Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

hypothesis is demonstrably true of those beings.

It follows that the


hypothesis is meaningful. It may also be true of us, or it may not. Perhaps
we will never know the answer to the question, but the hypothesis is either
true or false all the same.
What about the original simulation hypothesis, saying that our world is
a computer simulation? Is this a scientific hypothesis or a philosophical
hypothesis?
The philosopher of science Karl Popper insisted that the hallmark of a
scientific hypothesis is that it is falsifiable—capable of being proved false
using scientific evidence. We’ve seen that the simulation hypothesis is not
falsifiable because any evidence against it could be simulated. So Popper
would say that it’s not a scientific hypothesis.
Like many philosophers these days, I think Popper’s criterion is too
strong. There can be scientific hypotheses—for example, about the early
universe—that could never be falsified. But I’m happy to say that the
simulation hypothesis is not a squarely scientific hypothesis but one that is
partly scientific and partly philosophical. Some versions of it are subject to
test, while other versions of it are impossible to test. But whether testable or
not, the simulation hypothesis is a perfectly meaningful hypothesis about
our world.

The simulation hypothesis and the virtual-world


hypothesis

What is the relationship between computer simulations and virtual worlds?


Recall that a virtual world is an interactive, computer-generated space. Is
every virtual world a simulation? Is every simulation a virtual world?
Most virtual worlds found in video games can be regarded as
simulations. This is most obvious in games that simulate some physical-
world activity, like fishing or flying or playing basketball. These games are
closest to Baudrillard’s representations. They may not aim for perfect
realism, but they try to reflect the real world. More exotic games, like Space
Invaders and World of Warcraft, are closer to Baudrillard’s simulacra. They
don’t purport to reflect the real world, but they’re simulations of possible
worlds. Space Invaders loosely simulates an alien invasion of Earth. World
Chapter 6

What is reality?

A T THE END OF THE MOVIE READY PLAYER ONE, WHICH TAKES place mainly
in a virtual world, one of the characters raises a version of the Reality
Question. He says, “Reality is the only thing that’s real.”
At first, this looks like a tautology, akin to Boys are boys. Of course
only reality is real!
On second glance, it looks like a confusion, as with Happiness is happy.
How could reality itself be real?
Still, the underlying message is pretty clear. In the context of the movie,
the speaker is praising physical reality and downgrading virtual reality. The
intended message: Physical reality is the only thing that’s real. And Virtual
reality is not real.
Isn’t this still a confusion? What is it for physical or virtual reality to be
“real”? On my reading, the central idea is that a reality is real if things in
that reality are real. Read this way, the slogan is saying Physical things are
the only things that are real, and Virtual things are not real.
If the slogan is right, the planet Earth is real, whereas the planet Ludus
(a virtual planet in Ready Player One) is not. Earth and the things on it,
from ducks to mountains, exist as part of the objective world. Ludus and the
things on it, from avatars to virtual weapons, do not—they’re mere fictions
or illusions.
Virtual things are not real is the standard line on virtual reality. I think
it’s wrong. Virtual reality is real—that is, the entities in virtual reality really
exist.
My view is a sort of virtual realism. That phrase first appeared as the
title of the American philosopher Michael Heim’s pioneering 1998 book on
the ramifications of VR. Heim used the label primarily for a broad social
and political view of virtual reality, intermediate between that of “network
idealists who promote virtual communities” and “naive realists who blame
electronic culture” for various social ills. At the same time, Heim associated
the label with the view that “Virtual entities are indeed real, functional, and
even central to life in coming eras.” I’m using the label virtual realism in
the latter sense.
As I understand it, virtual realism is the thesis that virtual reality is
genuine reality, with emphasis especially on the view that virtual objects are
real and not an illusion. In general, “realism” is the word philosophers use
for the view that something is real. Someone who thinks morality is real is a
moral realist. Someone who thinks that colors are real is a color realist. By
analogy, someone who believes that virtual objects are real is a virtual
realist.
I also accept simulation realism: If we’re in a simulation, the objects
around us are real and not an illusion. Virtual realism is a view about virtual
reality in general, while simulation realism is a view specifically about the
simulation hypothesis. Simulation realism says that even if we’ve lived our
whole life in a simulation, the cats and chairs in the world around us really
exist. They aren’t illusions; things are as they seem. Most of what we
believe in the simulation is true. There are real trees and real cars. New
York, Sydney, Donald Trump, and Beyoncé are all real.
Simulation realism has major implications for skepticism about the
external world. We’ve seen that the Cartesian path to global skepticism says
no to the Reality Question (in a simulation, nothing is real) and says no to
the Knowledge Question (we can’t know we’re not in a simulation). When
we accept simulation realism, we say yes to the Reality Question. In a
simulation, things are real and not illusions. If so, the simulation hypothesis
and related scenarios no longer pose a global threat to our knowledge. Even
if we don’t know whether or not we’re in a simulation, we can still know
many things about the external world.
Of course, if we’re in a simulation, the trees and cars and Beyoncé are
not exactly how we thought they were. Deep down, there are some
differences. We thought that trees and cars and human bodies were
ultimately made of fundamental particles such as atoms and quarks; instead,
they’re made of bits.
I call this view virtual digitalism. Virtual digitalism says that objects in
virtual reality are digital objects—roughly speaking, structures of binary
information, or bits.
Virtual digitalism is a version of virtual realism, since digital objects are
perfectly real. Structures of bits are grounded in real processing, in a real
computer. If we’re in a simulation, the computer is in the next world up,
metaphorically speaking. But the digital objects are no less real for that. So
if we’re in a simulation, the cats, trees, and tables around us are all perfectly
real.
That might sound preposterous, but I’ll try to convince you it’s true. I’ll
argue for these views over the next few chapters. In chapters 7–9, I’ll make
the case that if we’re in a simulation, our world is still real. In chapters 10–
12, I’ll concentrate on more familiar virtual reality technology, making the
case that here, too, virtual worlds and objects are real.
First, though, we should elucidate what we mean by reality and real.

