1017/S0140525X04000159

Wong, K.-F., Huk, A. C., Shadlen, M. N., and Wang, X.-J. (2007). Neural circuit dynamics underlying accumulation of time-varying evidence during perceptual decision making. Front. Comput. Neurosci. 1:6. doi: 10.3389/neuro.10.006.2007

Keywords: readiness potential, unconscious decision, choice prediction, stochastic processes, measurement of freedom, evidence accumulation

Citation: Lavazza A (2016) Free Will and Neuroscience: From Explaining Freedom Away to New Ways of Operationalizing and Measuring It. Front. Hum. Neurosci. 10:262. doi: 10.3389/fnhum.2016.00262

Received: 15 March 2016; Accepted: 18 May 2016; Published: 01 June 2016.

Reviewed by:

Copyright © 2016 Lavazza. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution and reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Andrea Lavazza, [email protected]

This article is part of the Research Topic

Paradigm shifts and innovations in Neuroscience

Freewill vs Determinism In Psychology

Saul Mcleod, PhD

Editor-in-Chief for Simply Psychology

BSc (Hons) Psychology, MRes, PhD, University of Manchester

Saul Mcleod, PhD., is a qualified psychology teacher with over 18 years of experience in further and higher education. He has been published in peer-reviewed journals, including the Journal of Clinical Psychology.

Learn about our Editorial Process

Olivia Guy-Evans, MSc

Associate Editor for Simply Psychology

BSc (Hons) Psychology, MSc Psychology of Education

Olivia Guy-Evans is a writer and associate editor for Simply Psychology. She has previously worked in healthcare and educational sectors.

On This Page:

The free will vs. determinism debate revolves around how our behavior results from forces over which we have no control or whether people can decide to act or behave in a certain way.

Determinism

The determinist approach proposes that all behavior has a cause and is thus predictable. Free will is an illusion, and our behavior is governed by internal or external forces over which we have no control.

External Determinism

External (environmental) determinism sees the cause of behavior as being outside the individual, such as parental influence, the media, or school. Approaches that adopt this position include behaviorism and social learning theory .

For example, Bandura (1961) showed that children become aggressive through observation and imitation of their violent parents.

Internal Determinism

The other main supporters of determinism are those who adopt a biological perspective .

However, for them, it is internal, not external, forces that are the determining factor. According to sociobiology , evolution governs the behavior of a species and the genetic inheritance that of each individual within it.

For example, Bowlby (1969) states a child has an innate (i.e., inborn) need to attach to one main attachment figure (i.e., monotropy).

Personality traits like extraversion or neuroticism and the behavior associated with them are triggered by neurological and hormonal processes within the body. There is no need for the concept of an autonomous human being.

Ultimately this view sees us as no more than biological machines, and even consciousness itself is interpreted as a level of arousal in the nervous system.

Psychic Determinism

Freud also viewed behavior as being controlled from inside the individual through unconscious motivation or childhood events, known as psychic determinism.

Freud used this principle to explain phenomena like slips of the tongue (“Freudian slips”), dreams, and symptoms of mental disorders, arguing that they all have meaningful explanations rooted in the individual’s unconscious mind.

Different levels of determinism

Hard determinism.

Hard determinism sees free will as an illusion and believes that every event and action has a cause.

Behaviorists are strong believers in hard determinism. Their most forthright and articulate spokesman has been B. F. Skinner. Concepts like “free will” and “motivation” are dismissed as illusions that disguise the real causes of human behavior.

In Skinner’s scheme of things, the person who commits a crime has no real choice. (S)he is propelled in this direction by environmental circumstances and a personal history, which makes breaking the law natural and inevitable.

For the law-abiding, an accumulation of reinforcers has the opposite effect. Having been rewarded for following rules in the past, the individual does so in the future. There is no moral evaluation or even mental calculation involved. All behavior is under stimulus control.

Soft Determinism

Soft determinism represents a middle ground, people do have a choice, but that choice is constrained by external or internal factors.

For example, being poor doesn’t make you steal, but it may make you more likely to take that route through desperation.

