Something Deeply Hidden | Sean Carroll | Talks at Google

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments

"... why in the world does the world need another book on quantum mechanics? And I think that the answer is that I don't like any of the other books. Especially because what they tend to do is emphasize how difficult it is to make sense of quantum mechanics. How surprising and spooky and mysterious it is." Ω©(^α΄—^)ΫΆ

πŸ‘οΈŽ︎ 33 πŸ‘€οΈŽ︎ u/josephwb πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

I love his podcasts. Thanks to JRE having him on, lots of other people do too.

πŸ‘οΈŽ︎ 43 πŸ‘€οΈŽ︎ u/FlatRateForms πŸ“…οΈŽ︎ Oct 05 2019 πŸ—«︎ replies

His podcasts are a work of art. A somnant bliss through and through. Can't have enough of it, ever.

πŸ‘οΈŽ︎ 11 πŸ‘€οΈŽ︎ u/KvellingKevin πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

Thanks for sharing. The way he explained this was understandable.

πŸ‘οΈŽ︎ 4 πŸ‘€οΈŽ︎ u/zalf14 πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

Sean Carroll's blog and podcasts are dope

πŸ‘οΈŽ︎ 5 πŸ‘€οΈŽ︎ u/Woah_Mad_Frollick πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

There's one argument he also talks about for an Everettian view that has always been the most convincing argument about interpretations of QM to me, actually pretty much converted me to an Everttian - and I have wondered why it seems inconclusive to many. Perhaps I'm missing some essential flaw and somebody could help me understand better - or perhaps it really is approximately as good of an argument as I think, in which case why not reiterate it.

In QM, starting from a system in a some prepared state for some observables, its evolution will be described by a wavefunction (SchrΓΆdinger or Dirac). The possible choice of different bases for decomposition of states in the time evolution of systems and the superposition principle leads to a unitarity, but not uniqueness of our solution for the question what can be observed at a later time, with the specific probabilities given by the Born rule.

When we look at (prepare) two such systems to interact in a relevant way somewhere along the line, the most interesting consequence (I think) of QM happens - there ceases to be a way to describe the evolution of the state(s) of one of those prepared systems irrespective of the other, instead its state itself becomes relative to the system it interacted with in entangled superpositions (Decoherence notwithstanding).

If we forget for a minute about the Copenhagen View that was (likely) the first introduction to all those ideas for all of us - and ask ourselves what the consequence is when we turn the above conceptualization around on ourselves and look at observer and observed system - we can see that this is a paradigmatic example of systems in relative states. And thus we arrive at the Everettian Relative State conceptualization.

Of course the reason for inventing the idea of wavefunction collapse in the first place is the same question that we, at this point in our thoughts about relative states, still have to answer: How do we "bridge the gap" between the unique, determinate things we observe, and the wavefunction, or the distributions and superpositions we get when interpreting a wavefunction in terms of determinate states.

But we have to realize that this is (while supremely important) a separate question, independent of the logical conclusion that if QM describes interacting systems as evolving in relative states, and if we as observers have no reason to exclude ourselves from being such systems in such interactions with the things we observe, then it follows that the observer-observed relationship is also one of systems evolving in relative states (ie as a structured whole).

Everett's point was that everything else (like Collapse, or Pilot Waves or other hidden variables) are additional theoretical elements not motivated from within the theoretical framework itself, but auxiliary hypotheses to make it jive (and here's the thing) not with "what we observe" simpliciter, but with a relatively specific ontological conception of "we" and "observe".

Everett's proposal was to take the theory at face value first - and questioning parts of our general ontological assumptions before making such additions to be consistent with other, specific parts of our ontologies.

From here on, we can go the Everett-DeWitt way and just assume that the theory is in fact complete, there is no "missing link" - which in turn means it's our perspective that's limited. Our observations correspond to having a specific preferred basis for decomposition of the overall state. The link to the probability distributions in observations is then given by the Born rule which functions as a measure of weight of the number of worlds in which a certain value will be observed relative to the weight for the other possibilities (a set-theoretic measure of relative magnitude) - while decoherence of many of the branching futures from a specific state explains that our observations are mostly of "ordinary" things and events, not the more "absurd" possibilities of quantum mechanical probability distributions - reducing the enormous Hilbert space via einselection to the things we actually regularly observe.

Furthermore, one might extend the investigation from allowed states to allowed state-transitions and how that can be synthesized with the insights about relative states.

An additional benefit is that this theory retains usable notions of physical objects with unique states - and places the uncertainty again firmly in the epistemic, not ontic camp, thereby providing more coherence, consistency and parsimony of a scale-integrated view of "what there is", physically than views which thought it was necessary to abandon those concepts to jive with experimental data.

That is, I think - the main tragedy of the fact that Copenhagen was victorious. Several generations of physicists have been educated with an understanding that we by necessity have to throw overboard our very conceptions of what objects and properties are, that even fundamental logic has to be abandoned (the law of excluded middle: "x can not have property B and a property which amounts to not-B"), leaving us with a necessary, radical disconnect between the ontology of our theories and the world we actually experience, and a radical disconnect to the ontologies of theories at different (meso or macroscopic) levels of size. ... often enough, the result of being educated this way is a conviction that attempts to not go that way and salvage conceptions of objects and properties are invalid because such inconsistencies are thought to be irrelevant when the maths works.

Thankfully, more and more physicists are realizing that this is not true - it is not necessary to abandon those concepts to formulate an empirically adequate theory of quantum mechanics. Neither locality nor realism have to be abandoned - when we thought that, we tacitly assumed counter-factual definiteness. But it turns out, the former two can be salvaged for the price of the latter. And this is anything but irrelevant. Empirical adequacy is one of several criteria for the explanatory value and epistemic probability of a theory - but infinitely many empirically adequate theories can be constructed for any set of observations. To adjudicate, we have to look to non-empirical measures of explanatory value and epistemic probability - namely how adopting a theory or hypothesis affects coherence, consistency and parsimony of the overall network of hypotheses/theories/beliefs relative to adopting a rival hypothesis or theory.

That, of course - does not mean that Many Worlds has to be true - but it seems to me that the value of overall coherence, consistency and parsimony is often underestimated, and that in any case - the Everettian insight that it is in fact not necessary to postulate either hidden variables or a mysterious collapse of the wavefunction, and that QM-observer and observed are in relative states just like other systems evolving in entangled superposition appears to remain valid, with the question being open where we best go from there. I personally, find an Everett-DeWitt approach modified with decoherence and research into potential restrictions of state-transitions and the consequences for Many Worlds very appealing - but am aware it has its issues and will always gladly seek out good arguments for alternative views.

