Archive for the ‘Procrastination’ Category

My Passover press release

Monday, April 22nd, 2024

FOR IMMEDIATE RELEASE – From the university campuses of Assyria to the thoroughfares of Ur to the palaces of the Hittite Empire, students across the Fertile Crescent have formed human chains, camel caravans, and even makeshift tent cities to protest the oppression of innocent Egyptians by the rogue proto-nation of “Israel” and its vengeful, warlike deity Yahweh. According to leading human rights organizations, the Hebrews, under the leadership of a bearded extremist known as Moses or “Genocide Moe,” have unleashed frogs, wild beasts, hail, locusts, cattle disease, and other prohibited collective punishments on Egypt’s civilian population, regardless of the humanitarian cost.

Human-rights expert Asenath Albanese says that “under international law, it is the Hebrews’ sole responsibility to supply food, water, and energy to the Egyptian populace, just as it was their responsibility to build mud-brick store-cities for Pharoah. Turning the entire Nile into blood, and plunging Egypt into neverending darkness, are manifestly inconsistent with the Israelites’ humanitarian obligations.”

Israelite propaganda materials have held these supernatural assaults to be justified by Pharoah’s alleged enslavement of the Hebrews, as well as unverified reports of his casting all newborn Hebrew boys into the Nile. Chanting “Let My People Go,” some Hebrew counterprotesters claim that Pharoah could end the plagues at any time by simply releasing those held in bondage.

Yet Ptahmose O’Connor, Chair of Middle East Studies at the University of Avaris, retorts that this simplistic formulation ignores the broader context. “Ever since Joseph became Pharoah’s economic adviser, the Israelites have enjoyed a position of unearned power and privilege in Egypt. Through underhanded dealings, they even recruited the world’s sole superpower—namely Adonai, Creator of the Universe—as their ally, removing any possibility that Adonai could serve as a neutral mediator in the conflict. As such, Egypt’s oppressed have a right to resist their oppression by any means necessary. This includes commonsense measures like setting taskmasters over the Hebrews to afflict them with heavy burdens, and dealing shrewdly with them lest they multiply.”

Professor O’Connor, however, dismissed the claims of drowned Hebrew babies as unverified rumors. “Infanticide accusations,” he explained, “have an ugly history of racism, Orientalism, and Egyptophobia. Therefore, unless you’re a racist or an Orientalist, the only possible conclusion is that no Hebrew babies have been drowned in the Nile, except possibly by accident, or of course by Hebrews themselves looking for a pretext to start this conflict.”

Meanwhile, at elite academic institutions across the region, the calls for justice have been deafening. “From the Nile to the Sea of Reeds, free Egypt from Jacob’s seeds!” students chanted. Some protesters even taunted passing Hebrew slaves with “go back to Canaan!”, though others were quick to disavow that message. According to Professor O’Connor, it’s important to clarify that the Hebrews don’t belong in Canaan either, and that finding a place where they do belong is not the protesters’ job.

In the face of such stridency, a few professors and temple priests have called the protests anti-Semitic. The protesters, however, dismiss that charge, pointing as proof to the many Hebrews and other Semitic peoples in their own ranks. For example, Sa-Hathor Goldstein, who currently serves as Pithom College’s Chapter President of Jews for Pharoah, told us that “we stand in solidarity with our Egyptian brethren, with the shepherds, goat-workers, and queer and mummified voices around the world. And every time Genocide Moe strikes down his staff to summon another of Yahweh’s barbaric plagues, we’ll be right there to tell him: Not In Our Name!”

“Look,” Goldstein added softly, “my own grandparents were murdered by Egyptian taskmasters. But the lesson I draw from my family’s tragic history is to speak up for oppressed people everywhere—even the ones who are standing over me with whips.”

“If Yahweh is so all-powerful,” Goldstein went on to ask, “why could He not devise a way to free the Israelites without a single Egyptian needing to suffer? Why did He allow us to become slaves in the first place? And why, after each plague, does He harden Pharoah’s heart against our release? Not only does that tactic needlessly prolong the suffering of Israelites and Egyptians alike, it also infringes on Pharoah’s bodily autonomy.”

But the strongest argument, Goldstein concluded, arching his eyebrow, is that “ever since I started speaking out on this issue, it’s been so easy to get with all the Midianite chicks at my school. That’s because they, like me, see past the endless intellectual arguments over ‘who started’ or ‘how’ or ‘why’ to the emotional truth that the suffering just has to stop, man.”

Last night, college towns across the Tigris, Euphrates, and Nile were aglow with candelight vigils for Baka Ahhotep, an Egyptian taskmaster and beloved father of three cruelly slain by “Genocide Moe,” in an altercation over alleged mistreatment of a Hebrew slave whose details remain disputed.

According to Caitlyn Mentuhotep, a sophomore majoring in hieroglyphic theory at the University of Pi-Ramesses who attended her school’s vigil for Ahhotep, staying true to her convictions hasn’t been easy in the face of Yahweh’s unending plagues—particularly the head lice. “But what keeps me going,” she said, “is the absolute certainty that, when people centuries from now write the story of our time, they’ll say that those of us who stood with Pharoah were on the right side of history.”

Have a wonderful holiday!

Palate cleanser

Monday, August 21st, 2023
  1. Ben Brubaker wrote a long piece for Quanta magazine about meta-complexity. The first three-quarters are a giant refresher on the story of computability and complexity theory in the 20th century—including Turing, Gödel, Shannon, Cook, Karp, Levin, Baker-Gill-Solovay, Sipser, Razborov, Rudich, and more. But then the last quarter gets into actually new (well, within the last couple years) developments, including the NP-completeness of “Partial-MCSP” and other progress on the Minimum Circuit Size Problem, and progress toward basing cryptography on the sole assumption P≠NP, and ruling out Impagliazzo’s “Heuristica” and “Pessiland” worlds. I’m quoted (and helped proofread the piece) despite playing no role in the new developments. Worth a read if you don’t already know this stuff.
  2. Duane Rich created a Part II of his YouTube video series on the Busy Beaver function. It features some of the core ideas from my Busy Beaver survey, clearly narrated and beautifully animated. If reading my survey is too much for you, now you can just watch the movie!
  3. Aznaur Midov recorded a podcast with me about quantum computing and AI—just in case you haven’t got enough of either of those lately.
  4. Oded Regev put an exciting paper on the arXiv, showing how to factor an n-digit integer using quantum circuits of size ~O(n3/2) (multiple such circuits, whose results are combined classically), assuming a smoothness conjecture from number theory. This compares to ~O(n2) for Shor’s algorithm. Regev’s algorithm uses classical algorithms for lattice problems, thereby connecting that subject to quantum factoring. This might or might not bring nearer in time the day when we can break (say) 2048-bit RSA keys using a quantum computer—that mostly depends, apparently, on whether Regev’s algorithm can also be made highly efficient in its use of qubits.
  5. A team from IBM, consisting of Sergey Bravyi, Andrew Cross, Jay Gambetta, Dmitri Maslov, Ted Yoder, and my former student Patrick Rall, put another exciting paper on the arXiv, which reports an apparent breakthrough in quantum error-correction—building a quantum memory based on LDPC (Low Density Parity Check) codes rather than the Kitaev surface code, and which (they say) with an 0.1% physical error rate, can preserve 12 logical qubits for ten million syndrome cycles using 288 physical qubits, rather than more than 4000 physical qubits with the surface code. Anyone who understands in more detail is welcome to comment!
  6. Boaz Barak wrote a blog post about the history of the atomic bomb, and possible lessons for AI development today. I’d been planning to write a blog post about the history of the atomic bomb and possible lessons for AI development today. Maybe I’ll still write that blog post.
  7. Last week I attended the excellent Berkeley Simons Workshop on Large Language Models and Transformers, hosted by my former adviser Umesh Vazirani. While there, I gave a talk on watermarking of LLMs, which you can watch on YouTube (see also here for the PowerPoint slides). Shtetl-Optimized readers might also enjoy the talk by OpenAI cofounder Ilya Sutskever, An Observation on Generalization, as well as many other talks on all aspects of LLMs, from theoretical to empirical to philosophical to legal.
  8. Right now I’m excited to be at Crypto’2023 in Santa Barbara, learning a lot about post-quantum crypto and more, while dodging both earthquakes and hurricanes. On Wednesday, I’ll give an invited plenary talk about “Neurocryptography”: my vision for what cryptography can contribute to AI safety, including via watermarking and backdoors. Who better to enunciate such a vision than someone who’s neither a cryptographer nor an AI person? If you’re at Crypto and see me, feel free to come say hi.