Reality and realities

Before defining real, I’ll say something about defining reality. This word
has at least three meanings that are relevant for our purposes. I use the word
reality in all these ways in this book.
First, when we talk about reality as an entity, we mean something like
everything that exists: the entire cosmos. When we talk about physical
reality and virtual reality as entities, we mean something in the same spirit.
We mean roughly everything that’s physical and everything that’s virtual.
Second, when we talk about a reality, we mean something like a world.
When we talk about realities plural, we mean worlds. When we talk about a
virtual reality, we mean roughly a virtual world. A world is roughly
equivalent to a universe: a complete interconnected physical or virtual
space.
Reality (in the first sense) may contain many realities (in the second
sense). It’s a familiar idea—depicted for example in the movie Spider-Man:
Into the Spider-Verse—that we could be living in a multiverse: a cosmos
with multiple universes. In a multiverse, reality contains many worlds and
many realities. With the advent of virtual worlds, we have Reality+: a
multiverse of both physical and virtual realities. All of these realities
(worlds) are part of reality (the cosmos).
Third, we can talk about reality as a property like rigidity. Rigidity is
the property of being rigid. Some objects are rigid, and some are not. In this
sense, reality is real-ness. It’s the property of being real. Some things are
real, and others are not. Talking about reality as real-ness is a way of talking
about what it is to be real. That’s our focus in this chapter.
To really confuse things, we could sum up the Reality+ view of reality
(in all three senses) by saying: reality contains many realities, and those
realities are real. Or more mundanely: the cosmos (everything that exists)
contains many worlds (physical and virtual spaces), and the objects in those
worlds are real. Now we just need to unpack real.

Five ways of thinking about what’s real

In The Matrix, Morpheus asks, “What is real? How do you define real?”
That is, What does it mean to say something is real? What does it mean
when we say that Joe Biden is real but Santa Claus isn’t?
Philosophers have answered Morpheus’s question in various ways. I’ll
focus on five complementary answers here. Each answer illuminates part of
the story. Each is a different strand in our concept of the real.

Reality as existence. First and foremost, something is real if it really exists.


Joe Biden really exists; he’s part of the universe. Santa Claus doesn’t really
exist; he’s not part of the universe. There are stories about Santa Claus, and
there are beliefs about Santa Claus, but Santa Claus himself does not exist.
Figure 16 In a virtual world, Morpheus and Neo debate digital reality. Virtual digitalism holds
that virtual objects are real digital objects.