πŸ‘οΈŽ︎ 3 πŸ‘€οΈŽ︎ u/BlueHatScience πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

Idk, I've been to his talk at CERN and it seems like a general agreement between my colleagues that he's not really doing science. Mathematically his approach does not differ from regular quantum mechanics, and there is no new testable prediction. If it wasn't for his curriculum he would be considered a crackpot.

πŸ‘οΈŽ︎ 6 πŸ‘€οΈŽ︎ u/migasalfra πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

Has he published any papers related to his research program of inferring gravitation from QM principles?

πŸ‘οΈŽ︎ 2 πŸ‘€οΈŽ︎ u/AlexCoventry πŸ“…οΈŽ︎ Oct 06 2019 πŸ—«︎ replies

Damn. For a sec I misread this as Steve Carell. Sean Carroll is great too! But the former would have been much funnier...

πŸ‘οΈŽ︎ 1 πŸ‘€οΈŽ︎ u/telescopes_and_tacos πŸ“…οΈŽ︎ Oct 31 2019 πŸ—«︎ replies
Captions
[MUSIC PLAYING] [APPLAUSE] SEAN CARROLL: Thanks. It's great to be here at Google. I've used your service on the internet. [LAUGHTER] And yeah, I want to talk about quantum mechanics. And I'm presuming that-- I know there are some experts in the audience. I'm presuming that there are many complete non-experts. But probably even the most complete non-experts have heard the phrase "quantum mechanics" and know that there are many books out there on quantum mechanics. And why in the world does the world need another book on quantum mechanics? And I think that the answer is that I don't like any of the other books, especially because, what they tend to do is to emphasize how difficult it is to make sense of quantum mechanics, how surprising and spooky and mysterious it is. And I admit that there are things about quantum mechanics that are hard to wrap our minds around. But I don't think there's anything intrinsically unintelligible about it. So the message of my book is that you can understand quantum mechanics, not that you can't. Of course, I am flying in the face of my esteemed predecessor at Caltech, Richard Feynman, who thinks, "I think I can safely say that nobody understands quantum mechanics." And I know what he means. You know, he's not wrong. We-- quantum mechanics, for those of you who don't know, is this wonderfully successful theory that we developed mostly over the first quarter of the 20th century that's supposed to apply to everything in the universe, but really becomes manifest and necessary when we look at microscopic things, when we look at electrons or atoms or very, very small-scale things. And we can use this theory to extraordinary precision. We can make predictions that have been tested to 12 decimal places of accuracy. Quantum mechanics is absolutely necessary to understand why the sun shines or how transistors work or why this table is solid, OK? So why does Feynman say this? Because even though we can use quantum mechanics to make predictions, if you ask physicists, but what's really happening, they don't know. And that's OK. It's OK not to know things, but when you don't know, what you should try to do is to learn, is to figure it out. And as a field, physics has decided not to do that. Rather than taking people who are trying to understand quantum mechanics at the deepest level and treat them as the superstars and most important people in the field, they push them out of the field. They've decided that trying to understand quantum mechanics at a deep level is not our job as physicists. It's only to make predictions. I think that this is bad. The analogy I like to use is with Aesop's fable of "The Fox and the Grapes." If you remember this one, the Fox sees the grapes, and would like to eat these juicy, wonderful grapes, and jumps up, but can't reach the grapes. The Fox is unable to get to the grapes. So the Fox says, you know what? I never really wanted those grapes anyway. They were probably sour. The Fox represents physicists, and the grapes represent understanding quantum mechanics. We used to try very hard to understand quantum mechanics, and we subsequently stopped, and I think that was a mistake. So let me tell you how we got to quantum mechanics. This is a fake version of the history, but it gives you an idea. So by 1909, we had this picture of the atom, which is the same cartoon you will always see when people talk about atoms. I think it's literally the logo of the Atomic Energy Commission or the Nuclear Regulatory Commission. There is a nucleus, a middle of the atom, which we now know is full of protons and neutrons, and there are electrons orbiting around the atom, OK? The problem is that, even though this is consistent with certain pieces of data, it can't be true as the final answer, at least, according to the standards of classical Newtonian mechanics because these electrons zooming around in orbits should be emitting electromagnetic waves. When you take a charged particle like an electron, and you shake it in any way, it has an electric field radiating out in all directions. So when you move the electron, the field adjusts. And when you move it up and down, the field waves, and we see those. That's where the light that we're seeing right now comes from, from vibrating electrons. So these electrons that are moving around in circles should be emitting light. And that means they should be losing energy, and they should spiral in to the nucleus of the atom. And you can calculate how quickly that should happen. It's about 10 to the minus 11 seconds, which is a very short period of time. So in other words, according to the rules of classical mechanics applied to the Rutherford model of the atom, all matter should be dramatically unstable. You and the chair you're sitting on and this table and the Earth itself should collapse into a point in a tiny fraction of a second. That prediction does not fit the data, so we need to do something better. There were many ideas batted around. What we eventually hit upon was, don't think of the electron as a point particle in an orbit. Think of the electron as a wave. In particular, we dubbed this the wave function, which is the most boring, uninspiring name for the most important thing in all of physics. The wave function is the idea that, rather than being in an orbit around the nucleus of the atom, think of the electron as being described by a wave that is spread out around the atom. And just like a violin string that you pluck in different ways-- a string that is tied down at both ends has, I guess, the fundamental frequency that it can vibrate at. And then there are overtones, harmonics, that it can vibrate at in different ways. The point is that there's a discrete set of different ways it can vibrate. Likewise, if you think of the electron as a wave, rather than as a particle, there's a discrete set of shapes that wave can have in the atom. And there's a minimum energy shape. So rather than just spiraling in to the center of the nucleus, the electron goes to its minimum energy configuration, which is still spread out. It's the shape in the upper left here, spherically symmetric. These other shapes are higher-energy. Versions of the wave function of the electron. And then it just sits there forever. And if you remember high school chemistry, you were tortured by pictures of electrons doing these things and the various orbitals that they could be in. And so this became a way of thinking about why matter is stable. Even better, a couple of years later, Erwin Schrodinger came up with an equation for how these orbitals, electron wave functions, actually behave. And I give you the details of the Schrodinger equation, because I know you all want to go off and solve it in your spare time. There are people here at Google who spend a lot of time solving this equation on quantum computers. And you don't need to know the details of the Schrodinger equation if you are not a quantum mechanical expert. What you need to know is that there is an equation, OK? So physicists love equations for good reasons. It's because you can go from the idea, the original idea, the electron is a wave around the atom, and now you can just apply it widely. You can apply it to any other circumstance. You can take the wave function doing anything. And you can let it evolve. You can solve this equation. In words, the Schrodinger equation says, any one wave can be decomposed into different energy parts. And every energy part evolves at a separate rate. So the energy of the wave is proportional to how fast it evolves. That's the Schrodinger equation. That's the entire