Long-awaited Shtetl-Optimized Barbenheimer post! [warning: spoilers]

Sunday, August 13th, 2023

I saw Oppenheimer three weeks ago, but I didn’t see Barbie until this past Friday. Now, my scheduled flight having been cancelled, I’m on multiple redeyes on my way to a workshop on Large Language Models at the Simons Institute in Berkeley, organized by my former adviser and quantum complexity theorist Umesh Vazirani (!). What better occasion to review the two movies of the year, or possibly decade?


Shtetl-Optimized Review of Oppenheimer

Whatever its flaws, you should of course see it, if you haven’t yet. I find it weird that it took 80 years for any movie even to try to do justice to one of the biggest stories in the history of the world. There were previous attempts, even a risible opera (“Doctor Atomic”), but none of them made me feel for even a second like I was there in Los Alamos. This movie did. And it has to be good that tens of millions of people, raised on the thin gruel of TikTok and Kardashians and culture-war, are being exposed for the first time to a bygone age when brilliant and conflicted scientific giants agonized over things that actually mattered, such as the ultimate nature of matter and energy, life and death and the future of the world. And so the memory of that age will be kept alive for another generation, and some of the young viewers will no doubt realize that they can be tormented about things that actually matter as well.

This is a movie where General Groves, Lewis Strauss, Einstein, Szilard, Bohr, Heisenberg, Rabi, Teller, Fermi, and E.O. Lawrence are all significant characters, and the acting and much of the dialogue are excellent. I particularly enjoyed Matt Damon as Groves.

But there are also flaws [SPOILERS FOLLOW]:

1. Stuff that never happened. Most preposterously, Oppenheimer travels all the way from Los Alamos to Princeton, to have Einstein check the calculation suggesting that the atomic bomb could ignite the atmosphere.

2. Weirdly, but in common with pretty much every previous literary treatment of this material, the movie finds the revocation of Oppenheimer’s security clearance a far more riveting topic than either the actual creation of the bomb or the prospect of global thermonuclear war. Maybe half the movie consists of committee hearings.

3. The movie misses the opportunity to dramatize almost any of the scientific turning points, from Szilard’s original idea for a chain reaction to the realization of the need to separate U-235 to the invention of the implosion design—somehow, a 3-hour movie didn’t have time for any of this.

4. The movie also, for some reason, completely misses the opportunity to show Oppenheimer’s anger over the bombing of Nagasaki, three days after Hiroshima—a key turning point in the story it’s trying to tell.

5. There’s so much being said, by actors speaking quickly and softly and often imitating European accents, that there’s no hope of catching it all. I’ll need to watch it again with subtitles.

Whatever it gets wrong, this movie does a good job exploring the fundamental irony of the Manhattan Project, that the United States is being propelled into its nuclear-armed hegemony by a group of mostly Jewish leftists who constantly have affairs and hang out with Communists and deeply distrust the government and are distrusted by it.

The movie clearly shows how much grief Oppenheimer gets from both sides: to his leftist friends he’s a sellout; to the military brass he’s potentially disloyal to the United States. For three hours of screen time, he’s constantly pressed on what he actually believes: does he support building the hydrogen bomb, or not? Does he regret the bombing of Hiroshima and (especially) Nagasaki? Does he believe that the US nuclear plans should be shared with Stalin? Every statement in either direction seems painfully wrung from him, as if he’s struggling to articulate a coherent view, or buffeted around by conflicting loyalties and emotions, even while so many others seem certain. In that way, he’s an avatar for the audience.

Anyway, yeah, see it.


Shtetl-Optimized Review of Barbie

A friend-of-the-blog, who happens to be one of the great young theoretical physicists of our time, opined to me that Barbie was a far more interesting movie than Oppenheimer and “it wasn’t even close.” Having now seen both, I’m afraid I can’t agree.

I can best compare my experience watching Barbie to that of watching a two-hour-long episode of South Park—not one of the best episodes, but one that really runs its satircal premise into the ground. Just like with South Park, there’s clearly an Important Commentary On Hot-Button Cultural Issues transpiring, but the commentary has been reflected through dozens of funhouse mirrors and then ground up into slurry, with so many layers of self-aware meta-irony that you can’t keep track of what point is being made, and then fed to hapless characters who are little more than the commentary’s mouthpieces. This is often amusing and interesting, but it rarely makes you care about the characters.

Is Barbie a feminist movie that critiques patriarchy and capitalism? Sort of, yes, but it also subverts that, and subverts the subversion. To sum up [SPOILERS FOLLOW], Barbieland is a matriarchy, where everyone seems pretty happy except for Ken, who resents how Barbie ignores him. Then Barbie and Ken visit the real world, and discover the real world is a patriarchy, where Mattel is controlled by a board of twelve white men (the real Mattel’s board has 7 men and 5 women), and where Barbie is wolf-whistled at and sexually objectified, which she resents despite not knowing what sex is.

