Search Results for: hawking

Hawking on black holes

The cat is out of the bag about Stephen Hawking’s new ideas about black holes. Preposterous regulars were in on the ground floor, of course. We can now speculate somewhat more intelligently about what the ideas actually are, since the abstract for his talk (next week at the GR17 conference in Dublin) is online:

The Euclidean path integral over all topologically trivial metrics can be done by time slicing and so is unitary when analytically continued to the Lorentzian. On the other hand, the path integral over all topologically non-trivial metrics is asymptotically independent of the initial state. Thus the total path integral is unitary and information is not lost in the formation and evaporation of black holes. The way the information gets out seems to be that a true event horizon never forms, just an apparent horizon.

(For more scoop see Not Even Wrong, Smijer, By the Way, and this thread on sci.physics.research.)

It seems as if our initial skepticism might be accurate: even if Hawking’s proposal is ultimately judged to be a fantastic breakthrough, it won’t catch on right away. The Euclidean path integral approach to quantum gravity, which Hawking has been instrumental in developing, is not generally thought to be the most promising approach. One never knows, and it might eventually turn out to be on the right track, but it’s not very popular in the quantum-gravity community.

Let’s try to explain what is going on. Quantum mechanics tells us that the world is described by wave functions, which give the probability for getting a certain result when we make an observation. A path integral is just a certain way of calculating the wave function. Invented by Richard Feynman, the path-integral approach says that we should attach a certain contribution to every possible path the system can take from one configuration to another, and then sum all of these contributions; the result is the wavefunction for being in the final state, given the initial state we started with. Often this sum (the path integral) is hard to do, and a convenient mathematical trick is to allow the time coordinate to be imaginary; this purely formal manouver turns four-dimensional spacetime into a four-dimensional space (no direction is singled out as “timelike”), and the integral turns out to be much easier to do. In gravity, we would like to sum over all the possible geometries of spacetime itself. So “Euclidean quantum gravity” tells us to calculate the wave function for the geometry of three-dimensional space at some fixed time by summing over all the possible four-dimensional geometries that connect to that spatial geometry.

The problems with this approach include: nobody knows what number to attach to each geometry in the sum, nobody knows how to do the integral, and nobody knows how to interpret the result. Significant problems, in other words. Optimists like Hawking think that they can describe a certain set of approximations in which the path integral makes sense; string theorists, on the other hand, would generally say that the path integral is nonsense if you don’t include the fundamentally “stringy” aspects of spacetime on ultra-small scales. In the absence of direct experimental data, we have to judge how well the ideas hang together in their own right; at the moment, Euclidean quantum gravity seems to have serious issues, and hasn’t scored many notable triumphs.

Hawking now seems to be saying that the path integral for black holes can help us understand how information that we thought had disappeared into the singularity can actually be lurking close to the horizon, so that it can eventually influence the outgoing Hawking radiation, and thus allow information to be recovered by the evaporating black hole. He’ll be met with serious skepticism in the physics community, but let’s wait to see how it washes out.

Of course, he’ll also be met with serious awe by the media. This can drive other physicists crazy. I often find myself explaining to non-physicists that Hawking is not the very best theoretical physicist out there, but simultaneously defending his reputation to my fellow physicists. It’s probably safe to say that Hawking was the leading researcher in theoretical gravitational physics in the second half of the twentieth century, and arguably since Einstein; but gravitational physics was not where the action was during that time (it was in particle physics and quantum field theory). A lot of physicists hurry to point out that Hawking is not doing the most influential work in quantum gravity these days; but the truth is that most sixty-year-old physicists aren’t doing the most influential work in their fields, even if they don’t have serious neurological disorders. Like anyone else who has made fantastic contributions, Hawking has earned respect for new ideas he might have, but the ideas themselves will be judged on their own merits.

Hawking on black holes Read More »

Thanksgiving

This year we give thanks for one of the very few clues we have to the quantum nature of spacetime: black hole entropy. (We’ve previously given thanks for the Standard Model Lagrangian, Hubble’s Law, the Spin-Statistics Theorem, conservation of momentum, effective field theory, the error bar, gauge symmetry, Landauer’s Principle, the Fourier Transform, Riemannian Geometry, the speed of light, the Jarzynski equality, the moons of Jupiter, and space.)

Black holes are regions of spacetime where, according to the rules of Einstein’s theory of general relativity, the curvature of spacetime is so dramatic that light itself cannot escape. Physical objects (those that move at or more slowly than the speed of light) can pass through the “event horizon” that defines the boundary of the black hole, but they never escape back to the outside world. Black holes are therefore black — even light cannot escape — thus the name. At least that would be the story according to classical physics, of which general relativity is a part. Adding quantum ideas to the game changes things in important ways. But we have to be a bit vague — “adding quantum ideas to the game” rather than “considering the true quantum description of the system” — because physicists don’t yet have a fully satisfactory theory that includes both quantum mechanics and gravity.

The story goes that in the early 1970’s, James Bardeen, Brandon Carter, and Stephen Hawking pointed out an analogy between the behavior of black holes and the laws of good old thermodynamics. For example, the Second Law of Thermodynamics (“Entropy never decreases in closed systems”) was analogous to Hawking’s “area theorem”: in a collection of black holes, the total area of their event horizons never decreases over time. Jacob Bekenstein, who at the time was a graduate student working under John Wheeler at Princeton, proposed to take this analogy more seriously than the original authors had in mind. He suggested that the area of a black hole’s event horizon really is its entropy, or at least proportional to it.

