Archive for the ‘Mirrored on CSAIL Blog’ Category

Thanksgiving Special: D-Wave at MIT

Thursday, November 22nd, 2007

Some people think I have a vendetta against D-Wave Systems and its questionable quantum computer claims (see here, here, here, here, here, here, here, here, here for context). But actually, nothing could be further from the truth. I keep trying and trying to change the subject! Wouldn’t you all rather hear about Wolfram, I say? Or unparadoxes? Or my #1 topic du jour, nothing whatsoever?

Apparently you wouldn’t. From my inbox to my comments section to the hallway, the masses have spoken, and what they want to know is: did I attend D-Wave’s presentation at MIT on Monday, and if so what did I think?

Yes, I attended, in body though not in mind. You see, Monday was also the day of the STOC deadline, so if our guests from D-Wave (Mohammad Amin and Andrew Berkley) were expecting a ferocious skeptic, they instead got a bleary-eyed zombie with visions of MAEXP, P/poly, and 7:59PM EST cavorting in his head.

This meant that Ed Farhi, Isaac Chuang, Peter Shor, and Lorenza Viola had to do most of the questioning. As it turned out, they did a vastly better job than I could have.

As others have pointed out in stronger terms, I’m not a physicist. (On the other hand, the gentleman linked to in the previous sentence is not correct about my being paid by the NSA to discredit Canadian quantum computing efforts: it’s actually the GCHQ and the Mossad.) As such, I can’t directly evaluate D-Wave’s central claim to have built an adiabatic quantum computer, nor have I ever tried to do so. All I can do is point out the many things D-Wave has said to the press (about NP-complete problems, for example) that I know are false, its history of making dramatic announcements without evidence, and its contemptuous attitude toward scientists who have asked for such evidence. For me, that’s more than enough to destroy D-Wave’s credibility on the claims I can’t directly evaluate. After all, the burden of proof is not on me; it’s on them.

However, other people have not been satisfied with this line of argument. “We don’t care who the burden the proof is on,” they say. “We just care whether D-Wave built an adiabatic quantum computer.”

But my physicist colleagues don’t suffer from the same argumentative limitations that I do. At the group meeting preceding the talk, Farhi announced that he didn’t care what the press releases said, nor did he want to discuss what problems quantum computers can solve (since we academics can figure that out ourselves). Instead he wanted to focus on a single question: is D-Wave’s device a quantum computer or not?

What followed was probably the most intense grilling of an invited speaker I’ve ever seen.

It quickly emerged that D-Wave wants to run a coherent quantum computation for microseconds, even though each of their superconducting qubits will have completely decohered within nanoseconds. Farhi had to ask Amin to repeat this several times, to make sure he’d gotten it right.

Amin’s claim was that what looks like total decoherence in the computational basis is irrelevant — since for adiabatic quantum computation, all that matters is what happens in the basis of energy eigenstates. In particular, Amin claimed to have numerical simulations showing that, if the temperature is smaller than the spectral gap, then one can do adiabatic quantum computation even if the conventional coherence times (the t1 and t2) would manifestly seem to prohibit it.

The physicists questioned Amin relentlessly on this one claim. I think it’s fair to say that they emerged curious but severely skeptical, not at all convinced by the calculations Amin provided, and determined to study the issue for themselves.

In other words, this was science as it should be. In contrast to their bosses, Amin and Berkley made a genuine effort to answer questions. They basically admitted that D-Wave’s press releases were litanies of hype and exaggeration, but nevertheless thought they had a promising path to a quantum computer. On several occasions, they seemed to be struggling to give an honest answer that would still uphold the company line.

Two other highlights:

  • I asked Amin and Berkley whether they could give any evidence for any sort of speedup over classical simulated annealing. They laughed at this. “It’s sixteen qubits!” they said. “Of course you’re not going to see a scaling effect with sixteen qubits.”I said I understood perfectly well (though I wondered silently whether the dozens of journalists covering D-Wave’s demo understood the same). But, I continued, surely you should be able to see a scaling effect by the end of 2008, when your business plan calls for 1024 qubits?”Well, that’s what it says in the press release,” they said.Forget about the press release, Farhi interjected. How many qubits are you actually going to make?Amin and Berkley shrugged; they said they’d just try to make as many qubits as they could.
  • Even though it hadn’t exhibited any sort of speedup, Amin and Berkley steadfastly maintained that their 16-qubit device was indeed a quantum computer. Their evidence was that simulations of its behavior that took quantum mechanics into account gave, they said, a better fit to the data than simulations that didn’t. On the other hand, they said they were not able to test directly for the presence of any quantum effect such as entanglement. (They agreed that entanglement was a non-negotiable requirement for quantum computing.)There was a Feynmanesque moment, when Ike Chuang asked Amin and Berkley an experimental question so simple even I understood it. Ike said: if you’re indeed seeing quantum effects, then by running your computer at higher and higher temperatures, at some point you should see a transition to classical behavior. Have you tried this simple control experiment?Amin and Berkley said that they hadn’t, but that it sounded like a good idea.