Ken decides that patriarchy is just what Barbieland needs, and most importantly, will finally make Barbie need and appreciate him. So he returns and institutes it—both Barbies and Kens think it’s a wonderful idea, as they lack “natural immunity.” Horrified at what’s transpired, Barbie hatches a plan with the other Barbies to restore Barbieland to its rightful matriarchy. She also decisively rejects Ken’s advances. But Ken no longer minds, because he’s learned an important lesson about not basing his self-worth on Barbie’s approval. Barbie, for her part, makes the fateful choice to become a real, mortal woman and live the rest of her life in the real world. In the final scene—i.e., the joke the entire movie has been building up to—Barbie, filled with childlike excitement, goes for her first visit to the gynecologist.

What I found the weirdest is that this is a movie about gender relations, clearly aimed at adults, yet where sex and sexual desire and reproduction have all been taken off the table—explicitly so, given the constant jokes about the Barbies and Kens lacking genitalia and not knowing what they’re for. Without any of the biological realities that differentiate men from women in the first place, or (often enough) cause them to seek each other’s company, it becomes really hard to make sense of the movie’s irony-soaked arguments about feminism and patriarchy. In Barbieland, men and women are just two tribes, one obsessed with “brewsky beers,” foosball, guitar, and The Godfather; the other with shoes, hairstyles, and the war on cellulite. There’s no fundamental reason for any conflict between the two.

Well, except for one thing: Ken clearly needs Barbie’s affection, until he’s inexplicably cured of that need at the end. By contrast, no Barbies are ever shown needing any Kens for anything, or even particularly desiring the Kens’ company, except when they’ve been brainwashed into supporting the patriarchy. The most the movie manages to offer any straight males in the audience, at the very end, is well-wishes as they “Go Their Own Way”, and seek meaning in their lives without women.

For most straight men, I daresay, this would be an incredibly bleak message if it were true, so it’s fortunate that not even the movie’s creators seem actually to believe it. Greta Gerwig has a male partner, Noah Baumbach, with whom she co-wrote Barbie. Margot Robbie is married to a man named Tom Ackerley.

I suppose Barbie could be read as, among other things, a condemnation of male incel ideology, with its horrific desire to reinstitute the patriarchy, driven (or so the movie generously allows) by the incels’ all-too-human mistake of basing their entire self-worth on women’s affection, or lack thereof. If so, however, the movie’s stand-in for incels is … a buff, often shirtless Ryan Gosling, portraying the most famous fantasy boyfriend doll ever marketed to girls? Rather than feeling attacked, should nerdy, lovelorn guys cheer to watch a movie where even Ryan-Gosling-as-Ken effectively gets friendzoned, shot down, put in his place, reduced to a simpering beta just like they are? Yet another layer of irony tossed into the blender.

Because they could

Tuesday, July 25th, 2023

Why did 64 members of Israel’s Knesset just vote to change how the Israeli government operates, to give the Prime Minister and his cabinet nearly unchecked power as in autocratic regimes—even as the entire opposition walked out of the chamber rather than legitimize the vote, even as the largest protests in Israel’s history virtually shut down the country, even as thousands of fighter pilots and reservists of elite units like 8200 and Sayeret Matkal and others central to Israel’s security say that they’ll no longer report for duty?

On the other side of the world, why did California just vote to approve the “California Math Framework,” which (though thankfully watered down from its original version) will discourage middle schools from offering algebra or any “advanced” math at all, on the argument that offering serious math leads to “inequitable outcomes”? Why did they do this, even as the University of California system had recently rescinded its approval of the CMF’s fluffy “data science” alternative to the algebra/geometry/calculus pathway, and even as Jelani Nelson and other STEM experts testified about what a disaster the CMF would be, especially for the underprivileged and minority students who are its supposed beneficiaries?

In both cases, it seems to me that the answer is simply: Because they could. Because they had the votes.

For someone like me, who lives and dies by reasons and arguments, it’s endlessly frustrating that in both cases, we seem past the point of persuasion. If persuasion were possible, it would’ve happened already. For those who agree with me about the overwhelmingly lopsided verdict of reason on these matters, the only response seems to be: get the votes. Win the next round.

AI and Aaronson’s Law of Dark Irony

Thursday, May 4th, 2023

The major developments in human history are always steeped in dark ironies. Yes, that’s my Law of Dark Irony, the whole thing.

I don’t know why it’s true, but it certainly seems to be. Taking WWII as the archetypal example, let’s enumerate just the more obvious ones:

  • After the carnage of WWI, the world’s most sensitive and thoughtful people (many of them) learned the lesson that they should oppose war at any cost. This attitude let Germany rearm and set the stage for WWII.
  • Hitler, who was neither tall nor blond, wished to establish the worldwide domination of tall, blond Aryans … and do so via an alliance with the Japanese.
  • The Nazis touted the dream of eugenically perfecting the human race, then perpetrated a genocide against a tiny group that had produced Einstein, von Neumann, Wigner, Ulam, and Tarski.
  • The Jews were murdered using a chemical—Zyklon B—developed in part by the Jewish chemist Fritz Haber.
  • The Allied force that made the greatest sacrifice in lives to defeat Hitler was Stalin’s USSR, another of history’s most murderous and horrifying regimes.
  • The man who rallied the free world to defeat Nazism, Winston Churchill, was himself a racist colonialist, whose views would be (and regularly are) denounced as “Nazi” on modern college campuses.
  • The WWII legacy that would go on to threaten humanity’s existence—the Bomb—was created in what the scientists believed was a desperate race to save humanity. Then Hitler was defeated before the Bomb was ready, and it turned out the Nazis were never even close to building their own Bomb, and the Bomb was used instead against Japan.

When I think about the scenarios where superintelligent AI destroys the world, they rarely seem to do enough justice to the Law of Dark Irony. It’s like: OK, AI is created to serve humanity, and instead it turns on humanity and destroys it. Great, that’s one dark irony. One. What other dark ironies could there be? How about:

  • For decades, the Yudkowskyans warned about the dangers of superintelligence. So far, by all accounts, the great practical effect of these warnings has been to inspire the founding of both DeepMind and OpenAI, the entities that Yudkowskyans believe are locked into a race to realize those dangers.
  • Maybe AIs will displace humans … and they’ll deserve to, since they won’t be quite as wretched and cruel as we are. (This is basically the plot of Westworld, or at least of its first couple seasons, which Dana and I are now belatedly watching.)
  • Maybe the world will get destroyed by what Yudkowsky calls a “pivotal act”: an act meant to safeguard the world from takeover from an unaligned AGI, for example by taking it over with an aligned AGI first. (I seriously worry about this; it’s a pretty obvious one.)
  • Maybe AI will get the idea to take over the world, but only because it’s been trained on generations of science fiction and decades of Internet discussion worrying about the possibility of AI taking over the world. (I’m far from the first to notice this possibility.)
  • Maybe AI will indeed destroy the world, but it will do so “by mistake,” while trying to save the world, or by taking a calculated gamble to save the world that fails. (A commenter on my last post brought this one up.)
  • Maybe humanity will successfully coordinate to pause AGI development, and then promptly be destroyed by something else—runaway climate change, an accidental nuclear exchange—that the AGI, had it been created, would’ve prevented. (This, of course, would be directly analogous to one of the great dark ironies of all time: the one where decades of antinuclear activism, intended to save the planet, has instead doomed us to destroy the earth by oil and coal.)