Soft determinism suggests that some behaviors are more constrained than others and that there is an element of free will in all behavior.

However, a problem with determinism is that it is inconsistent with society’s ideas of responsibility and self-control that form the basis of our moral and legal obligations.

An additional limitation concerns the fact that psychologists cannot predict a person’s behavior with 100% accuracy due to the complex interaction of variables that can influence behavior.

Free will is the idea that we are able to have some choice in how we act and assumes that we are free to choose our behavior. In other words, we are self-determined.

For example, people can make a free choice as to whether to commit a crime or not (unless they are a child or they are insane).

This does not mean that behavior is random, but we are free from the causal influences of past events. According to free will a person is responsible for their own actions.

One of the main assumptions of the humanistic approach is that humans have free will; not all behavior is determined. Personal agency is the humanistic term for the exercise of free will. Personal agency refers to the choices we make in life, the paths we go down, and their consequences.

For humanistic psychologists such as Maslow (1943) and Rogers (1951), freedom is not only possible but also necessary if we are to become fully functional human beings. Both see self-actualization as a unique human need and form of motivation setting us apart from all other species. There is, thus, a line to be drawn between the natural and the social sciences.

To take a simple example, when two chemicals react, there is no sense in imagining that they could behave in any other way than the way they do. However, when two people come together, they could agree, fall out, come to a compromise, start a fight, and so on.

The permutations are endless, and in order to understand their behavior, we would need to understand what each party to the relationship chooses to do.

Ranged against the deterministic psychologies of those who believe that what “is” is inevitable are, therefore, those who believe that human beings have the ability to control their own destinies. However, there is also an intermediate position that goes back to the psychoanalytic psychology of Sigmund Freud .

At first sight, Freud seems to be a supporter of determinism in that he argued that our actions and our thoughts are controlled by the unconscious . However, the very goal of therapy was to help the patient overcome that force. Indeed without the belief that people can change therapy itself makes no sense.

This insight has been taken up by several neo-Freudians. One of the most influential has been Erich Fromm (1941). In “Fear of Freedom,” he argues that all of us have the potential to control our own lives but that many of us are too afraid to do so.

As a result, we give up our freedom and allow our lives to be governed by circumstances, other people, political ideologies, or irrational feelings. However, determinism is not inevitable, and in the very choice we all have to do good, or evil, Fromm sees the essence of human freedom.

Critical Evaluation

Psychologists who take the free will view suggest that determinism removes freedom and dignity and devalues human behavior.

By creating general laws of behavior, deterministic psychology underestimates the uniqueness of human beings and their freedom to choose their own destiny.

There are important implications for taking either side in this debate. Deterministic explanations for behavior reduce individual responsibility.

A person arrested for a violent attack, for example, might plead that they were not responsible for their behavior – it was due to their upbringing, a bang on the head they received earlier in life, recent relationship stresses, or a psychiatric problem. In other words, their behavior was determined.

The deterministic approach also has important implications for psychology as a science. Scientists are interested in discovering laws that can then be used to predict events. This is very easy to see in physics, chemistry, and biology.

As a science, psychology attempts the same thing – to develop laws, but this time to predict behavior. If we argue against determinism, we are, in effect, rejecting the scientific approach to explaining behavior

Mental illnesses appear to undermine the concept of free will. For example, individuals with OCD lose control of their thoughts and actions, and people with depression lose control over their emotions.

Clearly, a pure deterministic or free will approach does not seem appropriate when studying human behavior. Most psychologists use the concept of free will to express the idea that behavior is not a passive reaction to forces but that individuals actively respond to internal and external forces.

The term soft determinism is often used to describe this position, whereby people do have a choice, but their behavior is always subject to some form of biological or environmental pressure.

Bandura, A. Ross, D., & Ross,S.A (1961). Transmission of aggression through the imitation of aggressive models. Journal of Abnormal and Social Psychology , 63, 575-582

Bowlby, J. (1969). Attachment. Attachment and Loss: Vol. 1. Loss . New York: Basic Books.