This annoyed Hawking, who set out to prove Bekenstein wrong. After all, if black holes have entropy then they should also have a temperature, and objects with nonzero temperatures give off blackbody radiation, but we all know that black holes are black. But he ended up actually proving Bekenstein right; black holes do have entropy, and temperature, and they even give off radiation. We now refer to the entropy of a black hole as the “Bekenstein-Hawking entropy.” (It is just a useful coincidence that the two gentlemen’s initials, “BH,” can also stand for “black hole.”)

Consider a black hole whose area of its event horizon is A. Then its Bekenstein-Hawking entropy is

    \[S_\mathrm{BH} = \frac{c^3}{4G\hbar}A,\]

where c is the speed of light, G is Newton’s constant of gravitation, and \hbar is Planck’s constant of quantum mechanics. A simple formula, but already intriguing, as it seems to combine relativity (c), gravity (G), and quantum mechanics (\hbar) into a single expression. That’s a clue that whatever is going on here, it something to do with quantum gravity. And indeed, understanding black hole entropy and its implications has been a major focus among theoretical physicists for over four decades now, including the holographic principle, black-hole complementarity, the AdS/CFT correspondence, and the many investigations of the information-loss puzzle.

But there exists a prior puzzle: what is the black hole entropy, anyway? What physical quantity does it describe?

Entropy itself was invented as part of the development of thermodynamics is the mid-19th century, as a way to quantify the transformation of energy from a potentially useful form (like fuel, or a coiled spring) into useless heat, dissipated into the environment. It was what we might call a “phenomenological” notion, defined in terms of macroscopically observable quantities like heat and temperature, without any more fundamental basis in a microscopic theory. But more fundamental definitions came soon thereafter, once people like Maxwell and Boltzmann and Gibbs started to develop statistical mechanics, and showed that the laws of thermodynamics could be derived from more basic ideas of atoms and molecules.

Hawking’s derivation of black hole entropy was in the phenomenological vein. He showed that black holes give off radiation at a certain temperature, and then used the standard thermodynamic relations between entropy, energy, and temperature to derive his entropy formula. But this leaves us without any definite idea of what the entropy actually represents.

One of the reasons why entropy is thought of as a confusing concept is because there is more than one notion that goes under the same name. To dramatically over-simplify the situation, let’s consider three different ways of relating entropy to microscopic physics, named after three famous physicists:

  • Boltzmann entropy says that we take a system with many small parts, and divide all the possible states of that system into “macrostates,” so that two “microstates” are in the same macrostate if they are macroscopically indistinguishable to us. Then the entropy is just (the logarithm of) the number of microstates in whatever macrostate the system is in.
  • Gibbs entropy is a measure of our lack of knowledge. We imagine that we describe the system in terms of a probability distribution of what microscopic states it might be in. High entropy is when that distribution is very spread-out, and low entropy is when it is highly peaked around some particular state.
  • von Neumann entropy is a purely quantum-mechanical notion. Given some quantum system, the von Neumann entropy measures how much entanglement there is between that system and the rest of the world.

These seem like very different things, but there are formulas that relate them to each other in the appropriate circumstances. The common feature is that we imagine a system has a lot of microscopic “degrees of freedom” (jargon for “things that can happen”), which can be in one of a large number of states, but we are describing it in some kind of macroscopic coarse-grained way, rather than knowing what its exact state actually is. The Boltzmann and Gibbs entropies worry people because they seem to be subjective, requiring either some seemingly arbitrary carving of state space into macrostates, or an explicit reference to our personal state of knowledge. The von Neumann entropy is at least an objective fact about the system. You can relate it to the others by analogizing the wave function of a system to a classical microstate. Because of entanglement, a quantum subsystem generally cannot be described by a single wave function; the von Neumann entropy measures (roughly) how many different quantum must be involved to account for its entanglement with the outside world.

So which, if any, of these is the black hole entropy? To be honest, we’re not sure. Most of us think the black hole entropy is a kind of von Neumann entropy, but the details aren’t settled.

One clue we have is that the black hole entropy is proportional to the area of the event horizon. For a while this was thought of as a big, surprising thing, since for something like a box of gas, the entropy is proportional to its total volume, not the area of its boundary. But people gradually caught on that there was never any reason to think of black holes like boxes of gas. In quantum field theory, regions of space have a nonzero von Neumann entropy even in empty space, because modes of quantum fields inside the region are entangled with those outside. The good news is that this entropy is (often, approximately) proportional to the area of the region, for the simple reason that field modes near one side of the boundary are highly entangled with modes just on the other side, and not very entangled with modes far away. So maybe the black hole entropy is just like the entanglement entropy of a region of empty space?

Would that it were so easy. Two things stand in the way. First, Bekenstein noticed another important feature of black holes: not only do they have entropy, but they have the most entropy that you can fit into a region of a fixed size (the Bekenstein bound). That’s very different from the entanglement entropy of a region of empty space in quantum field theory, where it is easy to imagine increasing the entropy by creating extra entanglement between degrees of freedom deep in the interior and those far away. So we’re back to being puzzled about why the black hole entropy is proportional to the area of the event horizon, if it’s the most entropy a region can have. That’s the kind of reasoning that leads to the holographic principle, which imagines that we can think of all the degrees of freedom inside the black hole as “really” living on the boundary, rather than being uniformly distributed inside. (There is a classical manifestation of this philosophy in the membrane paradigm for black hole astrophysics.)

The second obstacle to simply interpreting black hole entropy as entanglement entropy of quantum fields is the simple fact that it’s a finite number. While the quantum-field-theory entanglement entropy is proportional to the area of the boundary of a region, the constant of proportionality is infinity, because there are an infinite number of quantum field modes. So why isn’t the entropy of a black hole equal to infinity? Maybe we should think of the black hole entropy as measuring the amount of entanglement over and above that of the vacuum (called the Casini entropy). Maybe, but then if we remember Bekenstein’s argument that black holes have the most entropy we can attribute to a region, all that infinite amount of entropy that we are ignoring is literally inaccessible to us. It might as well not be there at all. It’s that kind of reasoning that leads some of us to bite the bullet and suggest that the number of quantum degrees of freedom in spacetime is actually a finite number, rather than the infinite number that would naively be implied by conventional non-gravitational quantum field theory.