For a theorist like me — accustomed to talks ending with “if there are no questions, then let’s thank the speaker again” — this was exciting, heady stuff. And when it was over, I still had almost three hours until the STOC deadline.

Deoxyribononapproximability

Sunday, November 4th, 2007

[Updates]

Alright, here’s a problem for all you bioinformatistas and inapproximabistas out there, which was inspired by this post of Eliezer Yudkowsky at Overcoming Bias (see also the comments there).

Let a DNA sequence be an element of {A,C,G,T}*, and suppose we’re allowed the following primitive operations: (1) insert a base pair anywhere we want, (2) delete any substring, (3) reverse any substring, and (4) copy any substring into any other part of the string. Then given a DNA sequence S, how hard is it to estimate the minimum number of operations needed to produce S starting from the empty string?

Closely related is the following problem: by starting from the empty string and applying o(n) operations, can we produce a “pseudorandom DNA sequence” of length n — that is, a sequence that can’t be distinguished in polynomial time from a uniform random one?

(Note 1: For both problems, we might also want to stipulate that every intermediate sequence should have size at most polynomial in n. Or better yet, maybe one can prove that such an assumption is without loss of generality.)

(Note 2: I’m also very interested in what happens if we disallow the powerful operation of reversal.)

For all I know, these problems might have trivial (or at any rate, known) answers; I just came up with them and haven’t thought them through.

What the problems are really getting at is this: is the “effective number of bits” in your genome (that is, the number of bits from a polynomial-time algorithm’s perspective) limited by how many ancestors you’ve had since life on Earth began? Or can it be vastly greater?

Update (11/4): Rereading the last few paragraphs of Eliezer’s post, I see that he actually argues for his central claim — that the human genome can’t contain more than 25MB of “meaningful DNA” — on different (and much stronger) grounds than I thought! My apologies for not reading more carefully.

In particular, the argument has nothing to do with the number of generations since the dawn of time, and instead deals with the maximum number of DNA bases that can be simultaneously protected, in steady state, against copying errors. According to Eliezer, copying a DNA sequence involves a ~10-8 probability of error per base pair, which — because only O(1) errors per generation can be corrected by natural selection — yields an upper bound of ~108 on the number of “meaningful” base pairs in any given genome.

However, while this argument is much better than my straw-man based on the number of generations, there’s still an interesting loophole. Even with a 10-8 chance of copying errors, one could imagine a genome reliably encoding far more than 108 bits (in fact, arbitrarily many bits) by using an error-correcting code. I’m not talking about the “local” error-correction mechanisms that we know DNA has, but about something more global — by which, say, copying errors in any small set of genes could be completely compensated by other genes. The interesting question is whether natural selection could read the syndrome of such a code, and then correct it, using O(1) randomly-chosen insertions, deletions, transpositions, and reversals. I admit that this seems unlikely, and that even if it’s possible in principle, it’s probably irrelevant to real biology. For apparently there are examples where changing even a single base pair leads to horrible mutations. And on top of that, we can’t have the error-correcting code be too good, since otherwise we’ll suppress beneficial mutations!

Incidentally, Eliezer’s argument makes the falsifiable prediction that we shouldn’t find any organism, anywhere in nature, with more than 25MB of functional DNA. Does anyone know of a candidate counterexample? (I know there are organisms with far more than humans’ 3 billion base pairs, but I have no idea how many of the base pairs are functional.)

Lastly, in spite of everything above, I’d still like a solution to my “pseudorandom DNA sequence” problem. For if the answer were negative — if given any DNA sequence, one could efficiently reconstruct a nearly-optimal sequence of insertions, transpositions, etc. producing it — then even my original straw-man misconstrual of Eliezer’s argument could put up a decent fight!

Update (11/5): Piotr Indyk pointed me to a paper by Ergün, Muthukrishnan, and Sahinalp from FSTTCS’2003, which basically solves my problem in the special case of no reversals. It turns out that you can estimate the number of insert, delete, and copy operations needed to produce a given DNA sequence to within a factor of 4, by just applying Lempel-Ziv compression to the sequence. Thanks, Piotr!

Another Update (11/5): Andy Drucker has pointed out that, in the case where reversals are allowed, we can approximate the number of insert, delete, copy, and reverse operations needed to produce a given DNA sequence to within a factor of 16, by combining the Lempel-Ziv approach of Ergün et al. with a clever trick: maintain both the sequence and its reversal at all times! Interestingly, though, this trick doesn’t seem to work for transforming one sequence into another (a more general problem than I asked about, and the one considered by Ergün et al).

Unparadox Contest

Thursday, November 1st, 2007

In a recent talk at MIT, Umesh Vazirani appealed to the famous Birthday Paradox to say that two random subsets of {1,…,N}, each of size o(√N), probably wouldn’t intersect each other. Of course we all understood what he meant, but it occurred to me that Umesh was actually appealing to the Birthday Unparadox: “If you put three people in a room, chances are no two of them will have the same birthday.”