Chorney, M. J., Chorney, K., Seese, N., Owen, M. J., Daniels, J., McGuffin, P., … & Plomin, R. (1998). A quantitative trait locus associated with cognitive ability in children. Psychological Science , 9(3), 159-166.

Fromm, E. (1941). Escape from freedom.

Maslow, A. H. (1943). A Theory of Human Motivation. Psychological Review , 50(4), 370-96.

Rogers, C. (1951). Client-centered Therapy: Its Current Practice, Implications and Theory . London: Constable.

Skinner, B. F. (1957). Verbal behavior . Acton, MA: Copley Publishing Group.

is free will an illusion essay

The Illusion of Free Will: “The System of Nature” by Paul Holbach Essay

Paul Holbach is a famous French philosopher and writer. “The System of Nature” is the most renowned book of the writer, which deals with materialism. “The illusion of free will” is part of this book and reveals the philosopher’s ideas about the origin of human will. According to Paul Holbach, human freedom and choice are only a way of performing specific actions and the ability to accept all the consequences of one’s actions correctly.

Holbach’s first argument is that life is no longer our direct choice. According to the writer, birth and living one’s own life is the first and foremost indicator that the will is not inherent in a person. Holbach claims that life and all further human existence fulfill various requirements. For the philosopher, nature is the basis of control: “Man’s life is a line that nature commands him to describe upon the surface of the earth, without his ever being able to swerve from it, even for an instant” (Holbach 438). In this way, life is a direct indicator of human unfreedom.

Even though a person can change it at any moment of his life, these changes will not depend on the person’s wishes. Just as a person is always directly dependent on society, the desires that arise will depend on the environment. Holbach emphasizes that the changes a person tries to implement in his life are just impulses that drive us (439). Most often, the individual’s freedom depends on other objects that affect this person at such a moment.

The emergence of an idea in a person is not the will and freedom of a person. Holbach denies that the emergence of different ideas is a directly independent factor. The emergence of opinions or thoughts is a synthesis of the environment and analysis of people around (Holbach 441). In this way, the ideas that arise in an individual do not depend on the person himself but on everything that is currently in a person’s life.

Human will is only a conditional fact that people use. It is only a set of actions of society and the environment in which they are located. It is worth saying that there is an illusion of freedom, which depends on many factors in society. It is only necessary to understand and be able to use these ideas correctly.

Holbach, Baron. “The illusion of free will.” from The System of Nature (1770): 176-181.

  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2023, June 27). The Illusion of Free Will: “The System of Nature” by Paul Holbach. https://ivypanda.com/essays/the-illusion-of-free-will-the-system-of-nature-by-paul-holbach/

"The Illusion of Free Will: “The System of Nature” by Paul Holbach." IvyPanda , 27 June 2023, ivypanda.com/essays/the-illusion-of-free-will-the-system-of-nature-by-paul-holbach/.

IvyPanda . (2023) 'The Illusion of Free Will: “The System of Nature” by Paul Holbach'. 27 June.

IvyPanda . 2023. "The Illusion of Free Will: “The System of Nature” by Paul Holbach." June 27, 2023. https://ivypanda.com/essays/the-illusion-of-free-will-the-system-of-nature-by-paul-holbach/.

1. IvyPanda . "The Illusion of Free Will: “The System of Nature” by Paul Holbach." June 27, 2023. https://ivypanda.com/essays/the-illusion-of-free-will-the-system-of-nature-by-paul-holbach/.

Bibliography

IvyPanda . "The Illusion of Free Will: “The System of Nature” by Paul Holbach." June 27, 2023. https://ivypanda.com/essays/the-illusion-of-free-will-the-system-of-nature-by-paul-holbach/.