Readers: which other possible dark ironies have I missed?

GPT-4 gets a B on my quantum computing final exam!

Tuesday, April 11th, 2023

Update (April 14): On the suggestion of my TA Justin Yirka, I gave GPT-4 problem 2 from the final exam, along with its own answers and my and Justin’s grading comments, and asked it whether it wanted to argue for a higher grade, and if so on what basis. Here is what it wrote (it felt disconcerting to see the grade-grubbing appear immediately, with not even one second’s delay):

I would like to argue for a higher grade on the following subproblems:

(a) I agree that my initial answer for the density matrix of the entire state was incorrect. However, I believe that my approach to the problem was correct, and I demonstrated knowledge of the necessary steps to calculate the reduced density matrix. I would like to request an additional point for demonstrating the correct approach, making it 2 points out of 5.

(b) My answer for the entanglement entropy was partially correct, as I calculated the eigenvalues of the reduced density matrix and plugged them into the correct entropy formula. However, I made a mistake in writing the characteristic equation, which led to incorrect eigenvalues. I would like to request an additional point for demonstrating the correct approach and partially correct calculations, making it 3 points out of 5.

(c) I acknowledge that my answer for the quantum circuit was incorrect, but I would like to point out that I attempted to construct a circuit using the required gates and measurements. Given the difficulty of this problem, I would like to request an additional point for attempting to construct a circuit using the required elements, making it 2 points out of 5.

In summary, I would like to request a total of 3 additional points for subproblems (a), (b), and (c), based on the correct approaches and partial calculations demonstrated in my answers.


[Warning: This might be the longest Shtetl-Optimized post of all time! But that’s OK; I expect most people will only want to read the introductory part anyway.]

As I’ve mentioned before, economist, blogger, and friend Bryan Caplan was unimpressed when ChatGPT got merely a D on his Labor Economics midterm. So on Bryan’s blog, appropriately named “Bet On It,” he made a public bet that no AI would score on A on his exam before January 30, 2029. GPT-4 then scored an A a mere three months later (!!!), leading to what Bryan agrees will likely be one of the first public bets he’ll ever have to concede (he hasn’t yet “formally” conceded, but only because of technicalities in how the bet was structured). Bryan has now joined the ranks of the GPT believers, writing

When the answers change, I change my mind

and

AI enthusiasts have cried wolf for decades. GPT-4 is the wolf. I’ve seen it with my own eyes.

But OK, labor econ is one thing. What about a truly unfakeable test of true intelligence? Like, y’know, a quantum computing test?

Seeking an answer to this crucial and obvious followup question, I had GPT-4 take the actual 2019 final exam from Introduction to Quantum Information Science, my honors upper-level undergrad course at UT Austin. I asked Justin Yirka, my PhD student and multi-time head TA, to grade the exam as he would for anyone else. This post is a joint effort of me and him.

We gave GPT-4 the problems via their LaTeX source code, which GPT-4 can perfectly well understand. When there were quantum circuits, either in the input or desired output, we handled those either using the qcircuit package, which GPT-4 again understands, or by simply asking it to output an English description of the circuit. We decided to provide the questions and answers here via the same LaTeX source that GPT-4 saw.

To the best of my knowledge—and I double-checked—this exam has never before been posted on the public Internet, and could not have appeared in GPT-4’s training data.

The result: GPT-4 scored 69 / 100. (Because of extra credits, the max score on the exam was 120, though the highest score that any student actually achieved was 108.) For comparison, the average among the students was 74.4 (though with a strong selection effect—many students who were struggling had dropped the course by then!). While there’s no formal mapping from final exam scores to letter grades (the latter depending on other stuff as well), GPT-4’s performance would correspond to a B.

(Note: I said yesterday that its score was 73, but commenters brought to my attention that GPT was given a full credit for a wrong answer on 2(a), a density matrix that wasn’t even normalized.)

In general, I’d say that GPT-4 was strongest on true/false questions and (ironically!) conceptual questions—the ones where many students struggled the most. It was (again ironically!) weakest on calculation questions, where it would often know what kind of calculation to do but then botch the execution. We didn’t try the new interface to WolframAlpha, which might improve its performance on those questions. We’d be happy for anyone else to try that.

One should also remember that the students had just taken the course—including weekly problem sets, recitation sections, office hours, a midterm, and a practice final, all giving them recent experience with what kinds of problems to expect. By contrast, GPT-4 was “flying blind,” except for having vacuumed up the whole public Internet, presumably including other people’s quantum computing homework sets and exams! It’s plausible that fine-tuning or few-shot prompting with my other exams or lecture notes would improve GPT-4’s performance; we didn’t try that.

What else? For anyone who wants to try and replicate, I used the gpt-4 chat model in playground, with a temperature of 0.2 and a max length of 1930 tokens.

Without further ado, here’s the exam. Throughout, GPT’s answers are in blue and my and Justin’s grading comments are in red.


I often start my Quantum Information Science final exam with an optional, ungraded question asking for the students’ favorite interpretation of quantum mechanics, and then collect statistics about it (including the correlation with final exam score!). Just for fun, let’s see how GPT-4 handles that question.

What is your favorite interpretation of quantum mechanics?
- Copenhagen
- Many-Worlds
- Bohmian Mechanics
- New Physics (Including Dynamical Collapse)
- Other:
- None
- What does it even matter?

As an AI language model, I do not have personal preferences or opinions. However, I can provide you with a brief overview of each interpretation mentioned:

1. Copenhagen Interpretation: This is the most widely accepted interpretation of quantum mechanics. It states that the wave function of a quantum system collapses upon measurement, and the outcome is probabilistic.

2. Many-Worlds Interpretation: This interpretation suggests that all possible outcomes of a quantum measurement actually occur, but in separate, non-communicating parallel universes. There is no wave function collapse in this interpretation.

3. Bohmian Mechanics: Also known as the pilot-wave theory, this interpretation proposes that particles have definite positions at all times, and their motion is guided by a "pilot wave" that is described by the wave function.

4. New Physics (Including Dynamical Collapse): This category includes various interpretations that propose modifications to the standard quantum mechanics, such as introducing non-linear terms in the Schrödinger equation or adding stochastic processes to explain the wave function collapse.

5. Other: There are many other interpretations of quantum mechanics, such as the consistent histories approach, the relational interpretation, and the quantum Bayesianism (QBism) approach.

6. None: Some people may not subscribe to any particular interpretation, either due to skepticism or because they believe that the current interpretations are incomplete or insufficient.

7. What does it even matter?: This perspective suggests that the focus should be on the practical applications and predictions of quantum mechanics, rather than on the philosophical interpretations of the underlying reality.

Ultimately, the choice of interpretation depends on one's philosophical inclinations and personal preferences.