thing. And so this equation still-- you know, suggested in 1926-- is still right, as far as anyone knows. It's one of the fundamental rules of nature. So you might think that we're, like, triumphant, that quantum mechanics is basically done, that we know that now electrons are not particles; they're waves. This is the equation we obey. That's what you want in a good physical theory, an understanding of what nature is and which equation it obeys. The problem is the data don't stop there. So this is a wonderful little image of a chunk of uranium in a cloud chamber, OK? So a cloud chamber has some pressurized gas in it that, when a charged particle moves through the gas, it ionizes the particles around it, and it creates a little track, OK? So if you have a chunk of uranium-- it's radioactive-- it will actually emit electrons and alpha particles and things like that. And you can use the Schrodinger equation to say, well, what does the wave function of the emitted electron look like? The answer is, it looks like a spherical wave. The electron wave function comes out in all directions more or less equally. There's details, but that's the basic story. But then you look at it. You never see an electron coming out in all directions equally. What you see in this picture are straight lines, trajectories, as if the electron is a particle again. What's up with that? We still have not answered this question, what's up with that? This is the sad part about the modern understanding of quantum mechanics. So what people said was, look, it looks like the way that electrons behave in their wave functions is different when you're not looking at them, like when they're in the nucleus, around the nucleus of an atom, versus when you are looking at them, like when they're in a bubble chamber or a cloud chamber, OK? Now, surely that can't be the final answer. But what the strategy adopted by physicist in the 1920s was is, yes, that's the final answer. So they invented an idea called the collapse of the wave function. They said, sure, wave functions for electrons or whatever, they obey the Schrodinger equation when you're not looking at them. But when you measure a quantum mechanical system, its wave function suddenly and unpredictably changes. And rather than predicting exactly what's going to be able to happen, you can predict the probability of different things happening. And the probability is largest where the wave function was largest. And after the measurement has occurred, the wave function changes to be localized where you observed it. So on the left here, you have a picture of a wave function that might be all spread out for an electron. You imagine measuring the position of the electron. Its wave function collapses somewhere. And if you looked at it again right away, you would see it in the same place. So this is the attempt to understand why electrons are all spread out in wave functions when you're not looking at them, but look like particles when you do look at them. Because the act of measurement is something special. It really does something to the particle in a particular way. So this is what's called the "textbook" or "Copenhagen" interpretation of quantum mechanics. This is what we teach our undergraduates. I'm not, like, going to reveal the truth, you know? This is what we actually tell our kids. This is what I was told. The rules of quantum mechanics, according to this way of thinking, come in two groups. There's one set of rules for when you're not observing the system. Those rules say that the electron is described by wave functions, or whatever quantum system, there's a wave function. And that wave function obeys the Schrodinger equation. And that's exactly in parallel with the rules of classical mechanics. In classical mechanics, there's a system, and it obeys some equations. And that's it. Those are all the rules. In quantum mechanics, you have those rules, but then there's extra rules for what happens when you measure or look at or observe the system. You can only measure certain things. When you do, the wave function collapses. And the best you can do is to predict the probability of a certain collapse. That's the Copenhagen interpretation. This is clearly unacceptable as a fundamental theory of nature. I mean, what are you talking about? It's fine for making predictions and building things, but it's clearly not the final answer. This was the point of contention in the famous Bohr-Einstein debates in the '20s and '30s. Niels Bohr was the founder and defender of the Copenhagen interpretation. Einstein said, look, that's fine as far as it goes, but clearly it has not gone far enough. We need to look harder. This is why the title of my book is "Something Deeply Hidden." There's something else going on that we haven't yet explicated, and Bohr and his friends said, no, no, no. It's fine. Don't worry about it. And Bohr and his friends totally won the public relations battle. But I think that Einstein was actually right in this particular dispute. So let me just mention two reasons why the Copenhagen interpretation by itself can't be the right final answer. There's what we might call the reality problem. We started by saying that electrons should be thought of as wave functions, but what is the wave function, really? Is the wave function a complete description of nature? In other words, is there an isomorphism between the mathematical formalism of a wave function and what reality is doing? Or is it part of nature, but not the whole part? Maybe there's other variables there. That's what Einstein himself thought. He thought that there was a wave function, but there were also particles. So the wave function described what happened when you weren't looking at them, but when you measure the thing, you see the actual particle. And in the modern version, these are called hidden variable theories. Bohmian mechanics is the most famous version. But maybe the wave function doesn't represent reality in any way. There's other people who say, just use the wave function to make predictions. Don't think of it as part of reality. So the fact is that, even professional physicists who use quantum mechanics everyday, can't tell you what it says about reality. And if you ask them, they will tell you in all honesty, bless their hearts, that they don't care about reality. All they care about is making predictions for observations. I think that's a bad attitude to have, but I'm in the minority there. The other problem is called the measurement problem. And this one is even more obvious as a problem. What do you mean, make a measurement or look at something? Like, what counts as a measurement? Who is able to do measurements? Is it just people? Do you need to be human? Do you need to be conscious? What about when you're asleep? Can cats make measurements? What about video cameras? What if I don't look at it very closely? What if I glance at it? Does that count as a measurement? How fast does it happen? When does it happen? A well-defined, rigorous theory of physics should answer all of these questions. And the Copenhagen interpretation doesn't answer any of them. It just says, don't worry about that. You know a measurement when you see it. I think that's not quite good enough to be a fundamental theory of nature. So there are different alternatives. These days-- this was all put together in the 1920s, right? These days, we actually have perfectly well-defined theoretical physics frameworks that do answer these questions. The problem is, we have more than one, and we are having trouble deciding between them. So I am not going to be, in any sense, fair in this talk. I'm just going to tell you my favorite one and let you know there are others. Buy the book if you want to see the others. OK. Here is my favorite. It came from Hugh Everett, who was a graduate student and was incredibly criticized for his ideas and that left the field right away. And so Hugh Everett says the following. He says, here is the solution to all of your quantum problems. Chill out. Stop working so hard. You're trying too hard with all this stuff. You've made up a whole separate set of rules about quantum mechanics for what happens when you measure or observe a system. Just erase all those rules. Just forget about them. You want to know what reality is? It's the wave function. Do you want to know what wave functions do? They obey the Schrodinger equation, and that's it. That's all you need to know. So in other words, this is the