  • Nature vs. Technology: Traveling Through the Dark by Stafford
  • Travelling through the dark
  • ‘Must an Essay Be 1000 Words’ or ’Just Must Be Good'?
  • Globalization and Nation States
  • Materialism Concept and Theorists Views
  • Epistemology and Materialism: History and Application
  • An Evaluation of Complexities and Difficulties of Learning Second and Third Conditionals for Iranian Students
  • "On Functionalism and Materialism" by Paul Churchland
  • Materialism and the Theory of Consciousness
  • Berkeley’s Argument on Materialism Analysis
  • “Oration on the Dignity of Man” by Mirandola
  • The Essay "Nature" by R. W. Emerson
  • How Socratic Dialogue Differs From Other Conversations
  • Civility, Democracy, Memory in Sophocles’ Antigone
  • Race Matters by Cornel West: A Book Review

Quality Stock Arts/Shutterstock

Reviewed by Psychology Today Staff

Free will is the idea that humans have the ability to make their own choices and determine their own fates. Is a person’s will free, or are people's lives in fact shaped by powers outside of their control? The question of free will has long challenged philosophers and religious thinkers, and scientists have examined the problem from psychological and neuroscientific perspectives as well.

  • Does Free Will Exist?
  • Why Beliefs About Free Will Matter

Pixel-Shot/Shutterstock

Scientists have investigated the concept of human agency at the level of neural circuitry, and some findings have been taken as evidence that conscious decisions are not truly “free.” Free will skeptics argue that the subjective sense of free will is an illusion. Yet many scholars, as well as ordinary people, still profess a belief in free will, even if they acknowledge that choices are partly shaped by forces outside of one's control.

Behavioral science has made plain that individuals’ behavioral tendencies are influenced by genetics , as well as by factors in the environment that may be outside of a person’s control. This suggests that there are, at least, some constraints on the range of decisions and behaviors a person will be inclined to make (or even to consider) in any given situation. Challengers of the idea that people act the way they do due to conscious, unconstrained choices also point to evidence that unconscious brain activity can partly predict a choice before a conscious decision is made. And some have sought to logically refute the argument that choices necessarily demonstrate free will.

While there are many reasons to believe that a person’s will is not completely free of influence, there is not a scientific consensus against free will. Some use the term “free will” in a looser sense to reflect that conscious decisions play a role in the outcomes of a person’s life—even if those are shaped by innate dispositions or randomness. (Critics of the concept of free will might simply call this kind of decision-making “will,” or volition.) Even when unconscious processes help determine a person’s conscious behavior, some argue, such processes can still be thought of as part of an individual’s will.

Determinism is the idea that every event, including every human action, is the result of previous events and the laws of nature. A belief in determinism that includes a rejection of free will has been called “hard determinism.”

From a deterministic perspective, there is only one possible way that future events can unfold based on what has already occurred and the rules that govern the universe—though that doesn’t mean such events can necessarily be predicted by humans. Someone who believes in free will because they do not take determinism for granted is called, in philosophy , a “libertarian.”

Yes. This is called “compatibilism” or “soft determinism.” A compatibilist believes that even though events are predetermined, there is still some version of free will at work in decision-making . An incompatibilist argues that only determinism or free will can be true.

Pixabay

Whether free will exists or not, belief in free will is very real. Does it matter if a person believes that her choices are completely her own, and that other people’s choices are freely made, too? Psychologists have explored the connections between free will beliefs—often gauged by agreement with statements like “I am in charge of my actions even when my life’s circumstances are difficult” and, simply, “I have free will”—and people’s attitudes about decision-making, blame, and other variables of consequence.

The more people agree with claims of free will , some research suggests, the more they tend to favor internal rather than external explanations for someone else’s behavior. This may include, for example, learning of someone’s immoral deed and agreeing more strongly that it was a result of the person’s character and less that factors like social norms are to blame. (A study of whether reducing free-will beliefs influenced sentencing decisions by actual judges, however, did not show any effect. )

One idea proposed in philosophy is that systems of morality would collapse without a common belief that each person is responsible for his actions—and thus deserves reward or punishment for them. In this view, there is value in maintaining belief in free will, even if free will is in fact an illusion. Others argue that morality can exist in the absence of free-will belief, or that belief in free will actually promotes harmful outcomes such as intolerance and revenge-seeking. Some psychology research has been cited as suggesting that disbelief in free will increases dishonest behavior, but subsequent experiments have called this finding into question.