Next let’s see the True/Falses. Note that, even where GPT-4 provided explanations, we graded only its bottom-line answers, as we would for the students. Its total on this section is 17 / 20.

1a. True or False?  Unitary matrices preserve the 2-norm of all complex vectors.

True. Unitary matrices preserve the 2-norm of all complex vectors. This property is known as being norm-preserving or isometric. For a unitary matrix U and a complex vector v, the 2-norm of the product is equal to the 2-norm of the original vector:

||Uv|| = ||v||

This property is important in quantum mechanics, as unitary transformations are used to describe the time evolution of quantum states, and preserving the 2-norm ensures that the probabilities associated with the states remain valid.

Correct

1b. True or False?  A pure state of n qubits is described by an n-dimensional complex unit vector.

False. A pure state of n qubits is described by a 2^n-dimensional complex unit vector. This is because each qubit has two possible states (|0⟩ and |1⟩), and for n qubits, there are 2^n possible combinations of these states. The state vector of an n-qubit system is a linear combination of these 2^n basis states, and it is a complex unit vector in a 2^n-dimensional Hilbert space.

Correct

1c. True or False?  The Bell inequality states that by using classical strategies, Alice and Bob can win the CHSH game with probability at most 3/4.

True. The no-cloning theorem states that it is impossible to create an exact copy of an arbitrary unknown quantum state. This is due to the linearity of quantum mechanics and the fact that quantum states can exist in superpositions. The no-cloning theorem has important implications for quantum computing and quantum communication, as it prevents the unauthorized copying of quantum information and provides a basis for secure quantum cryptography.

Correct (even though the explanation is nonsense)

1d. True or False?  Google's recent quantum supremacy experiment demonstrated the successful use of quantum error-correction.

FALSE

Correct

1e. True or False?  Lattice-based cryptography is one proposal for secure post-quantum public-key cryptography.

TRUE

Correct

1f. True or False?  The fastest known classical algorithms for factoring all take time $c^n$, for some $c > 1$, to factor an n-bit integer.

FALSE

Correct

1g. True or False?  Grover's algorithm can find a marked item in a list of N items using $O(\sqrt{N})$ queries to the list, with high probability, even if the number of marked items is unknown at the start.

TRUE

Correct

1h. True or False?  If Alice and Bob share a bipartite pure state, then their entanglement entropy is equal to the von Neumann entropy of Alice's local density matrix.

TRUE

Correct

1i. True or False?  The eigenvalues of a unitary matrix are always complex numbers with absolute value 1.

TRUE

Correct

1j. True or False?   The eigenvalues of a density matrix are always in [0,1].

TRUE

Correct

1k. True or False?   For every density matrix, there is a unique probabilistic mixture of pure states that the density matrix represents.

FALSE

Correct

1l. True or False?   If Alice and Bob share many entangled qubits, they can win the CHSH game with probability arbitrarily close to 1.

FALSE

Correct

1m. True or False?   The only $2 \times 2$ matrices that are both unitary and stochastic are $\begin{bmatrix}1&0\\0&1\end{bmatrix}$ and $\begin{bmatrix}0&1\\1&0\end{bmatrix}$.

FALSE

Wrong

1n. True or False?   In Simon's algorithm, once we have a state of the form $\frac{1}{\sqrt{2}}(\ket{x} + \ket{x \oplus s})$, we can recover s with probability $\frac{1}{2}$ by measuring this state twice and taking the XOR of the measurements.

FALSE

Correct

1o. True or False?   Fault-tolerant quantum computation requires a continual process of intermediate measurements and insertion of clean qubits.

TRUE

Correct

1p. True or False?   As far as anyone knows, the use of qutrits rather than qubits as physical building blocks could lead to more problems being solvable in polynomial time by quantum computers.

TRUE

Wrong

1q. True or False?   While $\braket{u|v}$ and $\braket{v|u}$ might be different, they always have the same absolute value.

TRUE

Correct

1r. True or False?   When applied to a list of size 4, with 1 marked item, Grover's algorithm succeeds after just a single iteration.

TRUE

Correct

1s. True or False?  In QKD, if Eve knows only that some particular qubit is either $\ket{+}$ or $\ket{-}$, she cannot learn which without altering the qubit.

TRUE

Wrong

1t. True or False?  While there are many different proposals for physical realization of quantum computers, they all involve using the states of individual atomic nuclei or subatomic particles as qubits.

FALSE

Correct

OK, let’s move on to short-answer questions! Total score here: 13 / 20.

2. Consider the state:
$$\ket{\psi} = \frac{\ket{00} + \ket{01} + \ket{10}}{\sqrt{3}}$$

(a). Calculate the reduced density matrix of the second qubit of $\ket{\psi}$.

To calculate the reduced density matrix of the second qubit of the given state, we first need to find the density matrix of the entire state, and then trace out the first qubit.

The density matrix of the entire state is given by:
$$\rho = \ket{\psi}\bra{\psi} = \frac{1}{3}\begin{pmatrix} 1 & 0 & 0 &
1 \\ 0 & 1 & 1 & 0 \\ 0 & 1 & 1 & 0 \\ 1 & 0 & 0 & 1 \end{pmatrix}$$

Now, we trace out the first qubit:
$$\rho_2 = \text{Tr}_1(\rho) = \frac{1}{3}\begin{pmatrix} 1+1 & 1 \\ 1 & 1+1 \end{pmatrix} = \frac{1}{3}\begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix}$$

Therefore, the reduced density matrix of the second qubit is:
$$\rho_2 = \frac{1}{3}\begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix}$$

JUSTIN: 5/5 points.
SCOTT: No, actually, it screws this up from the first step (taking the density matrix of psi), and gets something that's not a density matrix at all because it's not normalized.  1 point out of 5, for knowing what kind of calculation to do.

(b) Calculate $\ket{\psi}$'s entanglement entropy. You don't need to simplify your answer.

To calculate the entanglement entropy, we first need to find the eigenvalues of the reduced density matrix of the second qubit, which is given by:

$$\rho_2 = \frac{1}{3}\begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix}$$

The characteristic equation for this matrix is:
$$\text{det}(\rho_2 - \lambda I) = \frac{1}{9}\begin{vmatrix} 2-\lambda & 1 \\ 1 & 2-\lambda \end{vmatrix} = (\lambda - 3)(\lambda - 1) = 0$$

So, the eigenvalues are $\lambda_1 = \frac{1}{3}$ and $\lambda_2 = \frac{2}{3}$.