Of course, this just raises a further question: “What is it to exist?” This


is a deep question, to which I can’t give a definitive answer. Many people
think there’s no definitive answer to be given.
Recall Berkeley’s dictum from chapter 4: esse is percipi. To be is to be
perceived—which means that something exists if it’s perceived or at least
could be perceived. As Morpheus put it, when we say something is “real,”
we might be talking about “what you can feel, what you can smell, what
you can taste and see.” A related, more scientific-sounding view is that
something is real if it can be measured.
I’ve already discussed why Berkeley’s dictum is too strong. There could
be real things that we can never perceive and never measure. It may be that
we can never measure certain physical entities around the time of the Big
Bang, or in distant galaxies, but they exist all the same. If I’m in a perfect
simulation, I may never be able to perceive the world outside the
simulation, but it’s real all the same. There are also some things we can
perceive and measure that aren’t real. We can measure the intensity of a
mirage, for example, but this doesn’t mean the mirage really exists.
Still, Berkeley’s dictum can function well as a heuristic for existence—
that is, an imperfect guide to what exists—rather than an absolute criterion.
If something is perceivable and measurable, that’s a strong indication that it
exists.

Reality as causal power. An even better heuristic for existence—one that


subsumes Berkeley’s dictum—is sometimes called the Eleatic dictum,
because a version of it is suggested by the mysterious “Eleatic Stranger”
from the ancient town of Elea in Plato’s dialogue The Sophist. The Stranger
says,

“I suggest that everything which possesses any power of any kind,


either to produce a change in anything of any nature or to be affected
even in the least degree by the slightest cause, though it be only on one
occasion, has real being.”

That’s the Eleatic dictum: to be real is to have causal powers. That is,
something exists if and only if it can affect things or be affected by things.
Joe Biden has causal powers; as US president, he can command the armed
forces and sign or veto legislation. Physical events around the Big Bang and
in distant galaxies have causal powers; they affect what happens next in
their locale. Even a stone has causal powers; it can leave an indentation on
the ground. Anything perceivable has causal powers because it has the
power to make a difference to a perceiver. As a result, anything that falls
under Berkeley’s dictum also falls under the Eleatic dictum.
Santa Claus does not fall under the Eleatic dictum. Santa has no causal
powers: He doesn’t make a difference in the world. According to stories
about Santa Claus, Santa has enormous causal powers: He delivers billions
of Christmas presents in a single night. But those stories are false; Santa has
no intrinsic causal powers. Of course, the stories themselves have causal
powers. They affect cards and costumes and children throughout the world.
So by the Eleatic dictum, the stories are real, but this doesn’t mean Santa is
real.
The Eleatic dictum isn’t the whole truth about reality. There could be
real things that have no causal powers. Maybe numbers are real without
having causal powers, for example. Maybe there could be a forgotten dream
with no causal powers. But causal powers provide at least a sufficient
condition for reality. If something has causal powers, then it exists and it is
real. In this way, causal powers add a second heuristic criterion for
something’s being real, alongside existence itself.

Reality as mind-independence. A third criterion for reality was proposed


by the science-fiction author Philip K. Dick in his 1980 short story “I Hope
I Shall Arrive Soon.” We can call it Philip K. Dick’s dictum: “Reality is that
which, when you stop believing in it, doesn’t go away.” The idea is that if
no one believed in Santa Claus, Santa Claus would not be on the scene, but
if no one believed in Joe Biden, he’d still be out there.
I don’t think Dick’s dictum is quite right as it stands. If we stop
believing in Gandalf but still see him on screen, he doesn’t go away. That
doesn’t mean that Gandalf is real. The same goes for illusions and
hallucinations: even if I believe that a distant mirage isn’t real, the mirage
doesn’t go away.
Still, Dick is onto something. In those cases, the presence of Gandalf or
the mirage still depends on people’s minds—their thoughts and experiences.
If no one had ever thought about Gandalf, he would never have entered our
lives. If we stop experiencing the mirage, it will go away. So we might say,
“Reality is that which, when it’s not in anyone’s mind, doesn’t go away.” Or
maybe better: “Reality is that which doesn’t depend on anyone’s mind for
its existence.”
Even this modified version of Dick’s dictum isn’t the whole truth about
reality. A counterpoint is what we might call Dumbledore’s dictum, spoken
by Hogwarts headmaster Albus Dumbledore toward the end of the Harry
Potter series: “Of course it is happening inside your head, Harry, but why
on earth should that mean that it is not real?” Your thoughts and experience
happen inside your head and depend on your mind, but they’re real all the
same. Social entities like money also depend on people’s minds. If no one
regarded dollar bills as valuable, they wouldn’t be money—but money is
real all the same.
Still, mind-independence can serve as a useful sufficient condition for
being real, and one that helps to explain at least one useful dimension of our
sense of reality. If something exists in a way that’s independent of anyone’s
mind, then it has an especially robust sort of reality. If something exists
only in a way that depends on our minds, then it’s less robustly part of the
external world.