Everett interpretation of quantum mechanics. There's only one set of rules. They're always obeyed. And all it says is, there are wave functions obeying the Schrodinger equation. Now, there's a problem with this interpretation, which is why it was not invented by Bohr or Einstein in the 1930s, in that, it doesn't seem to quite map onto our experience, right? Like, we saw the cloud chamber with all the little particles moving. They don't look like they're obeying Schrodinger's equation, so what's going on? And Everett says, there are two things that you've forgotten about that are already built into quantum mechanics. He doesn't add anything, right? He erased some rules. He says, there are two things that are already there in quantum mechanics that, if you just take them seriously, everything will be fine. One thing is, you, yourself are a quantum system, right? You are part of the wave function of the universe. You are not a classical system. In Copenhagen, depending on who is talking and what time of day it was, they would admit that they treated the observer as classical. There's even something called the Heisenberg cut, advocated by Werner Heisenberg, that separates the classical, macroscopic world of observers from the microscopic quantum world. And Everett says, look, you might be big and made of a lot of atoms, but the atoms you're made of obey the rules of quantum mechanics, so you do, too. Treat yourself as quantum mechanical. And the other point that he wanted to emphasize was entanglement-- again, a feature of quantum mechanics, not anything that he added to the theory. So let me mention what entanglement says. My favorite example is the Higgs boson, because I wrote a book about that, OK? The Higgs boson, we discovered back in 2012, and the reason it's my favorite example is because the Higgs boson is a spinless particle. Every elementary particle comes with a quantity called spin, which is really just like the spin of the Earth or a spinning top or anything like that. Except, when you get down to that microscopic realm, spin comes in definite, quantized values. So the spin of the Higgs boson is exactly 0. The spin of an electron is 1/2. Which means that, if you observe the spin of an electron in quantum mechanics, it will either be spinning clockwise by amount 1/2 or spinning counterclockwise by. An amount 1/2. And we call these spin up and spin down. So we know from the observations and from the theory, the Higgs boson can decay into two electrons, technically, an electron and an anti-electron, a positron, so the charge is conserved. But let's ignore that complication. So one particle with 0 spin is decaying into two particles with spin 1/2. But spin is conserved. Spin is angular momentum. It's part of that, right? So it must be that when the Higgs boson decays into an electron and an anti-electron, they're each spinning, but they'd better be spinning in opposite directions. The total spin has to add up to 0 because that's what it was for the Higgs boson all by itself. Now, we don't know which way it is, right? Like, one electron is spin down, and the other one is spin up. But we don't know which is up and which is down. So if you go through the Schrodinger equation, what you find is that it's not one or the other. It's a superposition of both. One of the crucial features of quantum mechanics is that you have different pieces of the universe. But unlike in classical mechanics, the different pieces don't have their own individual states that we can talk about. Rather, there's only one quantum state. There's only one wave function of the entire universe. So when you talk about what the two electrons are doing, you need to talk about both of them at once. So you know that electron one might be spin up or might be spin down. There's a 50/50 chance. You know that electron two might be spin up or spin down. There's a 50/50 chance. But you also know that they're not spinning the same way. So if you measure the spin of electron one and find that it's spinning up, you instantly know that electron two is spinning down. And this state of the Higgs boson decaying into something is that it decays into a superposition of spin up for electron one, spin down for electron two, plus spin down for electron one, spin up for electron two. That's entanglement. What is happening to particle one is entangled with what's happening to particle two. They're not independent from each other. This is so important, let me say exactly the same thing again in slightly different words to drive it home. There's only one wave function, the wave function of the universe, what Everett called the universal wave function. His first title for his thesis was "The Theory of the Universal Wave Function," but his advisor did not let him get away with that. So the lesson is, you might think that if you have two electrons, and you don't know the spin of either one, but you know they can be in superpositions, you might say, well, electron one is in a superposition of up and down. Electron two is in a superposition of up and down. Quantum mechanics says there's not two separate wave functions for the two electrons; there's only one, and they are entangled in the following way. And this was actually invented or at least appreciated by Einstein. Einstein and Schrodinger wrote letters back and forth. Because despite playing fundamental roles in the origin of quantum mechanics, neither Einstein nor Schrodinger liked how quantum mechanics was going. They didn't like the Copenhagen interpretation. And so the origin of entanglement was in Einstein's effort to show that quantum mechanics couldn't be complete as yet. Because you can send one of these particles light years away from the other one. And if you measure the spin of one, you instantly know the spin of the other. And Einstein says, look, I'm Einstein. I know about this thing called relativity. You can't send information faster than the speed of light. How does it know way over there? And that's the spooky action at a distance that he invented. But it's there. It's true. It's part of quantum mechanics, and you can actually test it experimentally. So let me show you, then, how entanglement explains the measurement problem according to the Everett interpretation. And we're going to use Schrodinger's cat to explain this. You've probably heard of the thought experiment put forward by Erwin Schrodinger. He has a box. He puts the cat in the box, closes the box. And there's a little apparatus there that has a quantum wave function that has a probability of opening some vial and emitting gas into the box or not doing that. The gas Schrodinger decided to contemplate was cyanide. So the cat's either alive or dead. His daughter literally said later, I think my father just didn't like cats. [LAUGHTER] There's no reason to kill the cat in the thought experiment. It plays no role. So in my version of the thought experiment, it's sleeping gas. AUDIENCE: Aww. SEAN CARROLL: And what happens is, there's a quantum system that evolves into a superposition of having decayed or not. So there is a Geiger counter or some other detector that evolves into a superposition of having clicked or not. So the box and the gas evolve into a superposition of having been emitted or not. So the cat evolves into a superposition of being awake or being asleep. The whole mess is just designed to take a superposition of a tiny, microscopic quantum system and amplify it to a macroscopic world. It doesn't matter that it's a cat. Honestly, the Geiger counter or the box by itself would have worked perfectly well. But remember, Schrodinger was unhappy about quantum mechanics. The point of the Schrodinger's cat experiment was not to go, like, look how weird quantum mechanics is. Schrodinger said, what you're telling me, Mr. Bohr, is that when I open the box, before I opened the box, the cat was in a superposition of awake or asleep. And after I open the box, it suddenly snaps into either awake or asleep. Surely, you don't believe that, right? That was his point. And so in other words, if you knew about classical mechanics, it could be perfectly plausible to say, there's a cat in the box. It might be awake. It might be asleep, but I don't know. But the only two possible states of the cat are, it's awake or it's asleep. In quantum mechanics, we have a new possibility of this superposition, that the cat can be in a combination of awake and asleep. And then we tell a story about what happens when we open the