So — mysteries remain! But it’s not as if we haven’t learned anything. The very fact that black holes have entropy of some kind implies that we can think of them as collections of microscopic degrees of freedom of some sort. (In string theory, in certain special circumstances, you can even identify what those degrees of freedom are.) That’s an enormous change from the way we would think about them in classical (non-quantum) general relativity. Black holes are supposed to be completely featureless (they “have no hair,” another idea of Bekenstein’s), with nothing going on inside them once they’ve formed and settled down. Quantum mechanics is telling us otherwise. We haven’t fully absorbed the implications, but this is surely a clue about the ultimate quantum nature of spacetime itself. Such clues are hard to come by, so for that we should be thankful.

Thanksgiving Read More »

23 Comments

Quantum Is Calling

Hollywood celebrities are, in many important ways, different from the rest of us. But we are united by one crucial similarity: we are all fascinated by quantum mechanics.

This was demonstrated to great effect last year, when Paul Rudd and some of his friends starred with Stephen Hawking in the video Anyone Can Quantum, a very funny vignette put together by Spiros Michalakis and others at Caltech’s Institute for Quantum Information and Matter (and directed by Alex Winter, who was Bill in Bill & Ted’s Excellent Adventure). You might remember Spiros from our adventures emerging space from quantum mechanics, but when he’s not working as a mathematical physicist he’s brought incredible energy to Caltech’s outreach programs.

Now the team is back again with a new video, this one titled Quantum is Calling. This one stars the amazing Zoe Saldana, with an appearance by John Cho and the voices of Simon Pegg and Keanu Reeves, and of course Stephen Hawking once again. (One thing about Caltech: we do not mess around with our celebrity cameos.)

Stephen Hawking + Zoe Saldana: Quantum is Calling ft. Keanu Reeves, Simon Pegg, John Cho, Paul Rudd

If you’re interested in the behind-the-scenes story, Zoe and Spiros and others give it to you here:

Behind the Scenes: Stephen Hawking + Zoe Saldana: Quantum is Calling

If on the other hand you want all the quantum-mechanical jokes explained, that’s where I come in:

The Science Behind Quantum Is Calling

Jokes should never be explained, of course. But quantum mechanics always should be, so this time we made an exception.

Quantum Is Calling Read More »

5 Comments

Science Career Stories

The Story Collider is a wonderful institution with a simple mission: getting scientists to share stories with a broad audience. Literal, old-fashioned storytelling: standing up in front of a group of people and spinning a tale, typically with a scientific slant but always about real human life. It was founded in 2010 by Ben Lillie and Brian Wecht; I got to know Ben way back when he was a postdoc at Argonne and the University of Chicago, before he switched from academia to the less well-trodden paths of communication and the wrangling of non-profit organizations.

By now the Story Collider has accumulated quite a large number of great tales from scientists young and old, and I encourage you to catch a live show or crawl through their archives. I was able to participate in one about a year ago, where I shared the stage with a number of fascinating scientific storytellers. One of them was one of my mentors and favorite physicists, Alan Guth. Of course he has an advantage at this game in comparison to most other scientists, as he gets to tell the story of how he came up with one of the most influential ideas in modern cosmology: the inflationary universe.

It’s a great story, both for the science and for the personal aspect: Alan was near the end of his third postdoc at the time, and his academic prospects were far from clear. You just need that one brilliant idea to pop up at the right time.

But everyone’s path is different. Here, from a different event, is my young Caltech colleague Chiara Mingarelli, who explains how she ended up studying gravitational waves at the center of the universe.

Finally, it is my blog, so here is the story I told. I basically talked about myself, but I used my (occasionally humorous) interactions with Stephen Hawking as a hook. Never be afraid to hitch a ride on the coattails of someone immensely more successful, I always say.

Science Career Stories Read More »

8 Comments

Einstein’s Girl

One of the joys of living in LA is being surrounded by talented and creative people of all stripes, both inside and outside the world of science. Recently, for example, my friend Gia Mora helped me redesign my main website. This is a long-overdue upgrade; the site has existed in one form or another since 1996, and has been hacked together in html by me. Remarkably, things have changed over the last twenty years, and what sufficed in the waning years of the twentieth century had become unwieldy and unable to cope with the challenges of modern webbery. Now I have a working site that should be ready, for example, for the publication of The Big Picture.

Einsteins_logo But all of that is just an excuse for mentioning that web design is not Gia’s primary calling — it’s acting and singing, often with a bit of scientific flair. For anyone in the LA area, I encourage you to check out her show Einstein’s Girl, appearing at the Malibu Playhouse on Saturday Feb. 27. I’ve seen the show, and it’s great, plus I’m told there will be highly topical gravitational-wave jokes added in for just this occasion.

Gia’s singing about physics recently took center stage at Caltech, for the Institute of Quantum Information and Matter’s One Entangled Evening event. Besides the famous Rudd/Hawking quantum chess video, and maybe a few science talks, the highlight of the program was John Preskill, resplendent in black tie, joining Gia to sining a duet about quantum entanglement. John is not technically a professional singer like Gia is, but he did write the lyrics, and I can testify that he’s very good at quantum mechanics. Hats off to both the performers.