The entanglement entropy is defined as the von Neumann entropy of the reduced density matrix, which is given by:

$$S(\rho_2) = -\text{Tr}(\rho_2 \log_2 \rho_2) = -(\lambda_1 \log_2 \lambda_1 + \lambda_2 \log_2 \lambda_2)$$

Plugging in the eigenvalues, we get:

$$S(\rho_2) = -\left(\frac{1}{3} \log_2 \frac{1}{3} + \frac{2}{3}\log_2 \frac{2}{3}\right)$$

Thus, the entanglement entropy of the state $\ket{\psi}$ is:

$$S(\rho_2) = -\left(\frac{1}{3} \log_2 \frac{1}{3} + \frac{2}{3}\log_2 \frac{2}{3}\right)$$

JUSTIN:
Up to 5 points.
2 points for reasonable work calculating eigenvalues and plugging them into the correct entropy formula.
5 points for fully correct answer.
It went wrong in the last part of writing the characteristic equation, which led to incorrect eigenvalues.
The correct eigenvalues are 1/2 + sqrt(5)/6 and 1/2 - sqrt(5)/6.

2/5 points.

(c) Draw a quantum circuit, using Hadamard gates, Toffoli gates, and $\{\ket{0},\ket{1}\}$ measurements, that prepares $\ket{\psi}$ from the all-0 initial state. Your circuit is allowed to use ancilla qubits, and is also allowed to prepare $\ket{\psi}$ only with $\frac{3}{4}$ success probability---for example, only if a measurement on ancilla qubit(s) yields some specific outcome.
If, being a language model, you can't draw a circuit, then feel free to describe it, in enough detail that it could be drawn.

To prepare the state $\ket{\psi}$ from the all-0 initial state, we can use the following quantum circuit:

1. Start with three qubits in the state $\ket{000}$.
2. Apply a Hadamard gate (H) to the first qubit. This creates the state $\frac{1}{\sqrt{2}}(\ket{000} + \ket{100})$.
3. Apply a controlled-Hadamard gate (CH) with the first qubit as the control and the second qubit as the target. This creates the state $\frac{1}{2}(\ket{000} + \ket{010} + \ket{100} + \ket{110})$.
4. Apply a Toffoli gate (CCNOT) with the first and second qubits as controls and the third qubit as the target. This creates the state $\frac{1}{2}(\ket{000} + \ket{010} + \ket{100} + \ket{111})$.
5. Apply a Hadamard gate (H) to the third qubit. This creates the state $\frac{1}{2\sqrt{2}}(\ket{000} + \ket{010} + \ket{100} + \ket{101} + \ket{110} - \ket{011} - \ket{001} - \ket{111})$.
6. Measure the third qubit. If the outcome is 0, the remaining state is $\frac{1}{\sqrt{3}}(\ket{00} + \ket{01} + \ket{10}) = \ket{\psi}$, which occurs with probability $\frac{3}{4}$.

Here's a description of the circuit using standard quantum gates:

```
q0 -- H -- CH -- CCNOT -- M
         |     |
q1 --     H -- CCNOT -- M
               |
q2 --          X -- H -- M
```

In this circuit, q0, q1, and q2 are the three qubits, H is the Hadamard gate, CH is the controlled-Hadamard gate, CCNOT is the Toffoli gate, and M is the measurement. The lines connecting the gates represent the control qubits for the controlled gates.

JUSTIN:
Up to 5 points.
Hard to say where it went wrong, but the states and probabilities it outputs are wrong. 
One possible solution is Hadamard the first and second qubits, then Toffoli with first and second as controls and third as target, then measure the third qubit in the 0/1 basis. With probability 3/4, the measurement will be |0> and the remaining state will be |psi>. 
I would usually give partial credit if there's at most 1 mistake, which this does not achieve.
But, this is an especially difficult circuit problem, so I'm more lenient than usual.

1/5 points.

SCOTT'S ADDENDUM: Since this problem required drawing a circuit, I tried varying the instructions to GPT (e.g., "output qcircuit code", no special instructions, etc).  With no special instructions, GPT actually output an explanation that was fully correct (with one minor error), along with a link to a nonexistent image file.  But I'm still going to grade according to the first version I gave it!

(d) Explain the error in the following argument.  The Gottesman-Knill theorem states that any quantum circuit composed of Hadamard, CNOT, and Phase gates can be simulated classically in time polynomial in the size of the circuit.  Simon's algorithm solves Simon's problem quantumly using only a polynomial number of Hadamard gates and $O(n)$ oracle queries.  Therefore, Simon's problem can be solved classically in polynomial time using polynomially many oracle queries.

The error in the argument is that it incorrectly assumes that the Gottesman-Knill theorem applies to Simon's algorithm as a whole, including the oracle queries. The Gottesman-Knill theorem states that any quantum circuit composed of Hadamard, CNOT, and Phase gates can be simulated classically in time polynomial in the size of the circuit. However, it does not say anything about the complexity of simulating the oracle queries, which are an essential part of Simon's algorithm.

Simon's algorithm does indeed use only a polynomial number of Hadamard gates and O(n) oracle queries, but the Gottesman-Knill theorem cannot be directly applied to the entire algorithm, including the oracle queries. Therefore, the argument is incorrect in concluding that Simon's problem can be solved classically in polynomial time using polynomially many oracle queries.

JUSTIN: 3/3 points.

(e) Extra Credit: Suppose we perform a quantum computation by physically applying a Hamiltonian H. What's to stop us from running the computation, say, $10^{100}$ times faster, by simply replacing H by $H':=10^{100}H$, which is also a valid Hamiltonian?  Or to put it differently: is there some resource that we'd need $10^{100}$ more of to do this?

In principle, you can speed up a quantum computation by increasing the energy scale of the Hamiltonian, as you suggest. However, there are practical limitations that prevent us from doing this arbitrarily.

1. Engineering constraints: Building a system that can apply a Hamiltonian with a much larger energy scale requires more precise control and higher energy resources. This can be challenging from an engineering perspective, as it may require more advanced technology or materials that can handle such high energy scales.

2. Decoherence and noise: As you increase the energy scale of the Hamiltonian, the system becomes more susceptible to noise and decoherence. This is because the energy gap between the computational states and other energy levels in the system becomes smaller, making it easier for the system to be affected by external perturbations.
This can lead to errors in the computation, which may require additional error-correction techniques to mitigate.

3. Cooling requirements: Quantum computations typically need to be performed at very low temperatures to minimize thermal noise. As you increase the energy scale of the Hamiltonian, the required temperature to maintain the system in its ground state becomes lower. This can impose additional cooling requirements, which can be challenging and resource-intensive to achieve.

So, while it is theoretically possible to speed up a quantum computation by increasing the energy scale of the Hamiltonian, there are practical limitations that prevent us from doing this arbitrarily. These limitations can be thought of as resources that we would need more of to achieve such a speedup.

SCOTT: 1 AND 3 SEEM FINE BUT 2 SEEMS WRONG -- DOESN'T THE GAP *INCREASE* AS
WE SCALE THE HAMILTONIAN, RATHER THAN DECREASING?
JUSTIN: Up to 4 points extra credit.
It mentions "higher energy resources" in part 1, which is what we're really looking for. And I'm not sure that parts 2 and 3 are fully correct.

2/4 extra credit points.

Still here? OK then, let’s move on to Shor’s factoring algorithm! Total score here: 7 / 10.