box and measure that superposition. And that story is very different in Copenhagen versus in Everett. So in Copenhagen, we treat the observer as classical. So I'm putting classical things in square brackets and quantum things in parentheses. So the role of the observer here is played by Niels Bohr. And we start with a system where the cat is quantum and in a superposition. And Niels Bohr is classical, and he has not yet opened the box. He opens the box, and that's observation or measurement. And according to the Copenhagen interpretation, the wave function collapses. And after that measurement, either the cat was awake and the observer measured it to be awake, or the cat was asleep and the observer saw it sleeping. Those are the two possibilities, and there's a certain probability you can calculate. So Everett says, ignore all of this business about wave functions collapsing. Just obey the Schrodinger equation. Just ask what the Schrodinger equation would predict for this particular setup. So there's no collapse, and there's only one wave function, and the observer is quantum in their own right. So this is Hugh Everett now playing the role of the observer. So measurement in Everett's version of quantum mechanics is just obeying the laws of physics. It's just, open the box and let Schrodinger's equation do its work. And what happens inevitably-- and everyone agrees on this-- is that the wave function of the universe evolves to a superposition of the cat was awake, and the observer saw the cat awake; and the cat was asleep, and the observer saw the cat asleep. Again, if you obey the Schrodinger equation, that's what happens. No ifs, ands, or buts. The problem is, it looks like you, if you're the observer, would be in a superposition of having seen two different things. And nobody in the history of human beings has ever felt like they are in a superposition of having seen one thing or having seen another. So the immediate rejection of Everett's theory was because it doesn't fit the data. It doesn't explain our experience. It doesn't explain why we see tracks that look like particles in the cloud chamber. So miraculously, Everett got it right, despite the fact that, in some sense, he had no right to. The real explanation for this comes from decoherence, which is a process which really wasn't understood until the 1970s. Remember, I said that the point of quantum mechanics is there's only one wave function for the whole universe, not separate wave functions for separate pieces. And then I put into the wave function the cat and the observer. I really should put in the entire rest of the universe, right, to be strictly correct. So here's the entire rest of the universe. It's usually, in this context, called the environment, so I made a picture of grass. But really, what you should think of is the light coming from these light bulbs or the air in the room, like, all the stuff, all the degrees of freedom that we're not explicitly keeping track of, that are bumping into us all the time, but we don't know where every single individual photon is. So long before we open the box, in the box, there are photons. There are air molecules, et cetera. They will interact with the cat. And they will interact with the cat differently depending on whether the cat's awake or asleep. Because the cat is in different locations in the box, right? So a photon will hit it or will not hit it depending on whether the cat is awake or asleep. So the environment becomes entangled with the cat long before you open the box. And then you finally open the box, and you say what you're doing is called measurement. But really you're just becoming entangled with the two different quantum states that were always there. Now, why does it matter? It looks a lot like this equation at the bottom is very similar to the equation we had before. What matters is that these two different states of the environment-- well, let me put it this way. The environment is doing two different, separate things in the two parts of the wave function, the part where the cat's awake and the part where the cat's asleep. The technical term is that these two environment states are orthogonal to each other. You can show mathematically this will happen very, very quickly. And what that means is that the separate two terms in this equation evolve separately. They obey their own equations of motion. If something happens in one part of that sum of two terms, it does not affect anything happening in the other term. So Everett says, this is the prediction of the Schrodinger equation, and it's right. Believe it. This is the final answer. What you haven't realized is, you're asking, why don't I feel like I am in a superposition? The answer is, because there are now two of you. There is one of you that saw the cat awake and one of you that saw the cat asleep. It's as if these two different parts of the superposition describe different worlds. So the crucial part here is that Everett doesn't put the worlds in. The Everett interpretation-- I'm not really revealing a surprise here. It's sometimes called the many-worlds interpretation of quantum mechanics. He never called it that. The name was not invented until 1970 by Bryce DeWitt. The point is, if you just have the Schrodinger equation and you just follow what it does. The worlds appear, like it or not. They were always there. Everett just points out that they naturally come to be. The process of decoherence separates the wave function into distinct branches that no longer interact with each other, so they are, for all intents and purposes, separate worlds. Now, there's a bunch of questions you can ask about the many-worlds interpretation. Some of them are easier to answer. Some of them are hard. Many of them are hard, to be honest. But I'm not going to dwell on all of them. I just want to give you a flavor for how we answer them. So one question is, how many worlds are there? We don't know is the short answer. We don't even know whether the number of worlds is finite or infinite, OK? The straightforward and simple answer is that there are infinitely many worlds. This is what Everett himself would have believed. At the technical level, for those of you who do know a little bit about quantum mechanics, the question is, what is the dimensionality of Hilbert space, the space of all the possible wave functions? For simple systems like an electron in nonrelativistic quantum mechanics or quantum field theory, Hilbert space is almost always infinite-dimensional. And therefore, for all intents and purposes, there are an infinite number of worlds. But we don't know that for sure. And people like me think that quantum gravity implies that the Hilbert space is actually finite-dimensional. So maybe there's only a finite number of worlds. But even if it's a finite number, it's a really, really big number. So rather than thinking of the splitting, the branching of the wave function, as happening at special events where you work hard to make it happen, it's happening all the time. There are radioactive decays in your body roughly 5,000 times a second, and every one of those duplicates the universe, OK? So rather than thinking of, like, bang, a special event that splits the universe, there's a constant whooshing as the universe is subdivided. And that's the way to think about it. It's not a duplication of the world with twice as much energy and everything. There's a certain amount of world-ness in the equations, and it's being subdivided and differentiated as time goes on. Think of the world as splitting, not as being copied in some way. Now, there are some objections that I think I can easily answer. One is that there's just too many universes-- sorry, I don't like it. This is not a very scientifically respectable worry. But I get it. The problem is that there's enough stuff going on in one universe. Doesn't it seem, like, ontologically extravagant to add all this extra stuff in? The response to this is that once you believe quantum mechanics, the potential for all these worlds was always there. Once you believe that electrons can be in superpositions, you should believe that people can be in superpositions and worlds can be in superposititons. Hilbert space is the name we attach to the space of all possible wave functions. It's not any bigger in many-worlds than it is in Copenhagen or anyone else's interpretation of quantum mechanics. It's just that many-worlds lets the wave function be wherever it wants to be in