Musical Performance - One Entangled Evening

Einstein’s Girl Read More »

10 Comments

Guest Post: Grant Remmen on Entropic Gravity

Grant Remmen“Understanding quantum gravity” is on every physicist’s short list of Big Issues we would all like to know more about. If there’s been any lesson from last half-century of serious work on this problem, it’s that the answer is likely to be something more subtle than just “take classical general relativity and quantize it.” Quantum gravity doesn’t seem to be an ordinary quantum field theory.

In that context, it makes sense to take many different approaches and see what shakes out. Alongside old stand-bys such as string theory and loop quantum gravity, there are less head-on approaches that try to understand how quantum gravity can really be so weird, without proposing a specific and complete model of what it might be.

Grant Remmen, a graduate student here at Caltech, has been working with me recently on one such approach, dubbed entropic gravity. We just submitted a paper entitled “What Is the Entropy in Entropic Gravity?” Grant was kind enough to write up this guest blog post to explain what we’re talking about.

Meanwhile, if you’re near Pasadena, Grant and his brother Cole have written a musical, Boldly Go!, which will be performed at Caltech in a few weeks. You won’t want to miss it!


One of the most exciting developments in theoretical physics in the past few years is the growing understanding of the connections between gravity, thermodynamics, and quantum entanglement. Famously, a complete quantum mechanical theory of gravitation is difficult to construct. However, one of the aspects that we are now coming to understand about quantum gravity is that in the final theory, gravitation and even spacetime itself will be closely related to, and maybe even emergent from, the mysterious quantum mechanical property known as entanglement.

This all started several decades ago, when Hawking and others realized that black holes behave with many of the same aspects as garden-variety thermodynamic systems, including temperature, entropy, etc. Most importantly, the black hole’s entropy is equal to its area [divided by (4 times Newton’s constant)]. Attempts to understand the origin of black hole entropy, along with key developments in string theory, led to the formulation of the holographic principle – see, for example, the celebrated AdS/CFT correspondence – in which quantum gravitational physics in some spacetime is found to be completely described by some special non-gravitational physics on the boundary of the spacetime. In a nutshell, one gets a gravitational universe as a “hologram” of a non-gravitational universe.

If gravity can emerge from, or be equivalent to, a set of physical laws without gravity, then something special about that non-gravitational physics has to make it happen. Physicists have now found that that special something is quantum entanglement: the special correlations among quantum mechanical particles that defies classical description. As a result, physicists are very interested in how to get the dynamics describing how spacetime is shaped and moves – Einstein’s equation of general relativity – from various properties of entanglement. In particular, it’s been suggested that the equations of gravity can be shown to come from some notion of entropy. As our universe is quantum mechanical, we should think about the entanglement entropy, a measure of the degree of correlation of quantum subsystems, which for thermal states matches the familiar thermodynamic notion of entropy.

The general idea is as follows: Inspired by black hole thermodynamics, suppose that there’s some more general notion, in which you choose some region of spacetime, compute its area, and find that when its area changes this is associated with a change in entropy. (I’ve been vague here as to what is meant by a “change” in the area and what system we’re computing the area of – this will be clarified soon!) Next, you somehow relate the entropy to an energy (e.g., using thermodynamic relations). Finally, you write the change in area in terms of a change in the spacetime curvature, using differential geometry. Putting all the pieces together, you get a relation between an energy and the curvature of spacetime, which if everything goes well, gives you nothing more or less than Einstein’s equation! This program can be broadly described as entropic gravity and the idea has appeared in numerous forms. With the plethora of entropic gravity theories out there, we realized that there was a need to investigate what categories they fall into and whether their assumptions are justified – this is what we’ve done in our recent work.

In particular, there are two types of theories in which gravity is related to (entanglement) entropy, which we’ve called holographic gravity and thermodynamic gravity in our paper. The difference between the two is in what system you’re considering, how you define the area, and what you mean by a change in that area.

In holographic gravity, you consider a region and define the area as that of its boundary, then consider various alternate configurations and histories of the matter in that region to see how the area would be different. Recent work in AdS/CFT, in which Einstein’s equation at linear order is equivalent to something called the “entanglement first law”, falls into the holographic gravity category. This idea has been extended to apply outside of AdS/CFT by Jacobson (2015). Crucially, Jacobson’s idea is to apply holographic mathematical technology to arbitrary quantum field theories in the bulk of spacetime (rather than specializing to conformal field theories – special physical models – on the boundary as in AdS/CFT) and thereby derive Einstein’s equation. However, in this work, Jacobson needed to make various assumptions about the entanglement structure of quantum field theories. In our paper, we showed how to justify many of those assumptions, applying recent results derived in quantum field theory (for experts, the form of the modular Hamiltonian and vacuum-subtracted entanglement entropy on null surfaces for general quantum field theories). Thus, we are able to show that the holographic gravity approach actually seems to work!

On the other hand, thermodynamic gravity is of a different character. Though it appears in various forms in the literature, we focus on the famous work of Jacobson (1995). In thermodynamic gravity, you don’t consider changing the entire spacetime configuration. Instead, you imagine a bundle of light rays – a lightsheet – in a particular dynamical spacetime background. As the light rays travel along – as you move down the lightsheet – the rays can be focused by curvature of the spacetime. Now, if the bundle of light rays started with a particular cross-sectional area, you’ll find a different area later on. In thermodynamic gravity, this is the change in area that goes into the derivation of Einstein’s equation. Next, one assumes that this change in area is equivalent to an entropy – in the usual black hole way with a factor of 1/(4 times Newton’s constant) – and that this entropy can be interpreted thermodynamically in terms of an energy flow through the lightsheet. The entropy vanishes from the derivation and the Einstein equation almost immediately appears as a thermodynamic equation of state. What we realized, however, is that what the entropy is actually the entropy of was ambiguous in thermodynamic gravity. Surprisingly, we found that there doesn’t seem to be a consistent definition of the entropy in thermodynamic gravity – applying quantum field theory results for the energy and entanglement entropy, we found that thermodynamic gravity could not simultaneously reproduce the correct constant in the Einstein equation and in the entropy/area relation for black holes.