Reality as non-illusoriness. What is the difference between illusion and


reality? So far, we’ve asked the question, “Do things really exist?” Another
crucial question, however, is “Are things as they seem?” We can use this
question as a fourth criterion for reality, capturing another strand in what we
mean when we say something is real. That is, something is real when it’s
the way it seems. Something is illusory when it’s not the way it seems.
Physical reality is real because it’s roughly the way it seems. When it
seems that there’s a ball in front of you in physical reality, there’s typically
a ball there. According to the standard view, virtual reality isn’t real because
it isn’t the way it seems. In VR it seems that there’s a ball there, but there’s
not. In virtual reality, the story goes, things are not as they seem—that is,
virtual reality is an illusion.
Right now, it seems to me that I’m sitting in a chair, using a desktop
computer, in a house somewhere in the Hudson Valley. It seems to me that
there’s an algae-covered pond outside the window, with geese wandering
around. It seems to me that I’m a philosopher working on a book on reality.
It seems to me that I grew up in Australia and that I now live in New York.
All this and more constitutes what we might think of as my apparent reality.
This apparent reality is real if, and only if, things are as they seem. If
I’m really sitting in a chair, and there is really a pond outside the window,
and I really grew up in Australia, then things are the way they seem, and all
this is real. If I’m not really sitting in a chair, and there is not really a pond
outside the window, and I didn’t really grow up in Australia, then all this is
not real.
Of course, I can be wrong about some things—maybe those aren’t really
geese outside the window—without my reality collapsing. But if I’m wrong
about almost everything, then it’s reasonable to say that my apparent reality
isn’t real.
What counts as the way things seem? There’s more than one candidate.
There’s the way we perceive things to be, using our senses. There’s also the
way we believe things to be, using our thinking and reasoning as well as
perception. I can perceive a pink elephant without believing that it exists. I
can believe that there was a Big Bang without having perceived it.
In these cases, it’s arguably what we believe that matters most for our
reality. For issues about skepticism and the simulation hypothesis, what
matters most to us is whether things in the world are as we believe them to
be. So where simulation realism is concerned, I’ll understand the fourth
criterion as saying things are real when they’re roughly as we believe them
to be.

Reality as genuineness. There’s a fifth way of thinking about reality that


comes from the British philosopher J. L. Austin in his 1947 lectures
collected as Sense and Sensibilia. (The title was inspired by Austin’s near-
namesake Jane Austen’s novel Sense and Sensibility, but the book is about
perception and reality.) Austin held that philosophy must always pay close
attention to the way words are used in ordinary language. To understand
what it is for something to be real, we must look at how normal English
speakers use the word “real.”
Austin said that in ordinary language we cannot just say of something
that it is real. If we did, it would invite the question “A real what?” The
issue would have to be whether it is a real diamond (as opposed to a fake
diamond), or a real duck (as opposed to a decoy duck). We might call this
Austin’s dictum: Instead of asking whether something is real, ask whether
it’s a real X.
I don’t think Austin’s dictum is quite right. It’s perfectly reasonable for
a child to ask whether Santa Claus is real or whether the Easter Bunny is
real. You could ask whether the Easter Bunny is a real rabbit, or a real
spirit, or whatever. But there’s also a question you can ask while staying
neutral on whether it’s a rabbit or a spirit: Is the Easter Bunny real? That is,
does it really exist? The answer seems to be no. It’s a folkloric figure that
has been passed down for generations; the folklore is real, but the bunny is
not. This is reflected in our first three dicta, all of which concern what it is
for an object to be real (rather than to be a real X). The Easter Bunny
doesn’t really exist, it doesn’t have causal powers, and it isn’t independent
of our minds.
Still, Austin’s dictum captures something crucial. We often don’t just
want to know whether something is real; we want to know whether it’s real
money or a real iPhone. If someone gives me an object that looks like a
Rolex watch, the object is indisputably real. It’s a real thing, at least. What
I’m interested in is whether it’s a real watch, and in particular whether it’s a
real Rolex. We could put this by asking whether the watch is genuine—that
is, is it a genuine watch? Is it a genuine Rolex?
The same goes for life in a simulation. It’s one question to ask whether
buildings and trees and animals we’re seeing are real; perhaps someone
could be convinced that they’re real digital entities. But it’s another
question to ask whether they’re real buildings, real trees, and real animals.
If they’re real objects but not real buildings, then things are not as they
seem. So when something seems to be an X, our fifth criterion for reality
asks: Is this genuine? That is, is it a real X?