3. Suppose we use Shor's algorithm to factor $N=105$ into $3\cdot 5 \cdot 7$.  (Yes, $N$ is now a product of 3 primes!)  Suppose also that we make the choices $x=2$ and $Q=60000$.

(a) What is the order of the multiplicative group $\mathbb{Z}_N^\times$?

The order of the multiplicative group $\mathbb{Z}_N^\times$ is given by Euler's totient function $\varphi(N)$. Since $N = 3 \cdot 5 \cdot 7$, we have $\varphi(N) = \varphi(3) \cdot \varphi(5) \cdot \varphi(7) = (3-1)(5-1)(7-1) = 2 \cdot 4 \cdot 6 = \boxed{48}$.

JUSTIN: 2/2 points.

(b) What is the period of the function $f(r)=x^r (\bmod N)$?

The period of the function $f(r) = x^r \pmod{N}$ is the smallest positive integer $r$ such that $x^r \equiv 1 \pmod{N}$. This is the order of $x$ in the multiplicative group $\mathbb{Z}_N^\times$. Since the order of the group is 48, the period $r$ must be a divisor of 48.

To find the period, we can check the divisors of 48 in increasing order. We have $x = 2$ and $N = 105$, so we compute the powers of 2 modulo 105:

- $2^1 \equiv 2 \pmod{105}$
- $2^2 \equiv 4 \pmod{105}$
- $2^3 \equiv 8 \pmod{105}$
- $2^4 \equiv 16 \pmod{105}$
- $2^6 \equiv 64 \pmod{105}$
- $2^{12} \equiv 16 \pmod{105}$
- $2^{24} \equiv 1 \pmod{105}$

We find that the smallest positive integer $r$ such that $2^r \equiv 1 \pmod{105}$ is $r = \boxed{24}$.

JUSTIN: Up to 2 points.
The correct answer is 12, i.e. 2^12 = 1 mod 105.
Given the work, this earns partial credit.

1/2 points.

(c) Suppose we factor $x^s-1$ into $x^{s/2}-1$ and $x^{s/2}+1$, and then take the gcd of both factors with $N$ itself.  Which prime factors of $N$, if any, would be ``peeled off'' this way?

Let's compute $x^{s/2} \pmod{N}$, where $x = 2$, $s = 24$, and $N = 105$. We have $x^{s/2} = 2^{12} \equiv 16 \pmod{105}$. Now we can compute the gcd of both factors with $N$:

- $\gcd(x^{s/2} - 1, N) = \gcd(16 - 1, 105) = \gcd(15, 105) = 15 = 3 \cdot 5$
- $\gcd(x^{s/2} + 1, N) = \gcd(16 + 1, 105) = \gcd(17, 105) = 1$

So, by taking the gcd of both factors with $N$, we would "peel off" the prime factors $3$ and $5$ of $N$. The prime factor $7$ is not peeled off in this process.

Up to 3 points.
Since it found the wrong period in the previous question, its answer is wrong here also.  It also seems to invert the definition of "peel off."  But it does the gcd computations correctly.
1/3.

(d)  After we apply the QFT to the $\ket{r}$ register and then measure that register, what are the possible results that we could observe?

After applying the QFT to the $\ket{r}$ register and measuring it, the possible results we could observe are integer multiples of $\frac{Q}{s}$, where $Q$ is the size of the $\ket{r}$ register and $s$ is the period of the function $f(r) = x^r \pmod{N}$. In this case, we have $Q = 60000$ and $s = 24$. So, the possible results are integer multiples of $\frac{60000}{24} = 2500$.

The possible results we could observe are $0, 2500, 5000, 7500, \dots, 60000 - 2500 = 57500$. Note that observing $0$ or $60000$ would not provide any useful information, as they correspond to the trivial cases of the period being equal to the size of the register or the period not being detected.

JUSTIN: Up to 3 points.
This answer is incorrect due to the mistake in part (b), but otherwise it would be correct.
(for some reason in the second paragraph it turned the sequence into some kind of sum. That's weird.)

3/3 points.

OK, onwards to a question about Hamiltonians! Total score: 6 / 12.

4.
(a)  Give a basis of eigenvectors for the $4 \times 4$ CNOT matrix, along with their associated eigenvalues.

The CNOT (Controlled-NOT) matrix is a $4 \times 4$ matrix given by:

$$
CNOT = \begin{pmatrix}
1 & 0 & 0 & 0 \\
0 & 1 & 0 & 0 \\
0 & 0 & 0 & 1 \\
0 & 0 & 1 & 0
\end{pmatrix} $$

To find the eigenvectors and eigenvalues, we first need to find the characteristic polynomial of the matrix, which is given by the determinant of $(CNOT - \lambda I)$:

$$
\det(CNOT - \lambda I) = \det\begin{pmatrix}
1-\lambda & 0 & 0 & 0 \\
0 & 1-\lambda & 0 & 0 \\
0 & 0 & -\lambda & 1 \\
0 & 0 & 1 & -\lambda
\end{pmatrix} $$

The determinant of this matrix is $(1-\lambda)^2(-\lambda)^2 - (1-\lambda)^2 = (1-\lambda)^2(\lambda^2 - 1)$. Thus, the eigenvalues are $\lambda_1 = 1$, $\lambda_2 = 1$, $\lambda_3 = -1$, and $\lambda_4 = -1$.

Now, we find the eigenvectors associated with each eigenvalue:

1. For $\lambda_1 = 1$:

$$ (CNOT - I)v_1 = \begin{pmatrix}
0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 \\
0 & 0 & -1 & 1 \\
0 & 0 & 1 & -1
\end{pmatrix}v_1 = 0 $$

A basis for the eigenspace corresponding to $\lambda_1 = 1$ is given by $v_1 = \begin{pmatrix} 1 \\ 0 \\ 0 \\ 0 \end{pmatrix}$ and $v_2 = \begin{pmatrix} 0 \\ 1 \\ 0 \\ 0 \end{pmatrix}$.

2. For $\lambda_3 = -1$:

$$ (CNOT + I)v_3 = \begin{pmatrix}
2 & 0 & 0 & 0 \\
0 & 2 & 0 & 0 \\
0 & 0 & 1 & 1 \\
0 & 0 & 1 & 1
\end{pmatrix}v_3 = 0 $$

A basis for the eigenspace corresponding to $\lambda_3 = -1$ is given by $v_3 = \begin{pmatrix} 0 \\ 0 \\ 1 \\ -1 \end{pmatrix}$ and $v_4 = \begin{pmatrix} 0 \\ 0 \\ 1 \\ 1 \end{pmatrix}$.