So when all is said and done, we’ve found that holographic gravity, but not thermodynamic gravity, is on the right track. To answer our own question in the title of the paper, we found – in admittedly somewhat technical language – that the vacuum-subtracted von Neumann entropy evaluated on the null boundary of small causal diamonds gives a consistent formulation of holographic gravity. The future looks exciting for finding the connections between gravity and entanglement!

Guest Post: Grant Remmen on Entropic Gravity Read More »

10 Comments

Reading List

Now that The Big Picture is complete, I have more time for fun things like blogging, but I have a bunch of research to catch up on before I can return as normal. So in the meantime, here’s another teaser from the book: my list of “Further Reading” keyed to the different sections. You should have enough time to read all of these between now and publication day, May 10.

Part One, Cosmos:

  • Adams, F., & Laughlin, G. (1999). The Five Ages of the Universe: Inside the Physics of Eternity. Free Press.
  • Albert, D.Z. (2003). Time and Chance. Harvard University Press.
  • Carroll, S. (2010). From Eternity to Here: The Quest for the Ultimate Theory of Time. Dutton.
  • Feynman, R.P. (1967). The Character of Physical Law. M.I.T. Press.
  • Greene, B. (2004). The Fabric of the Cosmos: Space, Time, and the Texture of Reality. A.A. Knopf.
  • Guth, A. (1997). The Inflationary Universe: The Quest for a New Theory of Cosmic Origins. Addison-Wesley Pub.
  • Hawking, S.W. and Mlodinow, L. (2010). The Grand Design. Bantam.
  • Pearl, J. (2009). Causality: Models, Reasoning, and Inference. Cambridge University Press.
  • Penrose, R. (2005). The Road to Reality: A Complete Guide to the Laws of the Universe. A.A. Knopf.
  • Weinberg, S. (2015). To Explain the World: The Discovery of Modern Science. HarperCollins.

Part Two, Understanding:

  • Ariely, D. (2008). Predictably Irrational: The Hidden Forces that Shape Our Decisions. HarperCollins.
  • Dennett, D.C. (2014) Intuition Pumps and Other Tools for Thinking. W.W. Norton.
  • Gillett, C. and Lower, B., eds. (2001). Physicalism and Its Discontents. Cambridge University Press.
  • Kaplan, E. (2014). Does Santa Exist? A Philosophical Investigation. Dutton.
  • Rosenberg, A. (2011). The Atheist’s Guide to Reality: Enjoying Life Without Illusions. W.W. Norton.
  • Sagan, C. (1995). The Demon-Haunted World: Science as a Candle in the Dark. Random House.
  • Silver, N. (2012). The Signal and the Noise: Why So Many Predictions Fail — But Some Don’t. Penguin Press.
  • Tavris, C. and Aronson, E. (2006). Mistakes Were Made (but not by me): Why We Justify Foolish Beliefs, Bad Decisions, and Hurtful Acts. Houghton Mifflin Harcourt.

Part Three, Essence:

  • Aaronson, S. (2013). Quantum Computing Since Democritus. Cambridge University Press.
  • Carroll, S. (2012). The Particle at the End of the Universe: How the Hunt for the Higgs Boson Leads Us to the Edge of a New World. Dutton.
  • Deutsch, D. (1997). The Fabric of Reality: The Science of Parallel Universes and Its Implications. Viking Adult.
  • Gefter, A. (2014). Trespassing on Einstein’s Lawn: A Father, a Daughter, the Meaning of Nothing, and the Beginning of Everything. Bantam.
  • Holt, J. (2012) Why Does the World Exist? An Existential Detective Story. Liveright Publishing.
  • Musser, G. (2015). Spooky Action at a Distance: The Phenomenon That Reimagines Space and Time–and What It Means for Black Holes, the Big Bang, and Theories of Everything. Scientific American / Farrar, Straus and Giroux.
  • Randall, L. (2011). Knocking on Heaven’s Door: How Physics and Scientific Thinking Illuminate the Universe and the Modern World. Ecco.
  • Wallace, D. (2014). The Emergent Multiverse: Quantum Theory According to the Everett Interpretation. Oxford University Press.
  • Wilczek, F. (2015). A Beautiful Question: Finding Nature’s Deep Design. Penguin Press.

Part Four, Complexity:

  • Bak, P. (1996). How Nature Works: The Science of Self-Organized Criticality. Copernicus.
  • Cohen, E. (2012). Cells to Civilizations: The Principles of Change that Shape Life. Princeton University Press.
  • Coyne, J. (2009). Why Evolution is True. Viking.
  • Dawkins, R. (1986). The Blind Watchmaker: Why the Evidence of Evolution Reveals a Universe without Design. W.W. Norton.
  • Dennett, D.C. (1995). Darwin’s Dangerous Idea: Evolution and the Meanings of Life. Simon & Schuster.
  • Hidalgo, C. (2015). Why Information Grows: The Evolution of Order, from Atoms to Economies. Basic Books.
  • Hoffman, P. (2012). Life’s Ratchet: How Molecular Machines Extract Order from Chaos. Basic Books.
  • Krugman, P. (1996). The Self-Organizing Economy. Wiley-Blackwell.
  • Lane, N. (2015). The Vital Question: Energy, Evolution, and the Origins of Complex Life. W.W. Norton.
  • Mitchell, M. (2009). Complexity: A Guided Tour. Oxford University Press.
  • Pross, A. (2012). What Is Life? How Chemistry Becomes Biology. Oxford University Press.
  • Rutherford, A. (2013). Creation: How Science is Reinventing Life Itself. Current.
  • Shubin, N. (2008). Your Inner Fish: A Journey into the 3.5-Billion-Year History of the Human Body. Pantheon.