Is simulated reality real?

We now have five criteria for reality. We can encapsulate them in a reality
checklist: five questions to ask when we’re wondering whether or not
something is real. Does it really exist? Does it have causal powers? Is it
independent of our minds? Is it as it seems? Is it a genuine X?
These five criteria—existence, causal powers, mind-independence, non-
illusoriness, and genuineness—capture five different strands in our concept
of being real. When we say that something is real, we sometimes mean one
of these things and sometimes a mix of them. There are other strands we
could have added—perhaps Reality as intersubjectivity, Reality as
theoretical utility, and Reality as fundamentality, among others I discuss in
an endnote—but these five are the strands that matter most for our
purposes.
None of these criteria for reality make it easy to figure out what’s real;
they just clarify the question a little. Different notions of reality are relevant
for different purposes, so we need to be clear about which notion we’re
using.
Let’s apply these criteria to the case of objects in a simulation. We can
start with the case of a perfect and permanent simulation—one that
simulates a world in full detail, which has generated all our experiences
throughout our history. If we’re in a perfect simulation, is our world real?
My purpose here isn’t so much to argue for my view about simulations as to
make clear what my view says.
The first criterion for reality asks: Do the objects we perceive in this
simulated world really exist? If I’m in a perfect simulation, does the tree
outside my window really exist? My opponent says, “No, the tree and the
window itself are mere hallucinations.” I say, “Yes, the tree and the window
really exist.” At some level they’re digital objects, grounded in digital
processes in a computer, but they’re no less real for that.
The second criterion asks: Do the objects we perceive have causal
powers? Do they make a difference? My opponent says “No, the tree
merely seems to make a difference.” I say, “Yes, the tree is a digital object
with many causal powers.” It produces leaves (which are themselves digital
objects), it supports (digital) birds, and it produces experiences in me and in
others who look at it.
The third criterion asks: Are the objects we perceive independent of the
mind? If minds go away, can they still exist? My opponent says, “No, the
tree I perceive exists only in my mind, and if we all went away, it wouldn’t
exist.” I say, “Yes. The tree is a digital object, which does not depend on me
for its existence.” Even if all (simulated and unsimulated) human life went
away, in principle the tree could continue to exist as a digital object.
The fourth criterion asks: Are things as they seem to be? Are flowers
really blooming in my garden, as they seem to be? Am I really a
philosopher from Australia, as I seem to be? My opponent says, “No, these
are mere illusions, and in reality there are no flowers and no Australia.” I
say, “Yes, there are really flowers blooming in the garden, and I am really
from Australia.” If my whole world is a simulation, flowers are ultimately
digital objects, and Australia is ultimately digital, too, but this is no obstacle
to the flowers blooming and to my being Australian.
The fifth criterion asks: Are the things I experience in a simulation real
flowers, and real books, and real people? My opponent says, “No. Even if
they’re real digital objects, these objects are at best fake flowers, not real
flowers.” I say, “Yes, these objects are real flowers, the books are real
books, and the people are real people.” If I’ve lived my whole life in a
simulation, every real flower I experience has been digital all along.
To sum up: if we’re in a perfect, permanent simulation, the objects we
perceive are real according to all five of those criteria.
I defined simulation realism this way: “If we’re in a perfect simulation,
the objects around us are real and not an illusion.” The reference to illusion
puts the greatest weight on the fourth criterion: Simulation realism holds
that things are largely as we believe them to be. Now, we believe that cats
exist, and that cats do things, and that they are real cats. Simulation realism
entails that in a simulation, these beliefs are largely true.
You might think that I’ve cherry-picked the criteria for being real.
Perhaps things would go differently if we defined reality as fundamentality
or originality. If we’re in a simulation, the simulated trees we perceive may
not be the original trees (simulators may have modeled them after
unsimulated trees) and the simulated physical world will not be
fundamental (since a simulation in the next universe up underlies it).
Perhaps these points capture some part of what people mean when they say
that virtual objects are not real. Still, these criteria seem marginal as criteria
for reality. Dolly the cloned lamb is not original (she’s a clone of another
lamb) and not fundamental (she’s made out of particles)—but she is
perfectly real.
To bring out what a strong view my simulation realism is, and to
reinforce that the criteria are not cherry-picked, we can contrast it with a
view laid out by the British theoretical physicist David Deutsch in his 1997
book The Fabric of Reality. In a fascinating discussion of VR, Deutsch
advocates a partial sort of virtual realism. He argues that VR environments
(including simulations) “pass the test for reality” because they “kick back”
at the user. That suggests versions of the second and third criteria: These
environments have causal powers and are independent of our minds. At the
same time, Deutsch does not endorse the first criterion: He says of one
scenario, “the simulated aircraft and its surroundings do not really exist.”
Nor does he endorse the fourth and fifth criteria. He says that in another VR
scenario you can see rain that’s not there in reality, and that an engine in the
scenario is not a real engine. In Deutsch’s view, VR environments are real
but the objects in them are illusions. Simulation realism makes a much
stronger claim than this.
Of course, if we’re in a perfect simulation, then some things aren’t
exactly as they seem. Some of what we believe will be wrong. Most people
believe that they’re not in a simulation. They believe that flowers are not
digital. They may believe that their universe is the ultimate reality. If we’re
in a simulation, those beliefs will be wrong. But the undermined beliefs
here are mostly scientific or philosophical beliefs about reality.
Undermining them does not undermine everyday beliefs such as There are
flowers blooming in the garden.
If we’re in an imperfect simulation, more of our beliefs will be wrong,
but plenty will still be right. If the solar system is fully simulated but the
rest of the universe is only a sketch, then my beliefs about the Sun may be
correct, but my beliefs about Alpha Centauri may be wrong. If 2019 is
simulated but 1789 was not, then my beliefs about the French Revolution
may be false, but my beliefs about the contemporary United States may be
correct. Still, we’ll be right about those closer-to-home aspects of the
universe that are present in the simulation. Assuming the deer I see in my
garden is part of the simulation, I’ll still be right that there’s a deer in my
garden.