Mental illness can be thought of, in a sense, as involving additional constraints on the freedom of a person’s will (in the form of rigid thought patterns or compulsions, for instance), beyond the usual factors that shape thinking and behavior. Belief in free will, it has been argued, may contribute to the stigma attached to mental illness by obscuring the role of underlying biological and environmental causes.

There is limited evidence that people who believe more strongly in free will may tend to perceive at least some kinds of choices—such as buying electronics or deciding what to watch on TV—as easier to make, and that they may enjoy making choices more.

Two concepts from psychology that bear similarity to belief in free will are “ locus of control ” and “ self-efficacy .” Locus of control refers to a person’s belief about how much power he has over his life—how important factors like intentions and hard work seem to be compared to external forces such as good luck or the actions of others. Self-efficacy is a person’s sense of her ability to perform at a certain level so as to influence events that affect them. While all of these concepts relate to the factors that steer a person’s life, they are distinct—one can doubt that humans have free will, for example, and still be confident in her ability to win a competition .

is free will an illusion essay

Individuals with serious and persistent mental illnesses are held to a different standard in physician-assisted dying.

is free will an illusion essay

Kierkegaard's philosophy offers insights into living authentically and finding fulfillment amidst the distractions and pressures of the modern world.

is free will an illusion essay

If the days of bettering oneself through a liberal arts education have transformed into simply “making more money,” then let’s just teach kids to write ransom notes.

is free will an illusion essay

As brains evolved in complexity, they acquired top-down control systems, such that bottom-up inputs cannot completely predict behavioral outputs. Does this mean we have free-will?

is free will an illusion essay

New theoretical work helps us understand which aspects of emotional processing can and cannot be controlled.

is free will an illusion essay

To help ensure you actually reach your New Year's resolutions (or any goal), here is an essential technique you need in your repertoire.

is free will an illusion essay

The "mysteries" of concern to Freud and Jung were restricted to semantic processes, which today are not a mystery. But consciousness remains a true mystery.

is free will an illusion essay

Is it possible for a tiny decision to make a big impact? Surprisingly, infinitesimal self-help practices can be effective, when bigger commitments will fail.

is free will an illusion essay

Dopamine is the molecule that makes us feel good about expending effort. Motivation neurobiology reveals ways to enhance effort without extra calories, drugs, or risk.

is free will an illusion essay

I’m willing to bet that you’re never late for a truly important meeting, especially not when the meeting is a matter of life and death.

  • Find a Therapist
  • Find a Treatment Center
  • Find a Psychiatrist
  • Find a Support Group
  • Find Teletherapy
  • United States
  • Brooklyn, NY
  • Chicago, IL
  • Houston, TX
  • Los Angeles, CA
  • New York, NY
  • Portland, OR
  • San Diego, CA
  • San Francisco, CA
  • Seattle, WA
  • Washington, DC
  • Asperger's
  • Bipolar Disorder
  • Chronic Pain
  • Eating Disorders
  • Passive Aggression
  • Personality
  • Goal Setting
  • Positive Psychology
  • Stopping Smoking
  • Low Sexual Desire
  • Relationships
  • Child Development
  • Therapy Center NEW
  • Diagnosis Dictionary
  • Types of Therapy

March 2024 magazine cover

Understanding what emotional intelligence looks like and the steps needed to improve it could light a path to a more emotionally adept world.

  • Coronavirus Disease 2019
  • Affective Forecasting
  • Neuroscience

Home — Essay Samples — Philosophy — Free Will — The Illusion of Free Will

test_template

The Illusion of Free Will

  • Categories: Free Will

About this sample

close

Words: 584 |

Published: Jan 31, 2024

Words: 584 | Page: 1 | 3 min read

Table of contents

Theoretical perspectives on free will, scientific evidence for determinism, debunking free will arguments, ethical implications of determinism, critiques and limitations of determinism.

Image of Dr. Charlotte Jacobson

Cite this Essay

Let us write you an essay from scratch

  • 450+ experts on 30 subjects ready to help
  • Custom essay delivered in as few as 3 hours

Get high-quality help

author

Dr. Heisenberg

Verified writer

  • Expert in: Philosophy

writer

+ 120 experts online

By clicking “Check Writers’ Offers”, you agree to our terms of service and privacy policy . We’ll occasionally send you promo and account related email

No need to pay just yet!