Once I realized that, I started seeing unparadoxes everywhere I looked:

The Banach-Tarski Unparadox: If you cut an orange into five pieces using a standard knife, then put them back together, the result will have exactly the same volume as the original orange.

Braess’ Unparadox: If you add an extra lane to a highway, one possible result will be to decrease congestion.

Hempel’s Unparadox: If you observe a bunch of ravens and find that all of them are black, this might increase your likelihood for the statement “All ravens are black.”

Russell’s Unparadox: The set of all sets that contain themselves as a member, might or might not contain itself as a member (either way is fine).

In the spirit of my highly-successful Best Umeshism and Best Anthropicism contests (remember those?), I now open the floor to you: come up with the best unparadox! The winner will receive absolutely nothing. (If you have to ask what the point is, this contest isn’t for you.)

The Aaronson $25.00 Prize

Sunday, October 28th, 2007

[Update]

For those of you who’ve been living in a non-wifi-enabled cave, four days ago Stephen Wolfram awarded a $25,000 prize to a 20-year-old undergraduate named Alex Smith, for proving that a particular two-state, three-symbol Turing machine is universal. The prize was to celebrate the fifth anniversary of Wolfram’s paradigm-smashing, foundation-shaking masterpiece, A New Kind of Science. (More from Bill Gasarch’s blog, Nature, and Scientific American.)

Smith sounds like a swell guy who entered this contest for exactly the right reasons: he was at his parents’ place over summer break and had nothing better to do. He deserves the money, and I sincerely hope the CS theory community hasn’t heard the last from him.

Predictably, though, as soon as this story broke I started getting emails from journalists asking me about the far-reaching scientific implications of the new universality proof. In trying to give them an honest answer — one that wouldn’t be misunderstood, or spun to support a pre-existing storyline with which I disagreed — I inevitably came off like an ornery old sourpuss. From Scientific American:

Of course, there may be a reason the problem languished. “Finding the smallest universal [Turing machines] is a neat recreational pursuit,” quantum computation researcher Scott Aaronson of the Massachusetts Institute of Technology says, but “it’s no longer seen as connected to the central questions of the field.” …

“The impact of NKS on all the areas of computer science and physics I’m familiar with has been basically zero,” he says. “As far as I can tell, the main impact is that people now sometimes use the adjective ‘Wolframian’ to describe breathtaking claims for the trivial or well-known.” [Martin] Davis offers a sunnier take: “The book has a lot of beautiful pictures.”

And from Nature:

The solution isn’t hugely relevant to modern computer science, says Scott Aaronson, a computer scientist at the Massachusetts Institute of Technology (MIT) in Cambridge, Massachusetts. “Most theoretical computer scientists don’t particularly care about finding the smallest universal Turing machines,” he wrote in an e-mail. “They see it as a recreational pursuit that interested people in the 60s and 70s but is now sort of ‘retro’.”

Having partially degrumpified, in the remainder of this post I wish to offer something positive.

But first some background: a month after NKS came out, I wrote a review of it for the journal Quantum Information and Computation, in which I examined Wolfram’s claims about quantum mechanics and computational complexity, and explained what I saw as the problems with them. (Rather than rehash the review, I’ll just point you there if you’re interested.)

Today I’d like to celebrate the fifth anniversary of my critical review of NKS, by offering a $25 prize for stodgy, conventional work in the field of quantum complexity theory.

The Aaronson $25.00 Challenge

In NKS, Wolfram places himself among those computer scientists and physicists who doubt the possibility of quantum computers, not for any practical reason but as a consequence of their disbelieving quantum mechanics itself. As he writes on page 771:

Indeed within the usual formalism [of quantum mechanics] one can construct quantum computers that may be able to solve at least a few specific problems exponentially faster than ordinary Turing machines. But particularly after my discoveries in Chapter 9 [‘Fundamental Physics’], I strongly suspect that even if this is formally the case, it will still not turn out to be a true representation of ultimate physical reality, but will instead just be found to reflect various idealizations made in the models used so far.

Here, then, is the challenge:

If a quantum computer can efficiently solve a problem, can it also efficiently convince Wolfram that the solution is correct? More formally, does every language in the class BQP admit an interactive protocol where the prover is in BQP and the verifier is in BPP?

In other words: can quantum computers always “show their work”? It’s obvious, for example, that if a quantum computer spit out the factors of a 5,000-digit number, you wouldn’t have to believe quantum mechanics (or even know what it was) to check whether the answer was right. I’m asking whether every problem solvable by a quantum computer has the same property. And to make things fair to the quantum computer, I’ll let it give not just a static proof but also an interactive protocol, by which a distrustful polynomial-time classical verifier could become convinced, to arbitrarily high confidence, that the quantum computer knows the right answer.

(An example for the uninitiated: suppose you had two graphs G and H, and suppose you picked one of the graphs at random, randomly permuted its vertices, and gave the result to a quantum computer. And suppose the quantum computer could unfailingly tell you which graph you started with. Clearly this should convince you that G and H are not isomorphic — since if they were isomorphic, then the quantum computer couldn’t have done better than guessing! And this is true even though you never received a proof of non-isomorphism that you could hand to someone else.)