The simulation hypothesis as metaphysics

I’m sure that some of these claims sound counterintuitive to you. One
source of resistance is that trees and flowers certainly don’t seem to be
digital objects. But they don’t seem to be quantum mechanical objects
either. Yet deep down they are, and few people think that the mere fact that
trees are grounded in quantum processes makes them less real. I think that
being digital is just like being quantum mechanical here.
Science has taught us that there’s much more to reality than initially
seems to be the case. For millennia, we didn’t know that cats and dogs and
trees are made of cells, let alone that the cells are made of atoms or that
those are fundamentally quantum mechanical. Yet these discoveries about
the nature of cats and dogs and trees have not undermined their reality.
If I’m right, the discovery that we’re in a simulation should be treated
the same way. It will be a discovery about the underlying nature of cats and
dogs and trees—that they spring from digital processes—but it won’t
undermine their reality.
Importantly, I’m not saying in an unqualified way that a simulated tree
is the same as a real tree. I’ve said that if we’re in a perfect permanent
simulation, then the real trees of our universe are simulated trees—that is,
real trees have been digital trees all along. On the other hand, if we’re not in
such a simulation and are merely looking at one from the outside, then the
trees in the simulation are completely different from the trees in our outside
world: Simulated trees are digital; real trees are not. The real trees of our
world have been nondigital entities all along. Either way, digital entities and
nondigital entities are quite different things.
I will argue that the simulation hypothesis should be seen as a variant of
the “it-from-bit” hypothesis that’s been widely discussed in physics. This
hypothesis posits a digital level underlying physics: roughly speaking,
molecules are made of atoms, atoms are made of quarks, and quarks are
made of bits. This is a view in which physical processes are real. There are
just underlying levels to reality that we don’t know about yet.
I think that’s the correct way to think about the simulation hypothesis. If
the simulation hypothesis is true, the it-from-bit hypothesis is true. Physical
reality is perfectly real—there’s just an underlying level consisting of the
interaction of bits, and perhaps still further levels still underlying that.
The it-from-bit hypothesis corresponds to one part of the simulation
hypothesis—the simulation itself. What about the other part—the simulator
who initiated the simulation? In the next chapter, I’ll argue that the
simulator is analogous to a god. At least, the simulator can be seen as the
creator of the it-from-bit universe. In chapter 9, I’ll argue that the
simulation hypothesis is itself a combination of the it-from-bit hypothesis
and the creation hypothesis (according to which a creator created the
universe).
If I’m right, the simulation hypothesis is not a skeptical hypothesis in
which nothing exists. Instead it is a metaphysical hypothesis, a hypothesis
about the nature of reality. It’s equivalent to a metaphysical hypothesis (the
creation hypothesis) about how our world was created, plus a separate
metaphysical hypothesis (the it-from-bit hypothesis) about what underlies
reality in our world. If the simulation hypothesis is right, the physical world
is made of bits, and a creator created the physical world by arranging those
bits.
In the next three chapters, I’ll lay out the creation hypothesis and the it-
from-bit-hypothesis and then argue that the simulation hypothesis is
roughly equivalent to the two hypotheses put together. Along the way, I’ll
discuss many connected issues about God and about reality.