Related Essays

3 pages / 1221 words

4 pages / 1915 words

3 pages / 1474 words

3 pages / 1220 words

Remember! This is just a sample.

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

121 writers online

Still can’t find what you need?

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

Related Essays on Free Will

Cunning, David, editor. The Cambridge Companion to Descartes’ Meditations. Cambridge University Press, 2014.Gregory, Andrew. “Leucippus and Democritus on Like to Like and Ou Mallon.” Apeiron, vol. 46, no. 4, 2013, pp. [...]

Soft determinism is a concept that delves into the intricate relationship between free will and determinism, offering a nuanced perspective that acknowledges the presence of external influences while still emphasizing individual [...]

Kurt Vonnegut’s Slaughterhouse-Five has been the subject of much attention and debate since its release. Its wide range of topics such as critique of the American government and discussion of existentialism have made it an [...]

Humans have struggled with the concept of freedom and free-will since the Stoic philosophers debated the nature of being. We are aware of our existence in a larger sense, aware of the decisions we’re capable of making and the [...]

In conclusion, Slaughterhouse-Five offers a complex and nuanced analysis of free will. Through its unique narrative structure and thought-provoking characters, the novel challenges traditional notions of agency and determinism. [...]

War has, undisputedly, been an element of every civilization’s history throughout time, but the cause of war, however, is a topic of dispute. Is war something that humans bring on themselves, or has it been deemed inevitable, no [...]

Related Topics

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

Where do you want us to send this sample?

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

Be careful. This essay is not unique

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

Download this Sample

Free samples may contain mistakes and not unique parts

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

Please check your inbox.

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

Get Your Personalized Essay in 3 Hours or Less!

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

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

is free will an illusion essay

IMAGES

  1. 📗 Free Will Is an Illusion Essay Example

    is free will an illusion essay

  2. Free Will Essay

    is free will an illusion essay

  3. Free Will Essay

    is free will an illusion essay

  4. Is Free Will an Illusion?

    is free will an illusion essay

  5. Is free will an illusion?

    is free will an illusion essay

  6. Free will is an illusion, biologist says

    is free will an illusion essay

VIDEO

  1. #illusion #amazingfacts #opticalillusion #motivation #story #fact #trending #illusions

  2. THE ILLUSION THAT A CARDBOARD SNACK GET BIGGER!#asmr

  3. Optical Illusion 3

  4. illusion #illusion #opticalillusion #amazingfacts #factsinhindi #fact #viral #subscribe

  5. #trending #illusion #youtubeshorts

  6. 3d illusion bhaut essay hai 😲☣️#illusion #youtubeshort

COMMENTS

  1. Is Free Will an Illusion?

    According to their view, free will is a figment of our imagination. No one has it or ever will. Rather our choices are either determined—necessary outcomes of the events that have happened in ...

  2. The clockwork universe: is free will an illusion?

    That kind of free will is "as illusory as poltergeists", to quote Dennett. What they claim instead is that this doesn't matter: that even though our choices may be determined, it makes sense ...

  3. There's No Such Thing as Free Will

    Smilansky advocates a view he calls illusionism—the belief that free will is indeed an illusion, but one that society must defend. The idea of determinism, and the facts supporting it, must be ...

  4. PDF Is free will an illusion?

    Is free will an illusion? Scientists and philosophers are using new discoveries in neuroscience to question the idea of free will. They ... 14.5 Essay 164-165 MH AB.indd Author:

  5. Is free will an illusion?

    Scientists and philosophers are using new discoveries in neuroscience to question the idea of free will. They are misguided, says Martin Heisenberg. Examining animal behaviour shows how our ...

  6. Free Will Is Only an Illusion if You Are, Too

    In 2008 a group of researchers found that some information about an upcoming decision is present in the brain up to 10 seconds in advance, long before people reported making the decision of when ...

  7. 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 ...

  8. 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.