Hilbert space. So this might be an objection to quantum mechanics, but it's not an objection to many-worlds, per se. The second question is, how can you test this theory? You're saying there's all these other worlds out there that I can't interact with. Doesn't that violate the spirit of science, that I should be able to test my predictions? Well, there's a longer conversation here, but let's appeal to Karl Popper, the philosopher who said that a good scientific theory should be falsifiable. Of course, every scientific theory makes some predictions that you can't test. The question is, are there any experiments you could imagine doing that could give you answers that would cause you to reject the theory? And remember, all Everett is is the statement, there are wave functions, and they obey the Schrodinger equation. That's the world. So this is the most falsifiable theory ever invented. All you have to do is either find variables other than the wave function, or find the wave function doing something other than obeying the Schrodinger equation. And there are ongoing experiments to do exactly these things. Karl Popper himself was a huge fan of the Everett interpretation. He thought that the Copenhagen interpretation was a philosophical monstrosity. So in terms of testability, Everett is just as good as anyone else's interpretation of quantum mechanics. Now, so those are, I think, the easily answered objections. There are two other questions that are harder. I'm going to just do, like, one minute on one of them and then a couple of minutes on the other one. The first hard question is, where does probability come from? So as an empirical matter, when we do measurements on quantum systems, it is certainly true that the best we can do is to predict the probability of certain measurement outcomes. And in Copenhagen, that's because we put in a postulate into the theory that says there's a probability rule, OK? Everett-- all the postulates are, there's a wave function. It obeys the Schrodinger equation. There's no mention of probability. In fact, the Schrodinger equation is 100% deterministic. If I know the wave function at one moment of time, I can predict it at any other moment. There's nothing probabilistic about the dynamics. So how in the world does probability enter what we know about quantum mechanics from experience? The answer is this idea of self-locating uncertainty. You can indeed know everything there is to know about the wave function, and there's still something you don't know, which is where you are within the wave function. So let's go back to the cat and the observer after decoherence happened but before the observer knows what branch of the wave function they're on. Decoherence happens really, really fast. Typical timescales are less than 10 to the minus 20 seconds in a big, macroscopic system that you're not trying to protect. So before the observer opens the box, decoherence has happened, and the wave function has branched. So there's already two branches of the wave function, the cat awake branch and the cat asleep branch. But because the observer hasn't looked, there are two copies of the observer that are exactly identical with each other. Because the observer has not opened the box, if you ask the observer, which branch are you on, cat awake or cat asleep, they don't know. They are identical. So even though they know the entire wave function of the universe, they don't know which branch they're on. That is self-locating uncertainty. And you can ask, given some reasonable assumptions, is there a uniquely rational way to assign credences, probabilities, to being on one branch or the other? And you go through the math, and the answer is yes, and guess what? It's exactly the same as the postulate that you had in the Copenhagen interpretation-- the probability is given by the wave function squared. So it is very natural for probability to arise even though the underlying dynamics are completely deterministic. Now, the harder question is, how do we relate Everettian quantum mechanics to the world of-- the classical world we see of tables and chairs? And this is a tough one for me to get across because most of my fellow physicists don't even think this is a problem. I think it's the hardest problem, and we should all be thinking about it. But we all grow up tending to think about the world classically. So we look around. We see there are cats. There are trees. There are people. Those are the starting point when we try to describe the world. So when even professional physicists make quantum mechanical models of reality, of, spins or materials or particles, they start with some classical description, and then they quantize it. So you start with some classical stuff, and then there's rules that you're taught in graduate school or undergraduate education for turning that into a good quantum theory. And at the end of the day, you have a wave function, which mathematically is a vector living in Hilbert space, living in this giant dimensional vector space. Presumably that's cheating. Presumably nature doesn't start with a classical theory and quantize it. Nature has no need for that. Nature just is quantum mechanical from the start, right? So in some sense, you should be starting with a wave function thought of as a vector in some abstract Hilbert space and deriving the rest of the world. This is how nature actually works. That turns out to be really, really hard. And it's harder in many-worlds than it is in other approaches to quantum mechanics. In other approaches to quantum mechanics, whether it's Copenhagen or hidden variables or whatever, they sneak in some features of the classical world into the definition of the theory. Despite the fact that there are many worlds, when it comes to assumptions, when it comes to pieces of the theory itself, Everett is the leanest and meanest. Wave functions in the Schrodinger equation, and that's it. Every other approach to quantum mechanics adds extra stuff, and that extra stuff gives you a handle on why the world looks classical in a certain way. In Everett, you have to work harder. "Why does the world look classical at all?" is a perfectly good question. So keep that question in mind. Why does the world of classical at all? There's another question that we have, which is quantum gravity, right? I'm sure that you might have heard, depending on which street corners you hang out on, that we don't yet have a good quantum theory of gravity itself. We can take this procedure I mentioned over here, and it works for every force of nature and every piece of matter in nature other than gravity. For electromagnetism, the nuclear forces, particles that we know about, electrons and quarks and so forth, you start with a classical theory, you quantize it, you get a reasonable answer. For gravity, the best classical theory we have is Einstein's general theory of relativity. And his point is that gravity is a feature of spacetime itself. Namely, it is the curvature of spacetime. Spacetime curves and warps and responds to matter and energy, and we experience that curvature as the force of gravity. So he has a perfectly good classical theory. We can put it into our black box, and we can quantize it, and we get a terrible mess. It doesn't work. We have not yet successfully taken classical general relativity and applied the usual rules of quantization in a successful way. That's why people are driven to consider alternatives, like string theory or loop quantum gravity, et cetera. Those have made some progress, but none of them is obviously the right answer in any sense. So let me suggest the following thing-- that maybe these two problems cancel each other out. The problem being, we don't understand quantum mechanics, and we don't understand quantum gravity. Well, why in the world should we understand quantum gravity if we don't even understand quantum mechanics, right? Maybe this fact that, in Everett, you should start from a quantum wave function and emerge out of it the classical world, is really the right way to think about gravity. In other words, don't quantize gravity in the sense of starting with some classical model and turning it into a quantum theory. Start with some quantum theory and finding gravity within it. Well, we're allowed to take clues from the world as we do know it, where quantum field theory is the best way we have of thinking about particles and forces, right? If it weren't for gravity, quantum field theory is the way