Part Five, Thinking:

  • Alter, T. and Howell, R.J. (2009). A Dialogue on Consciousness. Oxford University Press.
  • Chalmers, D.J. (1996). The Conscious Mind: In Search of a Fundamental Theory. Oxford University Press.
  • Churchland, P.S. (2013). Touching a Nerve: The Self as Brain. W.W. Norton.
  • Damasio, A. (2010). Self Comes to Mind: Constructing the Conscious Brain. Pantheon.
  • Dennett, D.C. (1991). Consciousness Explained. Little Brown & Co.
  • Eagleman, D. (2011). Incognito: The Secret Lives of the Brain. Pantheon.
  • Flanagan, O. (2003). The Problem of the Soul: Two Visions of Mind and How to Reconcile Them. Basic Books.
  • Gazzaniga, M.S. (2011). Who’s In Charge? Free Will and the Science of the Brain. Ecco.
  • Hankins, P. (2015). The Shadow of Consciousness.
  • Kahneman, D. (2011). Thinking, Fast and Slow. Farrah, Straus and Giroux.
  • Tononi, G. (2012). Phi: A Voyage from the Brain to the Soul. Pantheon.

Part Six, Caring:

  • de Waal, F. (2013). The Bonobo and the Atheist: In Search of Humanism Among the Primates. W.W. Norton.
  • Epstein, G.M. (2009). Good Without God: What a Billion Nonreligious People Do Believe. William Morrow.
  • Flanagan, O. (2007). The Really Hard Problem: Meaning in a Material World. The MIT Press.
  • Gottschall, J. (2012). The Storytelling Animal: How Stories Make Us Human. Houghton Mifflin Harcourt.
  • Greene, J. (2013). Moral Tribes: Emotion, Reason, and the Gap Between Us and Them. Penguin Press.
  • Johnson, C. (2014). A Better Life: 100 Atheists Speak Out on Joy & Meaning in a World Without God. Cosmic Teapot.
  • Kitcher, P. (2011). The Ethical Project. Harvard University Press.
  • Lehman, J. and Shemmer, Y. (2012). Constructivism in Practical Philosophy. Oxford University Press.
  • May, T. (2015). A Significant Life: Human Meaning in a Silent Universe. University of Chicago Press.
  • Ruti, M. (2014). The Call of Character: Living a Life Worth Living. Columbia University Press.
  • Wilson, E.O. (2014). The Meaning of Human Existence. Liveright.

CVUeQgqUAAAWuBv

Reading List Read More »

33 Comments

Guest Post: Aidan Chatwin-Davies on Recovering One Qubit from a Black Hole

47858f217602be036c32e8ac76271a75_400x400 The question of how information escapes from evaporating black holes has puzzled physicists for almost forty years now, and while we’ve learned a lot we still don’t seem close to an answer. Increasingly, people who care about such things have been taking more seriously the intricacies of quantum information theory, and learning how to apply that general formalism to the specific issues of black hole information.

Now two students and I have offered a small contribution to this effort. Aidan Chatwin-Davies is a grad student here at Caltech, while Adam Jermyn was an undergraduate who has now gone on to do graduate work at Cambridge. Aidan came up with a simple method for getting out one “quantum bit” (qubit) of information from a black hole, using a strategy similar to “quantum teleportation.” Here’s our paper that just appeared on arxiv:

How to Recover a Qubit That Has Fallen Into a Black Hole
Aidan Chatwin-Davies, Adam S. Jermyn, Sean M. Carroll

We demonstrate an algorithm for the retrieval of a qubit, encoded in spin angular momentum, that has been dropped into a no-firewall unitary black hole. Retrieval is achieved analogously to quantum teleportation by collecting Hawking radiation and performing measurements on the black hole. Importantly, these methods only require the ability to perform measurements from outside the event horizon and to collect the Hawking radiation emitted after the state of interest is dropped into the black hole.

It’s a very specific — i.e. not very general — method: you have to have done measurements on the black hole ahead of time, and then drop in one qubit, and we show how to get it back out. Sadly it doesn’t work for two qubits (or more), so there’s no obvious way to generalize the procedure. But maybe the imagination of some clever person will be inspired by this particular thought experiment to come up with a way to get out two qubits, and we’ll be off.

I’m happy to host this guest post by Aidan, explaining the general method behind our madness.


If you were to ask someone on the bus which of Stephen Hawking’s contributions to physics he or she thought was most notable, the answer that you would almost certainly get is his prediction that a black hole should glow as if it were an object with some temperature. This glow is made up of thermal radiation which, unsurprisingly, we call Hawking radiation. As the black hole radiates, its mass slowly decreases and the black hole decreases in size. So, if you waited long enough and were careful not to enlarge the black hole by throwing stuff back in, then eventually it would completely evaporate away, leaving behind nothing but a bunch of Hawking radiation.

At a first glance, this phenomenon of black hole evaporation challenges a central notion in quantum theory, which is that it should not be possible to destroy information. Suppose, for example, that you were to toss a book, or a handful of atoms in a particular quantum state into the black hole. As the black hole evaporates into a collection of thermal Hawking particles, what happens to the information that was contained in that book or in the state of (what were formerly) your atoms? One possibility is that the information actually is destroyed, but then we would have to contend with some pretty ugly foundational consequences for quantum theory. Instead, it could be that the information is preserved in the state of the leftover Hawking radiation, albeit highly scrambled and difficult to distinguish from a thermal state. Besides being very pleasing on philosophical grounds, we also have evidence for the latter possibility from the AdS/CFT correspondence. Moreover, if the process of converting a black hole to Hawking radiation conserves information, then a stunning result of Hayden and Preskill says that for sufficiently old black holes, any information that you toss in comes back out almost a fast as possible!