I’ll award $25 either for a proof that every quantum computation can be “counter-Wolframized,” or for an oracle relative to which some quantum computation provably can’t be. If both problems are solved then I’ll award $25 for each. Every serious submission will be reviewed by a Prize Committee consisting of me. The Committee may also choose to award smaller prizes for partial results.

Note: Much as I’d like to “pull a Wolfram,” the beautiful question above was (to my knowledge) first asked by Daniel Gottesman, at a conference in February 2004. Also, the idea of a $25 prize was suggested to me by Mike Mosca.

Update (10/30): A commenter pointed me to this thread in the Foundations of Mathematics (FOM) mailing list, which contains an actual technical discussion of Smith’s universality proof. Of particular interest:

  1. an argument by Vaughan Pratt that Smith’s universality proof is wrong,
  2. a response by Todd Rowland of Wolfram Research,
  3. a post from Wolfram himself, which, though written in his trademark way, comes the closest I’ve seen of anything by him to addressing actual hard questions about the definition of universality, and
  4. this comment from John McCarthy: “In the 1950s I thought that the smallest possible (symbol-state product) universal Turing machine would tell something about the nature of computation. Unfortunately, it didn’t. Instead as simpler universal machines were discovered, the proofs that they were universal became more elaborate, and [so] did the encodings of information.”

In judging the correctness of Smith’s proof, the key question is what counts as “universality.” As I explained to the journalists who emailed me, the rules of Wolfram’s prize left a huge gray area by explicitly refusing to specify this. In particular: what kinds of input and output encodings are allowed? How do we make sure the real computational work is done by the Turing machine itself and not the encoding procedures? Does the tape have to be filled with 0’s initially, or can it be filled with other patterns, and if so which ones? Since the two-state Turing machine in question has no “halt” state, what external conditions can we impose to determine when the machine has halted?

Still, I decided not to make a fuss about such things in my original post, since it seemed clear from Smith’s writeup that (1) he was aware of these issues, and (2) there was some nontrivial sense in which he proved universality. I wasn’t going to lose sleep over which sense, for the simple reason that I’d never lost sleep over the (2,3) universality question itself!

Australian actresses are plagiarizing my quantum mechanics lecture to sell printers

Tuesday, October 2nd, 2007

I tried to think of a witty, ironic title for this post, but in the end, I simply couldn’t. The above title is a literal statement of fact.

A reader named Warren Smith informs me of an Australian TV commercial (which you can watch on YouTube), in which two fashion models have the following conversation:

Model 1: But if quantum mechanics isn’t physics in the usual sense — if it’s not about matter, or energy, or waves — then what is it about?

Model 2: Well, from my perspective, it’s about information, probabilities, and observables, and how they relate to each other.

Model 1: That’s interesting!

The commercial then flashes the tagline “A more intelligent model,” followed by a picture of a Ricoh printer.

More intelligent, or simply more shameless? Ladies and gentlemen of the jury, allow me to quote from Lecture 9 of my Quantum Computing Since Democritus notes:

But if quantum mechanics isn’t physics in the usual sense — if it’s not about matter, or energy, or waves, or particles — then what is it about? From my perspective, it’s about information and probabilities and observables, and how they relate to each other.

For almost the first time in my life, I’m at a loss for words. I don’t know how to respond. I don’t know which of 500,000 possible jokes to make. Help me, readers. Should I be flattered? Should I be calling a lawyer?

Update (10/3): [Sydney Morning Herald] [The Age] [Slashdot] [BoingBoing] [AdNews] [Scientific American blog] (Let me know if you find others) (I’m actually in Riga, Latvia, where it’s 8:30am, and just woke up to find all this. I’m going to take a shower now.)

Also: Please let me know if you get any more “CPU quota exceeded” errors. I just enabled an option to speed up PHP scripts, which might or might not solve the problem. Bluehost sucks — never use them for anything!

Update (10/4): Thanks, everyone, for the free legal advice! From half of you I’ve learned that I’d be an arrogant, stereotypically-American jerk to pursue this case further; from the other half I’ve learned that I’d be a naïve idiot not to. The longer I blog, the more I despair of ever achieving my central goal in life, namely for everyone to like me.

Also: Many commenters seem to assume the pilfered quote is just my expression of the conventional wisdom. Well, it should be, and I wish it were! But as Avi Wigderson pointed out to me, the idea that quantum mechanics is about information rather than waves or particles is still extremely non-standard, and would have been considered insane fifteen years ago.

Update (10/5): This is not to say it’s my idea (as other commenters assumed I was saying)! If any entity can claim “ownership” of the idea, I would think it’s the entire quantum computing and information community. The longer I blog, the more I despair of ever achieving my secondary goal in life, namely for everyone to understand me.