nature works. So here's a picture of a field. This is the magnetic field, OK? If you have a magnet. There's an invisible force field around it. We know that because you put the magnet near your refrigerator, and it reaches out and grabs it. Likewise, there's a gravitational field, an electric field, et cetera. In modern physics, everything is fields. So even particles, even matter, like electrons and quarks and neutrinos-- there are fields that fill all of space. And what you perceive as a particle is really a vibration in that field. So the fundamental ingredients of nature, according to modern physics, are fields filling space vibrating in certain ways, in a quantum mechanical framework, the whole thing we call a quantum field theory. And there's an immediate consequence of that, which is that empty space is fun. It's full and lively, and it's a busy place, OK? If the world were made of particles, you can imagine there are particles, and the space in between them is just space. It's just sitting there, and nothing is happening. If the world is made of fields, fields fill all of space. And even where you say space is empty, and there's nothing there, what you really mean is that there are fields there, but they're in their lowest-energy state. They're in their vacuum state. And we can take space and subdivide it into little regions and talk about what the fields are doing in every region. What you mean by a particle is that the field is vibrating more than it does in its vacuum state. But in most of space, it's just sitting there quietly. And there's a feature that you can find in quantum field theory that different regions of space, the fields that are vibrating there, are entangled with each other. And the amount of entanglement is related to the distance between them, to the geometry of space itself. If two regions of space are nearby, they will be highly entangled. If they're far away, they're not very entangled. That is the conventional story of quantum field theory. Now, let's ask ourselves, can we use this as a clue to find gravity, to emerge space itself? The point is, we can't put space in. We have to find it in the wave function. So the idea, the suggestion is, we can turn this idea around, the relationship between geometry and entanglement. Rather than saying, when two parts of space are close together they're highly entangled, we say, when two quantum parts of Hilbert space are highly entangled, they are close together. That's what it means to be close together. And if you do that for all the different parts of the wave function, a geometry emerges on space. And you cross your fingers. If everything works out nicely, it's the kind of three-dimensional space that you and I know and love and live in. So we can easily imagine a relationship between geometry and entanglement. There is also a relationship between entanglement and energy. Remember, I said that, in the vacuum, in empty space, there's a very specific entanglement structure in quantum field theory. If you want to put a particle somewhere-- so you want to add some energy to a region of space by putting a particle there-- you basically have to break the entanglement of that little vibrating quantum field with the rest of its surroundings. So you decrease the amount of entanglement in a region, and you necessarily add energy and vice versa. There's a direct correlation between the amount of entanglement somewhere and what we would traditionally refer to as the energy. So look what we got. By referring to entanglement rather than begging the question, by assuming space and stuff like that, we find that there is a relationship between the geometry of emergent space and the amount of entanglement. There is separately a relationship between the energy and the amount of entanglement. Therefore, there is a relationship between the amount of geometry-- the geometry of space and the amount of energy in one region. Modulo some mathematical details that I encourage you to check out. But that's general relativity. That's what Einstein said. He said that the geometry of space is sourced by the amount of stuff, energy and matter and so forth. What I'm saying here is that, if you didn't know that there was any such thing as space and you treated space as something that emerged naturally from the quantum mechanical entanglement between different parts of the wave function of the universe, the natural consequence would be that space has a geometry, and it's curved and that geometry obeys an equation that is very similar to Einstein's equation of general relativity. Now, we haven't proven this. Truth in advertising. This is an ongoing research program that is not done yet. We've made progress, but the progress is the following. Rather than saying, this is true, we have a long list of assumptions. And if all those assumptions are correct, then this is true, OK? So it's full employment for graduate students in the future generations. They need to prove every one of our assumptions is true. But it's very encouraging to me to imagine that the problem with quantum gravity was just that we shouldn't be quantizing gravity at all. We should be taking quantum mechanics seriously and finding it. So I know that there's probably some details there that I went through too quickly. So happily, you can buy my book, which I think is free to you. I think you're giving away copies for free. But all the details are in there. The details that are not in there are referenced in there. It's an exciting time. I think that we're actually making progress for the first time in understanding how nature works. And taking reality seriously turns out to be a smart strategy for theoretical physicists. Thank you very much. [APPLAUSE] And we have time for questions. SPEAKER 1: Yes, we have some time. So does anyone have a question? Hi. Look alive. SEAN CARROLL: That seems very dangerous. AUDIENCE: (LAUGHING) It's very soft. SEAN CARROLL: We live on the edge here. AUDIENCE: OK. So many-worlds theory is the fodder for a lot of science fiction. Just off the top of your head, where does modern science fiction get it right, and where does it get it wrong? SEAN CARROLL: Yeah. So that's a very good question. Modern science fiction almost always gets it wrong for the following very good reason-- that once the worlds branch, according to the real, true theory of Everett, they can't talk to each other anymore. So you can download-- if you have an iPhone, there is an app called Universe Splitter, where it will send a signal to a beamsplitter that will send a photon either left or right. And if you agree ahead of time, if the photons go left, I'm going to have Chinese food for dinner; if the photon goes right, I'm going to have pizza for dinner, there will be a universe in which you had Chinese food for dinner and a universe in which you had pizza. But you can no longer talk to the other one and say, well, how was yours, right? So that does not make for very good drama when that happens. Now, there are-- maybe the Schrodinger equation isn't quite right. Steven Weinberg, the famous physicist, suggested a modification of the Schrodinger equation. And Joe Polchinski, another famous physicist pointed out that, if that modification were true, it would enable communications between different branches of the wave function, which he called an Everett phone. So that's a perfectly respectable thing to write a TV show or a movie about if that's what you want to do. [LAUGHTER] It's crossed my mind, yeah. AUDIENCE: If the, uh-- so if there's just a single wave function and the universe is just a ton of forking events that are creating all these different universes, I guess, is the implication then that there was ever or is ever just a single point or instance of entropy or, like, true probability or randomness that occurred? SEAN CARROLL: Yeah, well, I think the way that I would put it is, there is absolutely a direction of time in this picture, right? There were fewer separate branches of the wave function in the past than there are in the future. Now, there's separately an arrow of time given by statistical mechanics, given by entropy increasing. Entropy used to be low, and it's growing. I think it's the same thing. I think both cases are just facts that the initial conditions of our observable universe, 14 billion years ago near the big bang, were highly non-generic, were very special and very low-entropy. And nobody knows why in either case. But whatever caused the wave