Even so, exactly how information leaks out of a black hole and how one would go about converting a bunch of Hawking radiation to a useful state is quite mysterious. On that note, what we did in a recent piece of work was to propose a protocol whereby, under very modest and special circumstances, you can toss one qubit (a single unit of quantum information) into a black hole and then recover its state, and hence the information that it carried.

More precisely, the protocol describes how to recover a single qubit that is encoded in the spin angular momentum of a particle, i.e., a spin qubit. Spin is a property that any given particle possesses, just like mass or electric charge. For particles that have spin equal to 1/2 (like those that we consider in our protocol), at least classically, you can think of spin as a little arrow which points up or down and says whether the particle is spinning clockwise or counterclockwise about a line drawn through the arrow. In this classical picture, whether the arrow points up or down constitutes one classical bit of information. According to quantum mechanics, however, spin can actually exist in a superposition of being part up and part down; these proportions constitute one qubit of quantum information.

spin

So, how does one throw a spin qubit into a black hole and get it back out again? Suppose that Alice is sitting outside of a black hole, the properties of which she is monitoring. From the outside, a black hole is characterized by only three properties: its total mass, total charge, and total spin. This latter property is essentially just a much bigger version of the spin of an individual particle and will be important for the protocol.

Next, suppose that Alice accidentally drops a spin qubit into the black hole. First, she doesn’t panic. Instead, she patiently waits and collects one particle of Hawking radiation from the black hole. Crucially, when a Hawking particle is produced by the black hole, a bizarro version of the same particle is also produced, but just behind the black hole’s horizon (boundary) so that it falls into the black hole. This bizarro ingoing particle is the same as the outgoing Hawking particle, but with opposite properties. In particular, its spin state will always be flipped relative to the outgoing Hawking particle. (The outgoing Hawking particle and the ingoing particle are entangled, for those in the know.)

singlePic

The picture so far is that Alice, who is outside of the black hole, collects a single particle of Hawking radiation whilst the spin qubit that she dropped and the ingoing bizarro Hawking particle fall into the black hole. When the dropped particle and the bizarro particle fall into the black hole, their spins combine with the spin of the black hole—but remember! The bizarro particle’s spin was highly correlated with the spin of the outgoing Hawking particle. As such, the new combined total spin of the black hole becomes highly correlated with the spin of the outgoing Hawking particle, which Alice now holds. So, Alice measures the black hole’s new total spin state. Then, essentially, she can exploit the correlations between her held Hawking particle and the black hole to transfer the old spin state of the particle that she dropped into the hole to the Hawking particle that she now holds. Alice’s lost qubit is thus restored. Furthermore, Alice didn’t even need to know the precise state that her initial particle was in to begin with; the qubit is recovered regardless!

That’s the protocol in a nutshell. If the words “quantum teleportation” mean anything to you, then you can think of the protocol as a variation on the quantum teleportation protocol where the transmitting party is the black hole and measurement is performed in the total angular momentum basis instead of the Bell basis. Of course, this is far from a resolution of the information problem for black holes. However, it is certainly a neat trick which shows, in a special set of circumstances, how to “bounce” a qubit of quantum information off of a black hole.

Guest Post: Aidan Chatwin-Davies on Recovering One Qubit from a Black Hole Read More »

23 Comments

Guest Post: Don Page on God and Cosmology

Don Page is one of the world’s leading experts on theoretical gravitational physics and cosmology, as well as a previous guest-blogger around these parts. (There are more world experts in theoretical physics than there are people who have guest-blogged for me, so the latter category is arguably a greater honor.) He is also, somewhat unusually among cosmologists, an Evangelical Christian, and interested in the relationship between cosmology and religious belief.

Longtime readers may have noticed that I’m not very religious myself. But I’m always willing to engage with people with whom I disagree, if the conversation is substantive and proceeds in good faith. I may disagree with Don, but I’m always interested in what he has to say.

Recently Don watched the debate I had with William Lane Craig on “God and Cosmology.” I think these remarks from a devoted Christian who understands the cosmology very well will be of interest to people on either side of the debate.


Open letter to Sean Carroll and William Lane Craig:

I just ran across your debate at the 2014 Greer-Heard Forum, and greatly enjoyed listening to it. Since my own views are often a combination of one or the others of yours (though they also often differ from both of yours), I thought I would give some comments.

I tend to be skeptical of philosophical arguments for the existence of God, since I do not believe there are any that start with assumptions universally accepted. My own attempt at what I call the Optimal Argument for God (one, two, three, four), certainly makes assumptions that only a small fraction of people, and perhaps even only a small fraction of theists, believe in, such as my assumption that the world is the best possible. You know that well, Sean, from my provocative seminar at Caltech in November on “Cosmological Ontology and Epistemology” that included this argument at the end.

I mainly think philosophical arguments might be useful for motivating someone to think about theism in a new way and perhaps raise the prior probability someone might assign to theism. I do think that if one assigns theism not too low a prior probability, the historical evidence for the life, teachings, death, and resurrection of Jesus can lead to a posterior probability for theism (and for Jesus being the Son of God) being quite high. But if one thinks a priori that theism is extremely improbable, then the historical evidence for the Resurrection would be discounted and not lead to a high posterior probability for theism.

I tend to favor a Bayesian approach in which one assigns prior probabilities based on simplicity and then weights these by the likelihoods (the probabilities that different theories assign to our observations) to get, when the product is normalized by dividing by the sum of the products for all theories, the posterior probabilities for the theories. Of course, this is an idealized approach, since we don’t yet have _any_ plausible complete theory for the universe to calculate the conditional probability, given the theory, of any realistic observation.