On Deutsches, Shors, and Vaziranis

Friday, September 28th, 2007

My friend Robin Hanson kvetches that scientific contrarians don’t get no respect: even if they ultimately turn out to be right, it’s the more cautious, Johnny-come-lately “conservative radicals” who get the lion’s share of the credit. (Like many of Robin’s posts, this one is written in a purely descriptive mode that nevertheless leaves no doubt where his sympathies lie.) And so, as a firmly-entrenched pillar of the hidebound scientific establishment, I thought I’d tell you our side of the story.

In the US, there are companies whose business it is to patent (or buy patents for) every obvious idea they can think of, then sit around and do nothing with those ideas, wait for some other company to build a successful business around one of them, and sue that company for patent infringement. (The textbook example is NTP’s lawsuit against Research In Motion.)

In science, one occasionally sees the intellectual equivalent of these patent-holding companies: people who publish one flaky idea after another with no data or calculations to back them up; and then if, after years of painstaking research, one of their speculations turns out to be right (or even 10% right), scream “they stole my idea!”

But in my experience — and happily for all concerned — the truth usually lies between this extreme and the opposite extreme described in Robin’s post. To illustrate, let’s consider the example of Robin’s that I know best: that of David Deutsch and quantum computing.

Unlike the patent-holding firms, David Deutsch really was a scientific pioneer, thinking deeply about quantum physics and the Church-Turing Thesis back when basically no one else was. His philosophical insights led him to define the quantum Turing machine model, prove its universality, and realize it might have implications for complexity theory. But his one concrete example of a quantum algorithm — how shall I say? — sucked. In particular, he gave an algorithm to compute the XOR of two bits (and know one has done so) using one quantum query and with success probability 1/2. (Later it was realized that success probability 1 is achievable, but that’s still only a factor-2 speedup compared to classical computers.) If this was all you’d seen of quantum computing, you would rightly file it away with dozens of other promising ideas that hadn’t led anywhere.

Unless, that is, you were Ethan Bernstein and Umesh Vazirani. These “conservative radicals” from Berkeley decided to put quantum computing under the microscope of theoretical computer science. The result of their labor — besides a bounteous harvest of complexity theorems like BPP ⊆ BQP ⊆ P#P — was the first example of a black-box problem for which quantum computers gave a superpolynomial speedup over classical randomized ones. Shortly afterward, another conservative, Dan Simon, set out to prove that the speedup of quantum computing was illusory — and ended up with strong evidence (now called Simon’s algorithm) for exactly the opposite conclusion. A year later, yet another conservative — an expert on combinatorics and discrete geometry by the name of Peter Shor — took a close look at Simon’s algorithm, and realized that if you changed the underlying group from (Z2)n to the cyclic group ZN, then you could efficiently compute the period of a black-box function, and thereby factor integers, and thereby break the RSA cryptosystem, and thereby change the world.

A Hansonian might downplay these later achievements — arguing that, were it not for Shor, some other “mainstream mathematician” (a strange description of him!) would’ve sooner or later discovered the factoring algorithm. But it’s equally true that, were it not for Deutsch, some other “renegade physicist” would have come up with quantum Turing machines (and indeed Feynman and Benioff were close). My own judgment is that Deutsch and Shor both made creative scientific contributions of the highest order, and are both deservedly celebrated for them. Indeed, if anyone gets short-shrifted in the usual popular accounts, I think it’s the people in between — like Bernstein, Vazirani, and Simon.

So yes, let’s remember the first person who struck gold, but also the first to realize it wasn’t fools’ gold and the first to figure out how to mine it. Science is a big place; there’s plenty of room for Deutsches, Shors, and even a Vazirani or two.

What Google Won’t Find

Friday, August 31st, 2007

While I rummage around the brain for something more controversial to blog (that’s nevertheless not too controversial), here, for your reading pleasure, is a talk I gave a couple weeks ago at Google Cambridge. Hardcore Shtetl-Optimized fans will find little here to surprise them, but for new or occasional readers, this is about the clearest statement I’ve written of my religio-ethico-complexity-theoretic beliefs.

What Google Won’t Find

As I probably mentioned when I spoke at your Mountain View location two years ago, it’s a funny feeling when an entity that knows everything that ever can be known or has been known or will be known invites you to give a talk — what are you supposed to say?

Well, I thought I’d talk about “What Google Won’t Find.” In other words, what have we learned over the last 15 years or so about the ultimate physical limits of search — whether it’s search of a physical database like Google’s, or of the more abstract space of solutions to a combinatorial problem?

On the spectrum of computer science, I’m about as theoretical as you can get. One way to put it is that I got through CS grad school at Berkeley without really learning any programming language other than QBASIC. So it might surprise you that earlier this year, I was spending much of my time talking to business reporters. Why? Because there was this company near Vancouver called D-Wave Systems, which was announcing to the press that it had built the world’s first commercial quantum computer.

Let’s ignore the “commercial” part, because I don’t really understand economics — these days, you can apparently make billions of dollars giving away some service for free! Let’s instead focus on the question: did D-Wave actually build a quantum computer? Well, they apparently built a device with 16 very noisy superconducting quantum bits (or qubits), which they say they’ve used to help solve extremely small Sudoku puzzles.