The no-illusion view in the history of philosophy


My response to Cartesian skepticism rests on a positive answer to the
Reality Question. In a perfect simulation, things are perfectly real. The
same goes for other Cartesian scenarios, such as Descartes’s evil-demon
scenario and Hilary Putnam’s brain-in-a-vat scenario. Generalizing
simulation realism to these scenarios, we arrive at the no-illusion view of
Cartesian scenarios. In these scenarios, things are largely as they seem.
Subjects in these situations aren’t deceived; they have largely true beliefs
about the world.
If the no-illusion view is right, then Cartesian scenarios are no bar to
knowing about the external world. If most of our beliefs are correct in a
Cartesian scenario, then our inability to rule out that scenario does nothing
to cast doubt on our beliefs. Our beliefs are more robust than many
Cartesians have thought.
I’m not the first to advocate a response to skepticism holding that
subjects in these scenarios have largely true beliefs. Still, it’s striking how
uncommon this view has been in the history of philosophy. Recently, I
asked a number of historians of philosophy whether they knew of anyone
explicitly endorsing this view before the 20th century. No one could come
up with a clear case. Even in the 20th century, there were only a handful of
brief discussions of it.
I suspected that George Berkeley, the 18th-century Anglo-Irish
philosopher we first encountered in chapter 4, might have held this view.
Recall that Berkeley was an idealist who argued that appearance is reality.
In an evil-demon scenario, the tables and chairs have the same appearance
as tables and chairs in the physical world. If appearance is reality, it follows
that people in that scenario will experience real tables and chairs, and will
have mostly true beliefs about their reality.
Alas, it looks as if Berkeley never endorsed this view. He never
discusses Descartes’s evil-demon scenario at all. He would probably have
considered the evil demon an impossibility; like Descartes, he thought God
alone was perfect enough to produce sensations and perceptions like ours.
Still, he insisted that God is doing this and when he’s doing it, we’re getting
things right. You could think of that as a cousin of my view on skeptical
scenarios, with God playing the role of the simulator or evil demon.
The first really clear statement of the no-illusion view that I know of
appears in a 1949 essay by the University of Nebraska philosopher Oets
Kolk Bouwsma. Bouwsma was a student of Ludwig Wittgenstein, who
inspired Jonathan Harrison’s hallucinating Baby Ludwig in chapter 4.
Bouwsma’s article “Descartes’ Evil Genius” is a wonderful fable about a
hapless evil demon who sets out to deceive us about reality and always
fails. (Bouwsma translates “malignus genium” as “evil genius,” but I’ll use
the more common “evil demon,” as in chapter 3.) The demon first turns
everything into paper, but people quickly notice. Tom, the human
protagonist of the story, who has himself been turned into paper, detects the
illusion. He finds that paper flowers don’t smell or feel the same, and the
evil demon is exposed.

Figure 17 Bouwsma’s evil demon tries to deceive Tom, first with paper flowers and then with
perfect simulations.