function in the universe to have few branches and the physical thermodynamic entropy of the universe to be low explains everything since then. AUDIENCE: So my question is related to this one. So if the branching needs to happen constantly, does that mean some sort of serialization needs to be done? Like-- SEAN CARROLL: Some what? AUDIENCE: Serialization. Like events need to wait for each other before they can branch the wave. Like, how do you explain that? SEAN CARROLL: Yeah, so I'm going to give you the correct answer, and it's going to be unsatisfying, OK? Let me make it concrete by thinking about the Higgs bosons and decaying and the EPR experiment where Einstein said, look, I'm going to take particle one and particle two, separate them by many light years, and observe particle one. And you're telling me, instantly, particle two changes. You can prove a theorem that says that cannot be used for communication. You cannot send information that way, because the observer of particle two doesn't know what the result was from particle one. But following my philosophy that we want to understand what's really happening, you're still allowed to ask what's really happening, OK? So the answer is, the wave function of the universe branches. What you want to know is, should I think of branching when I observe particle one as happening simultaneously throughout the whole universe, or am I allowed to think of it as sort of spreading within the light cone of that particular point? The answer is, either way. The answer is that the whole idea of branches of the wave function are convenient human constructs. Everett's idea is that physics is just the wave function of the whole universe. And this is what it means to say that the classical world, the branches of the wave function, are emergent. They are approximate descriptions that give us a handle on what's happening in a convenient way that is based on a lot less information than just giving us the whole wave function of the universe. So you can slice up the wave function of the universe by letting it branch instantaneously throughout space. Or you can keep things moving slower than the speed of light. Both descriptions give exactly the same predictions for every conceivable experiment. It's up-- whatever makes you feel better. [LAUGHTER] I told you that it would be unsatisfying. AUDIENCE: That's not satisfying. You were right. AUDIENCE: OK. So I was curious what you might think are some of the most, I guess, telling criticisms of this view of the universe. SEAN CARROLL: Yeah. I think I would say that there are two big looming-- I don't even want to say problems-- but research questions that have not been completely answered, both of which I talked about, one of which is the probability question. I gave you my favorite answer, but there's definitely perfectly legitimate worries and very smart people who worry about that. The best worry is the following simple-minded one, which I think is simple-minded, but it's hard to answer, which is, look, OK, fine. You say that there's self-locating uncertainty. And one of these branches, I don't know which one, and you say there's a good way to assign probabilities to being on one or the other. But what forces me to assign them that way? Like, why can't I just say I don't know? Like, what's wrong with saying, like, I don't know. I'm on some branch. Why, in my exper-- how do I-- am being forced to map this particular mathematical fact about the wave function onto my experience of flipping coins or measuring spins. That's a perfectly good objection. And the other one is this structure question. Like I said, we, made some bold assertions and assumptions and derived curved spacetime. It's only in the weak field limit. We don't yet have things like black holes. And there's still a bunch of questions that are just tricky. Like, one thing that we assume is that the Hamiltonian, the fundamental energy function of the universe, takes a very specific form, which seems to be experimentally true. But if you want to be really broad-minded, you say, well, why? Why does the-- why do the laws of physics look like that? Why is there locality at all? Why is there space at all? And there's no answer there in Everett. There's no answer there is anyone else's theory, either, by the way. But the ability to answer this question is more obvious in Everett, and we haven't answered it yet. AUDIENCE: If you assume a finite-dimensional Hilbert space for the universe, is there a notion of-- and you're sort of branching all the time-- is there a notion of running out a room and some of the branches coming back and hitting each other? SEAN CARROLL: There absolutely is. So if there's a finite number of branches possible, and branching is happening all the time, math theorems tell us we're going to run out of branches. Happily, we're nowhere close to doing that. And what it would look like-- it's just the approach to thermal equilibrium within any one branch. What happens is all the branches become indistinguishable from each other because equilibrium looks the same no matter where you started, right? So in some very real sense, once that happens, it's not so much that branches re-fuse together as there's no difference between the different branches. There's no obvious way to divide the universe into branches at all. AUDIENCE: Why can you say with such confidence that we're not close to that? SEAN CARROLL: Because it doesn't look like thermal equilibrium. I mean, happily. There are still stars shining in the sky. That's why. Anything else? I can sign books if we have any minutes left over, but yeah. [LAUGHTER] AUDIENCE: This is kind of a different style of question, but-- so I think quantum mechanics has a reputation as being a very arcane thing, discipline. SEAN CARROLL: Yeah. AUDIENCE: And yet, I think the kind of thinking that is fit for theoretical physics is probably pretty widespread and common. And I think this mainly because I'm a software engineer and because I find myself at home reading and listening to theoretical physics books and things. You know, bad theories smell bad in the same way that bad software smells bad. So if this is really the most important problem in science, arguably, how do we have a sea change that allows, regular old people who like thinking this way to do something about it and isn't just, like, increasing the number of crackpots in the world? SEAN CARROLL: Yeah, I mean, this is a good question. And it would be easy to joke about it, but it's hard to give the right answer. Here's the good news. Of course, quantum mechanics involves math-- complex analysis, linear algebra, and so forth. But it's not that hard math. And furthermore, these kinds of questions, these foundations of physics questions, the math is not the obstacle. It's really like-- it's undergraduate math. And you can get, as a professional theoretical physicist, into field theory your string theory or whatever, and the math gets harder and harder. But none of that's really relevant here, honestly. So I think that, if you're an honest autodidactic who wants to study quantum mechanics from a textbook and learn enough of it to think deeply about the foundations of physics questions, it's very doable. I mean, one of the best-- two of the best books on foundations of physics are David Albert's book "Quantum Mechanics and Experience" and David Wallace's book "The Emergent Multiverse." And they're both pretty accessible to people who like matrices and calculus. So I would say, get those, read them, take them seriously, and see what you have to do. AUDIENCE: I was just going to make an observation that I enjoy a podcast that talks about a lot of these things called "Mindscape." So people-- SEAN CARROLL: Yes, I have a podcast. Somehow I didn't mention that. Thank you. Thank you. Your $5 is in the mail. [LAUGHTER] Subscribe to my podcast. I talk sometimes about quantum mechanics. But other times I talk about music or movies or whatever we want to talk about that day. So it's a fun scape of the mind. Thank you very much for having me. [APPLAUSE]
Info
Channel: Talks at Google
Views: 431,323
Rating: 4.847743 out of 5
Keywords: talks at google, ted talks, inspirational talks, educational talks, something deeply hidden, sean carroll, science rules, sean carroll mindscape, sean carroll interview
Id: F6FR08VylO4
Channel Id: undefined
Length: 57min 3sec (3423 seconds)
Published: Fri Oct 04 2019
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.