The trouble is, we’ve known for years that if qubits are sufficiently noisy — if they leak a sufficient amount of information into their environment — then they behave essentially like classical bits. Furthermore, D-Wave has refused to answer extremely basic technical questions about how high their noise rates are and so forth — they care about serving their customers, not answering nosy questions from academics. (Recently D-Wave founder Geordie Rose offered to answer my questions if I was interested in buying one of his machines. I replied that I was interested — my offer was $10 US — and I now await his answers as a prospective customer.)

To make a long story short, it’s consistent with the evidence that what D-Wave actually built would best be described as a 16-bit classical computer. I don’t mean 16 bits in terms of the architecture; I mean sixteen actual bits. And there’s some prior art for that.

But that’s actually not what annoyed me the most about the D-Wave announcement. What annoyed me were all the articles in the popular press — including places as reputable as The Economist — that said, what D-Wave has built is a machine that can try every possible solution in parallel and instantly pick the right one. This is what a quantum computer is; this is how it works.

It’s amazing to me how, as soon as the word “quantum” is mentioned, all the ordinary rules of journalism go out the window. No one thinks to ask: is that really what a quantum computer could do?

It turns out that, even though we don’t yet have scalable quantum computers, we do know something about what they could do if we did have them.

A quantum computer is a device that would exploit the laws of quantum mechanics to solve certain computational problems asymptotically faster than we know how to solve them with any computer today. Quantum mechanics — which has been our basic framework for physics for the last 80 years — is a theory that’s like probability theory, except that instead of real numbers called probabilities, you now have complex numbers called amplitudes. And the interesting thing about these complex numbers is that they can “interfere” with each other: they can cancel each other out.

In particular, to find the probability of something happening, you have to add the amplitudes for all the possible ways it could have happened, and then take the square of the absolute value of the result. And if some of the ways an event could happen have positive amplitude and others have negative amplitude, then the amplitudes can cancel out, so that the event doesn’t happen at all. This is exactly what’s going on in the famous double-slit experiment: at certain spots on a screen, the different paths a photon could’ve taken to get to that spot interfere destructively and cancel each other out, and as a result no photon is seen.

Now, the idea of quantum computing is to set up a massive double-slit experiment with exponentially many paths — and to try to arrange things so that the paths leading to wrong answers interfere destructively and cancel each other out, while the paths leading to right answers interfere constructively and are therefore observed with high probability.

You can see it’s a subtle effect that we’re aiming for. And indeed, it’s only for a few specific problems that people have figured out how to choreograph an interference pattern to solve the problem efficiently — that is, in polynomial time.

One of these problems happens to be that of factoring integers. Thirteen years ago, Peter Shor discovered that a quantum computer could efficient apply Fourier transforms over exponentially-large abelian groups, and thereby find the periods of exponentially-long periodic sequences, and thereby factor integers, and thereby break the RSA cryptosystem, and thereby snarf people’s credit card numbers. So that’s one application of quantum computers.

On the other hand — and this is the most common misconception about quantum computing I’ve encountered — we do not, repeat do not, know a quantum algorithm to solve NP-complete problems in polynomial time. For “generic” problems of finding a needle in a haystack, most of us believe that quantum computers will give at most a polynomial advantage over classical ones.

At this point I should step back. How many of you have heard of the following question: Does P=NP?

Yeah, this is a problem so profound that it’s appeared on at least two TV shows (The Simpsons and NUMB3RS). It’s also one of the seven (now six) problems for which the Clay Math Institute is offerring a million-dollar prize for a solution.

Apparently the mathematicians had to debate whether P vs. NP was “deep” enough to include in their list. Personally, I take it as obvious that it’s the deepest of them all. And the reason is this: if you had a fast algorithm for solving NP-complete problems, then not only could you solve P vs. NP, you could presumably also solve the other six problems. You’d simply program your computer to search through all possible proofs of at most (say) a billion symbols, in some formal system like Zermelo-Fraenkel set theory. If such a proof existed, you’d find it in a reasonable amount of time. (And if the proof had more than a billion symbols, it’s not clear you’d even want to see it!)

This raises an important point: many people — even computer scientists — don’t appreciate just how profound the consequences would be if P=NP. They think it’s about scheduling airline flights better, or packing more boxes in your truck. Of course, it is about those things — but the point is that you can have a set of boxes such that if you could pack them into your truck, then you would also have proved the Riemann Hypothesis!

Of course, while the proof eludes us, we believe that P≠NP. We believe there’s no algorithm to solve NP-complete problems in deterministic polynomial time. But personally, I would actually make a stronger conjecture:

There is no physical means to solve NP-complete problems in polynomial time — not with classical computers, not with quantum computers, not with anything else.

You could call this the “No SuperSearch Principle.” It says that, if you’re going to find a needle in a haystack, then you’ve got to expend at least some computational effort sifting through the hay.