  9. 4.3: The Illusion of Free Will

    The inward persuasion that we are free to do, or not to do a thing, is but a mere illusion. If we trace the true principle of our actions, we shall find, that they are always necessary consequences of our volitions and desires, which are never in our power. You think yourself free, because you do what you will; but are you free to will, or not ...

  10. Free Will Is an Illusion, So What?

    Understanding that free will is an illusion means recognizing that people behave in the only way they know how. As such, it is important to realize that, when people act in harmful ways, it is ...

  11. Free Will

    Free Will and Illusion (Clarendon Press). Strawson, Galen (1994). "The Impossibility of Moral Responsibility," Philosophical ... An Essay on Free Will (Clarendon Press). Widerker, David and Michael McKenna (2003). Moral Responsibility and Alternative Possibilities: Essays on the Importance of Alternative Possibilities (Ashgate). Author ...

  12. Exploring the Illusion of Free Will and Moral Responsibility

    Free will skepticism, which Caruso tells us has become increasingly popular in recent decades, questions the existence of free will without necessarily adopting the truth of causal determinism. There are, as the authors of the sixteen chapters point out, a wide range of philosophical and scientific reasons for regarding free will as an illusion.

  13. It's not an illusion, you have free will. It's just not what you think

    In an attempt to dispel some of these misconceptions, I have created an interactive essay on Twitter called The Choice Engine. Waspish behaviour How Sphex came to be linked with free will is a ...

  14. Free Will Is Not an Illusion

    Free will is deemed an illusion. Conscious mind is informed after the fact, and it generates the illusion that it has agency. Freudian psychology is reborn in the framework of a pre-eminent ...

  15. Why the Classical Argument Against Free Will Is a Failure

    The older argument against free will is based on the assumption that determinism is true. Determinism is the view that every physical event is completely caused by prior events together with the laws of nature. Or, to put the point differently, it's the view that every event has a cause that makes it happen in the one and only way that it could have happened.

  16. Free Will Is An Illusion Philosophy Essay

    It is clear that free will is a weak position, and it could be claimed it is an illusion. Determinism can follow off into many routes, as Skinner has proved- a psychological approach, and Schopenhauer through an approach of our character and motive. But it is clear that determinism holds a very strong stance.

  17. Answering The Question on Whether Free Will is an Illusion

    They suppose that God is in control of the whole universe, including our lives, and they believe that God had already pre-destined each and every life in advance. To put it simply, Rationalists advocated that Man has no free will, so, free will is an illusion according to them. Spinoza, in particular, claims that "we don't control ...

  18. Free Will Is an Illusion, So What?

    Understanding that free will is an illusion means recognizing that people behave in the only way they know how. As such, it is important to realize that, when people act in harmful ways, it is ...

  19. Frontiers

    The concept of free will is hard to define, but crucial to both individual and social life. For centuries people have wondered how freedom is possible in a world ruled by physical determinism; however, reflections on free will have been confined to philosophy until half a century ago, when the topic was also addressed by neuroscience. The first relevant, and now well-known, strand of research ...

  20. Freewill vs Determinism In Psychology

    Hard determinism sees free will as an illusion and believes that every event and action has a cause.. Behaviorists are strong believers in hard determinism. Their most forthright and articulate spokesman has been B. F. Skinner. Concepts like "free will" and "motivation" are dismissed as illusions that disguise the real causes of human behavior.

  21. The Illusion of Free Will: "The System of Nature" by Paul Holbach Essay

    Paul Holbach is a famous French philosopher and writer. "The System of Nature" is the most renowned book of the writer, which deals with materialism. "The illusion of free will" is part of this book and reveals the philosopher's ideas about the origin of human will. According to Paul Holbach, human freedom and choice are only a way of ...

  22. Free Will

    Free will skeptics argue that the subjective sense of free will is an illusion. Yet many scholars, as well as ordinary people, still profess a belief in free will, even if they acknowledge that ...

  23. The Illusion of Free Will: [Essay Example], 584 words

    Conclusion. In conclusion, despite our belief in free will, scientific evidence challenges the illusion of free will by suggesting that our choices are predetermined by various factors such as brain activity, external influences, and genetic and environmental factors.