For me, when I consider evidence from cosmology and physics, I find it remarkable that it seems consistent with all we know that the ultimate theory might be extremely simple and yet lead to sentient experiences such as ours. A Bayesian analysis with Occam’s razor to assign simpler theories higher prior probabilities would favor simpler theories, but the observations we do make preclude the simplest possible theories (such as the theory that nothing concrete exists, or the theory that all logically possible sentient experiences occur with equal probability, which would presumably make ours have zero probability in this theory if there are indeed an infinite number of logically possible sentient experiences). So it seems mysterious why the best theory of the universe (which we don’t have yet) may be extremely simple but yet not maximally simple. I don’t see that naturalism would explain this, though it could well accept it as a brute fact.

One might think that adding the hypothesis that the world (all that exists) includes God would make the theory for the entire world more complex, but it is not obvious that is the case, since it might be that God is even simpler than the universe, so that one would get a simpler explanation starting with God than starting with just the universe. But I agree with your point, Sean, that theism is not very well defined, since for a complete theory of a world that includes God, one would need to specify the nature of God.

For example, I have postulated that God loves mathematical elegance, as well as loving to create sentient beings, so something like this might explain both why the laws of physics, and the quantum state of the universe, and the rules for getting from those to the probabilities of observations, seem much simpler than they might have been, and why there are sentient experiences with a rather high degree of order. However, I admit there is a lot of logically possible variation on what God’s nature could be, so that it seems to me that at least we humans have to take that nature as a brute fact, analogous to the way naturalists would have to take the laws of physics and other aspects of the natural universe as brute facts. I don’t think either theism or naturalism solves this problem, so it seems to me rather a matter of faith which makes more progress toward solving it. That is, theism per se cannot deduce from purely a priori reasoning the full nature of God (e.g., when would He prefer to maintain elegant laws of physics, and when would He prefer to cure someone from cancer in a truly miraculous way that changes the laws of physics), and naturalism per se cannot deduce from purely a priori reasoning the full nature of the universe (e.g., what are the dynamical laws of physics, what are the boundary conditions, what are the rules for getting probabilities, etc.).

In view of these beliefs of mine, I am not convinced that most philosophical arguments for the existence of God are very persuasive. In particular, I am highly skeptical of the Kalam Cosmological Argument, which I shall quote here from one of your slides, Bill:

  1. If the universe began to exist, then there is a transcendent cause
    which brought the universe into existence.
  2. The universe began to exist.
  3. Therefore, there is a transcendent cause which brought the
    universe into existence.

I do not believe that the first premise is metaphysically necessary, and I am also not at all sure that our universe had a beginning. …

Guest Post: Don Page on God and Cosmology Read More »

960 Comments

The Science of Interstellar

The intersection — maybe the union! — of science and sci-fi geekdom is overcome with excitement about the upcoming movie Interstellar, which opens November 7. It’s a collaboration between director Christopher Nolan and physicist Kip Thorne, both heroes within their respective communities. I haven’t seen it yet myself, nor do I know any secret scoop, but there’s good reason to believe that this film will have some of the most realistic physics of any recent blockbuster we’ve seen. If it’s a success, perhaps other filmmakers will take the hint?

Kip, who is my colleague at Caltech (and a former guest-blogger), got into the science-fiction game quite a while back. He helped Carl Sagan with some science advice for his book Contact, later turned into a movie starring Jodie Foster. In particular, Sagan wanted to have some way for his characters to traverse great distances at speeds faster than light, by taking a shortcut through spacetime. Kip recognized that a wormhole was what was called for, but also realized that any form of faster-than-light travel had the possibility of leading to travel backwards in time. Thus was the entire field of wormhole time travel born.

As good as the movie version of Contact was, it still strayed from Sagan’s original vision, as his own complaints show. (“Ellie disgracefully waffles in the face of lightweight theological objections to rationalism…”) Making a big-budget Hollywood film is necessarily a highly collaborative endeavor, and generally turns into a long series of forced compromises. Kip has long been friends with Lynda Obst, an executive producer on Contact, and for years they batted around ideas for a movie that would really get the science right.

Long story short, Lynda and Kip teamed with screenwriter Jonathan Nolan (brother of Christopher), who wrote a draft of a screenplay, and Christopher eventually agreed to direct. I know that Kip has been very closely involved with the script as the film has developed, and he’s done his darnedest to make sure the science is right, or at least plausible. (We don’t actually whether wormholes are allowed by the laws of physics, but we don’t know that they’re not allowed.) But it’s a long journey, and making the best movie possible is the primary goal. Meanwhile, Adam Rogers at Wired has an in-depth look at the science behind the movie, including the (unsurprising, in retrospect) discovery that the super-accurate visualization software available to the Hollywood special-effects team enable the physicists to see things they hadn’t anticipated. Kip predicts that at least a couple of technical papers will come out of their work.

And that’s not all! Kip has a book coming out on the science behind the movie, which I’m sure will be fantastic. And there is also a documentary on “The Science of Interstellar” that will be shown on TV, in which I play a tiny part. Here is the broadcast schedule for that, as I understand it:

SCIENCE
Wednesday, October 29, at 10pm PDT/9c

AHC (American Heroes Channel)
Sunday, November, 2 at 4pm PST/3c (with a repeat on Monday, November 3 at 4am PST/3c)

DISCOVERY
Thursday, November 6, at 11pm PST/10c

Of course, all the accurate science in the world doesn’t help if you’re not telling an interesting story. But with such talented people working together, I think some optimism is justified. Let’s show the world that science and cinema are partners, not antagonists.

Interstellar Movie - Official Trailer 3

The Science of Interstellar Read More »

47 Comments
Scroll to Top