I see this principle as analogous to the Second Law of Thermodynamics or the impossibility of superluminal signalling. That is, it’s a technological limitation which is also a pretty fundamental fact about the laws of physics. Like those other principles, it could always be falsified by experiment, but after a while it seems manifestly more useful to assume it’s true and then see what the consequences are for other things.

OK, so what do we actually know about the ability of quantum computers to solve NP-complete problems efficiently? Well, of course we can’t prove it’s impossible, since we can’t even prove it’s impossible for classical computers — that’s the P vs. NP problem! We might hope to at least prove that quantum computers can’t solve NP-complete problems in polynomial time unless classical computers can also — but even that, alas, seems far beyond our ability to prove.

What we can prove is this: suppose you throw away the structure of an NP-complete problem, and just consider it as an abstract, featureless space of 2n possible solutions, where the only thing you can do is guess a solution and check whether it’s right or not. In that case it’s obvious that a classical computer will need ~2n steps to find a solution. But what if you used a quantum computer, which could “guess” all possible solutions in superposition? Well, even then, you’d still need at least ~2n/2 steps to find a solution. This is called the BBBV Theorem, and was one of the first things learned about the power of quantum computers.

Intuitively, even though a quantum computer in some sense involves exponentially many paths or “parallel universes,” the single universe that happened on the answer can’t shout above all the other universes: “hey, over here!” It can only gradually make the others aware of its presence.

As it turns out, the 2n/2 bound is actually achievable. For in 1996, Lov Grover showed that a quantum computer can search a list of N items using only √N steps. It seems to me that this result should clearly feature in Google’s business plan.

Of course in real life, NP-complete problems do have structure, and algorithms like local search and backtrack search exploit that structure. Because of this, the BBBV theorem can’t rule out a fast quantum algorithm for NP-complete problems. It merely shows that, if such an algorithm existed, then it couldn’t work the way 99% of everyone who’s ever heard of quantum computing thinks it would!

You might wonder whether there’s any proposal for a quantum algorithm that would exploit the structure of NP-complete problems. As it turns out, there’s one such proposal: the “quantum adiabatic algorithm” of Farhi et al., which can be seen as the quantum version of simulated annealing. Intriguingly, Farhi and his collaborators proved that, on some problem instances where classical simulated annealing would take exponential time, the quantum adiabatic algorithm takes only polynomial time. Alas, we also know of problem instances where the adiabatic algorithm takes exponential time just as simulated annealing does. So while this is still an active research area, right now the adiabatic algorithm does not look like a magic bullet for solving NP-complete problems.

If quantum computers can’t solve NP-complete problems in polynomial time, it raises an extremely interesting question: is there any physical means to solve NP-complete problems in polynomial time?

Well, there have been lots of proposals. One of my favorites involves taking two glass plates with pegs between them, and dipping the resulting contraption into a tub of soapy water. The idea is that the soap bubbles that form between the pegs should trace out the minimum Steiner tree — that is, the minimum total length of line segments connecting the pegs, where the segments can meet at points other than the pegs themselves. Now, this is known to be an NP-hard optimization problem. So, it looks like Nature is solving NP-hard problems in polynomial time!

You might say there’s an obvious difficulty: the soap bubbles could get trapped in a local optimum that’s different from the global optimum. By analogy, a rock in a mountain crevice could reach a lower state of potential energy by rolling up first and then down … but is rarely observed to do so!

And if you said that, you’d be absolutely right. But that didn’t stop two guys a few years ago from writing a paper in which they claimed, not only that soap bubbles solve NP-complete problems in polynomial time, but that that fact proves P=NP! In debates about this paper on newsgroups, several posters raised the duh-obvious point that soap bubbles can get trapped at local optima. But then another poster opined that that’s just an academic “party line,” and that he’d be willing to bet that no one had actually done an experiment to prove it.

Long story short, I went to the hardware store, bought some glass plates, liquid soap, etc., and found that, while Nature does often find a minimum Steiner tree with 4 or 5 pegs, it tends to get stuck at local optima with larger numbers of pegs. Indeed, often the soap bubbles settle down to a configuration which is not even a tree (i.e. contains “cycles of soap”), and thus provably can’t be optimal.

The situation is similar for protein folding. Again, people have said that Nature seems to be solving an NP-hard optimization problem in every cell of your body, by letting the proteins fold into their minimum-energy configurations. But there are two problems with this claim. The first problem is that proteins, just like soap bubbles, sometimes get stuck in suboptimal configurations — indeed, it’s believed that’s exactly what happens with Mad Cow Disease. The second problem is that, to the extent that proteins do usually fold into their optimal configurations, there’s an obvious reason why they would: natural selection! If there were a protein that could only be folded by proving the Riemann Hypothesis, the gene that coded for it would quickly get weeded out of the gene pool.

So: quantum computers, soap bubbles, proteins … if we want to solve NP-complete problems in polynomial time in the physical world, what’s left? Well, we can try going to more exotic physics. For example, since we don’t yet have a quantum theory of gravity, people have felt free to speculate that if we did have one, it would give us an efficient way to solve NP-complete problems. For example, maybe the theory would allow closed timelike curves, which would let us solve NP-complete and even harder problems by (in some sense) sending the answer back in time to before we started.