In his second attempt, the evil demon destroys everything except


people’s minds. He grows cocky and whispers to Tom, “Your flowers are
nothing but illusions.” Tom replies that he can tell the difference between
flowers and illusions, and these are plainly not illusions. He understands
what the evil demon has done, but he says that what the demon calls an
illusion, he calls a flower. The demon has not created an illusion, he has
created flowers. The demon rides off on a corpuscle, defeated.
Like me, Bouwsma thinks that the subject in Descartes’s evil-demon
scenario is not undergoing an illusion. His reasons differ from mine,
however. Bouwsma thinks that an illusion is an illusion only if it can be
discovered. If an illusion cannot be discovered, as in Descartes’s evil-
demon scenario, it’s not an illusion at all, and it’s not a deception.
More precisely, Bouwsma’s evil demon can discover the illusion, but
his subject, Tom, cannot. So when the demon says, “This is an illusion,”
he’s correct, but if Tom were to say this, Tom would be wrong. For similar
reasons, what Tom experiences counts as a “flower” for Tom but not for the
demon. Consequently, when Tom says, “That’s a flower,” he’s correct.
Bouwsma’s reasoning here is grounded in a sort of verificationism,
which says that all meaningful claims are verifiable. You’ll recall that
earlier we discussed the verificationism of Rudolf Carnap and other logical
positivists, who held that skeptical hypotheses are unverifiable and
therefore meaningless. Bouwsma takes a related line. He thinks that the
hypothesis that there’s an illusion is meaningless if it cannot be verified. If
we can never verify that the evil-demon scenario is an illusion, it’s not an
illusion at all.
For reasons I discussed in chapter 4, I reject verificationism. Many
meaningful claims are unverifiable, and there may well be illusions we
never discover. So I reject Bouwsma’s analysis of the situation.
Nevertheless, I think he’s right about the crucial point. The subject in
Descartes’s evil-demon scenario is not undergoing an illusion.
A related route to the no-illusion view that shares some of the spirit of
Berkeley’s idealism is suggested by the Chinese philosopher Philip Zhai
(who now publishes as Zhenming Zhai) in his important 1998 book Get
Real: A Philosophical Adventure in Virtual Reality. Zhai holds that as long
as we have stable and coherent perception of an object, that object is real. In
the spirit of Berkeley, we might say that stable and coherent appearance is
reality. Zhai does not apply this thesis to issues about skepticism, but he
applies it to virtual reality and perfect simulations. A subject in a perfect
simulation will have stable and coherent perception of the world, so by
Zhai’s lights the world is real and not an illusion. I reject Zhai’s idealist
framework for much the same reasons that I rejected Berkeley’s idealism in
chapter 4, but like verificationism, idealism provides one important route to
the no-illusion view.
A third route to the no-illusion view is suggested by Hilary Putnam in
his 1981 book Reason, Truth and History. In chapter 4, we encountered
Putnam’s argument that the brain-in-a-vat hypothesis is contradictory based
on externalism, his theory that the meaning of a word depends on its
external context. Putnam suggests that the beliefs of a brain in a vat are
about the electronic impulses in its environment, and that these beliefs are
mostly true. This is a version of the no-illusion view that can potentially
drive an entirely different response to skepticism (though Putnam doesn’t
make the connection to skepticism). Instead of arguing that the brain-in-a-
vat hypothesis is contradictory, one could argue that the beliefs of a brain in
a vat are mostly true. In effect, where Bouwsma argues from
verificationism to the no-illusion view, Putnam argues from externalism to
the no-illusion view.
I’ll analyze Putnam’s argument for the no-illusion view in chapter 20,
where I discuss externalism. In my view, the argument doesn’t quite
succeed. Just as Bouwsma’s argument requires an implausible
verificationism and Zhai’s view requires an implausible idealism, Putnam’s
argument requires a strong and implausible version of externalism. Still,
like Bouwsma and Zhai, Putnam is on the right track in holding the no-
illusion view.
Bouwsma, Putnam, and Zhai offer three different routes to the no-
illusion view of simulation scenarios. All three routes can lead to what I
think is the correct view of the Cartesian skeptical problem: the subject in
Cartesian scenarios has largely true beliefs and is not deceived. However,
all three routes rest on strong and implausible philosophical views that I
reject. As a result, I think that none of these three philosophers has given a
strong argument for the no-illusion view nor a plausible analysis of just why
it is true. So we’re still in search of a good argument and a plausible
analysis.
I think the best argument for the no-illusion view comes not from
verificationism, idealism, or externalism but from a sort of structuralism
about the external world. I’ll work up to this argument gradually in the next
three chapters.

You might also like