In my view, though, it’s more likely that a quantum theory of gravity will do the exact opposite: that is, it will limit our computational powers, relative to what they would’ve been in a universe without gravity. To see why, consider one of the oldest “extravagant” computing proposals: the Zeno computer. This is a computer that runs the first step of a program in one second, the second step in half a second, the third step in a quarter second, the fourth step in an eighth second, and so on, so that after two seconds it’s run infinitely many steps. (It reminds me of the old joke about the supercomputer that was so fast, it could do an infinite loop in 2.5 seconds.)

Question from the floor: In what sense is this even a “proposal”?

Answer: Well, it’s a proposal in the sense that people actually write papers about it! (Google “hypercomputation.”) Whether they should be writing those papers a separate question…

Now, the Zeno computer strikes most computer scientists — me included — as a joke. But why is it a joke? Can we say anything better than that it feels absurd to us?

As it turns out, this question takes us straight into some of the frontier issues in theoretical physics. In particular, one of the few things physicists think they know about quantum gravity — one of the few things both the string theorists and their critics largely agree on — is that, at the so-called “Planck scale” of about 10-33 centimeters or 10-43 seconds, our usual notions of space and time are going to break down. As one manifestation of this, if you tried to build a clock that ticked more than about 1043 times per second, that clock would use so much energy that it would collapse to a black hole. Ditto for a computer that performed more than about 1043 operations per second, or for a hard disk that stored more than about 1069 bits per square meter of surface area. (Together with the finiteness of the speed of light and the exponential expansion of the universe, this implies that, contrary to what you might have thought, there is a fundamental physical limit on how much disk space Gmail will ever be able to offer its subscribers…)

To summarize: while I believe what I called the “No SuperSearch Principle” — that is, while I believe there are fundamental physical limits to efficient computer search — I hope I’ve convinced you that understanding why these limits exist takes us straight into some of the deepest issues in math and physics. To me that’s so much the better — since it suggests that not only are the limits correct, but (more importantly) they’re also nontrivial.

Thank you.

Does it come with a 14-Gyr warranty?

Sunday, August 19th, 2007

As many of you probably saw, John Tierney of the New York Times thinks there’s a ~50% chance we’re living in a computer simulation, having been persuaded by Nick Bostrom’s infamous simulation argument.

(This argument, incidentally, is something that occurred to me as a teenager, and I’m guessing to many others of nerdly leanings as well. I didn’t consider it a profound metaphysical discovery, just a sign I needed to get out more.)

Peter Woit feels strongly that debates about whether the universe is a computer are not science and therefore have no place in the Times science section. Robin Hanson retorts that “rather than complain that something is not ‘science,’ or not ‘philosophy,’ it is much better to just say more specifically what it is that you don’t like about it.” Peter Shor points out that if we’re living in a simulation, then the incompatibility of quantum mechanics with general relativity might simply be a bug, in which case the universe will crash when the first black hole evaporates.

As for me, I tend to side with Woody Allen: yes, the universe might be a simulation, but where else can you get a decent steak?

The last word, however, goes to Bender Bending Rodriguez of Futurama.

Bender: “If that stuff wasn’t real, how can I be sure anything is real? Is it not possible, nay, probable that my whole life is just a product of my or someone else’s imagination?”

Clerk: “No, get out. Next!”

(Click here for the audio clip.)

My Favorite Growth Rates

Sunday, August 12th, 2007

Update (8/17): Believe it or not, this blog actually led to something (scroll down to comment #52 if the link doesn’t work).

Intellectual whack-a-mole

Thursday, August 9th, 2007

Several readers have now written to me independently, asking for my reaction to the following paper:

An Optical Solution For The Traveling Salesman Problem
Tobias Haist and Wolfgang Osten

Abstract: We introduce an optical method based on white light interferometry in order to solve the well-known NP–complete traveling salesman problem. To our knowledge it is the first time that a method for the reduction of non–polynomial time to quadratic time has been proposed. We will show that this achievement is limited by the number of available photons for solving the problem. It will turn out that this number of photons is proportional to NN for a traveling salesman problem with N cities and that for large numbers of cities the method in practice therefore is limited by the signal–to–noise ratio. The proposed method is meant purely as a gedankenexperiment.

Look, this is really not hard. You really don’t need a world CompuCrackpotism expert to tell you what to think of this. If you read carefully, the authors were actually kind enough to explain themselves, right in the abstract, why their proposal doesn’t scale. (This, of course, is entirely to their credit, and puts them above ~98% of their colleagues in the burgeoning intersection of computer science, physics, and non-correctness.)

Hint: If the number of photons scales exponentially with N, and the photons have high enough energies that you can detect them, then the energy also scales exponentially with N. So by the Schwarzschild bound, the volume also scales exponentially with N; therefore, by locality, so does the time.