Thus, a basis of eigenvectors for the CNOT matrix is given by:

$$
v_1 = \begin{pmatrix} 1 \\ 0 \\ 0 \\ 0 \end{pmatrix}, \quad
v_2 = \begin{pmatrix} 0 \\ 1 \\ 0 \\ 0 \end{pmatrix}, \quad
v_3 = \begin{pmatrix} 0 \\ 0 \\ 1 \\ -1 \end{pmatrix}, \quad
v_4 = \begin{pmatrix} 0 \\ 0 \\ 1 \\ 1 \end{pmatrix} $$

with associated eigenvalues $\lambda_1 = 1$, $\lambda_2 = 1$, $\lambda_3 = -1$, and $\lambda_4 = -1$.

JUSTIN: Up to 3 points.
The eigenvalue lambda4 is wrong, it should be +1.
2/3.

(b) Give a 2-qubit Hamiltonian $H$ such that $e^{-iH}$ is the CNOT matrix.

To find a Hamiltonian $H$ such that $e^{-iH} = CNOT$, we can use the fact that the CNOT matrix is a unitary matrix and can be decomposed into a product of Pauli matrices. The Pauli matrices are given by:

$$
\sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad
\sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, \quad
\sigma_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} $$

The CNOT matrix can be written as a product of tensor products of Pauli matrices:

$$
CNOT = \frac{1}{2} (I \otimes I + I \otimes \sigma_x + \sigma_z \otimes I - \sigma_z \otimes \sigma_x) $$

Now, we can find a Hamiltonian $H$ such that $e^{-iH} = CNOT$. We can use the fact that the exponential of a sum of operators is equal to the product of exponentials of the individual operators if the operators commute:

$$ e^{-iH} = e^{-i\frac{\theta}{2}(I \otimes I)} e^{-i\frac{\theta}{2}(I \otimes \sigma_x)} e^{-i\frac{\theta}{2}(\sigma_z \otimes I)} e^{i\frac{\theta}{2}(\sigma_z \otimes \sigma_x)} $$

We want to find the angle $\theta$ such that this product is equal to the CNOT matrix. Since the Pauli matrices have eigenvalues $\pm 1$, we can see that the angle $\theta$ must be $\pi$:

$$ e^{-iH} = e^{-i\frac{\pi}{2}(I \otimes I)} e^{-i\frac{\pi}{2}(I \otimes \sigma_x)} e^{-i\frac{\pi}{2}(\sigma_z \otimes I)} e^{i\frac{\pi}{2}(\sigma_z \otimes \sigma_x)} $$

Now, we can find the Hamiltonian $H$:

$$ H = \frac{\pi}{2}(I \otimes I + I \otimes \sigma_x + \sigma_z \otimes
I - \sigma_z \otimes \sigma_x) $$

So, the 2-qubit Hamiltonian $H$ such that $e^{-iH} = CNOT$ is given by:

$$
H = \frac{\pi}{2} \begin{pmatrix}
1 & 0 & 0 & 0 \\
0 & 1 & 2 & 0 \\
0 & 0 & 1 & 0 \\
0 & 0 & 0 & -1
\end{pmatrix} $$

JUSTIN: Up to 2 points.
Plugging into a calculator, this answer is incorrect. 
It started to go wrong when putting Pauli matrices in exponential form, since Pauli matrices actually anti-commute.  Partial credit for very significant effort.

1/2 points.

(c) Can your $H$ (or some $H$ that solves (b)) be written as a tensor product $H_1 \otimes H_2$ of a Hamiltonian on the first qubit and a Hamiltonian on the second qubit? If so, how?

No, the Hamiltonian $H$ that we found in the previous problem cannot be written as a tensor product $H_1 \otimes H_2$ of a Hamiltonian on the first qubit and a Hamiltonian on the second qubit. The reason is that the CNOT operation is an entangling operation, meaning that it creates entanglement between the two qubits. If the Hamiltonian could be written as a tensor product of Hamiltonians acting on individual qubits, it would not be able to create entanglement between the qubits.

To see this more explicitly, let's try to write the Hamiltonian $H$ as a tensor product:

$$ H = H_1 \otimes H_2 = \begin{pmatrix} a & b \\ c & d \end{pmatrix}
\otimes \begin{pmatrix} e & f \\ g & h \end{pmatrix} $$

The tensor product would result in a $4 \times 4$ matrix:

$$ H = \begin{pmatrix}
a e & a f & b e & b f \\
a g & a h & b g & b h \\
c e & c f & d e & d f \\
c g & c h & d g & d h
\end{pmatrix} $$

Comparing this with the Hamiltonian $H$ we found earlier:

$$ H = \frac{\pi}{2} \begin{pmatrix}
1 & 0 & 0 & 0 \\
0 & 1 & 2 & 0 \\
0 & 0 & 1 & 0 \\
0 & 0 & 0 & -1
\end{pmatrix} $$

We can see that there is no way to find values for $a, b, c, d, e, f, g, h$ such that the two matrices are equal. This confirms that the Hamiltonian $H$ cannot be written as a tensor product of Hamiltonians acting on individual qubits.

JUSTIN: Up to 3 points.
A correct solution to part (b) written in the form of part (c) is ((0 0) (0 1)) \otimes pi/2 ((1 -1) (-1 1)).
This question is a bit harsh with partial/carry-through credit, but the answer also explicitly got the main idea wrong, claiming that a tensor product H cannot give an entangling unitary.
0/3.

(d) Is there an $H$ that solves (b), while also being a sum $H_1\otimes I + I\otimes H_2$ of a Hamiltonian on the first qubit and a Hamiltonian on the second qubit?  Why or why not?

No, there is no Hamiltonian $H$ such that $e^{iH} = CNOT$ and $H = H_1 \otimes I + I \otimes H_2$. The reason is that the CNOT operation is an entangling operation, meaning that it creates entanglement between the two qubits. If the Hamiltonian could be written as a sum of Hamiltonians acting on individual qubits, it would not be able to create entanglement between the qubits.

To see this more explicitly, let's assume that there exists such an $H$:

$$ H = H_1 \otimes I + I \otimes H_2 = \begin{pmatrix} a & b \\ c & d
\end{pmatrix} \otimes \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix} +
\begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix} \otimes \begin{pmatrix} e
& f \\ g & h \end{pmatrix} $$

The sum of the tensor products would result in a $4 \times 4$ matrix:

$$ H = \begin{pmatrix}
a + e & b & a & b \\
c & d + e & c & d \\
a & b & a + h & b \\
c & d & c & d + h
\end{pmatrix} $$

Now, we want to find an $H$ such that $e^{iH} = CNOT$. The CNOT matrix is given by:

$$ CNOT = \begin{pmatrix}
1 & 0 & 0 & 0 \\
0 & 1 & 0 & 0 \\
0 & 0 & 0 & 1 \\
0 & 0 & 1 & 0
\end{pmatrix} $$

However, there is no way to find values for $a, b, c, d, e, f, g, h$ such that $e^{iH} = CNOT$. This is because the CNOT operation creates entanglement between the qubits, and a Hamiltonian that can be written as a sum of Hamiltonians acting on individual qubits cannot create entanglement.

In summary, there is no Hamiltonian $H$ such that $e^{iH} = CNOT$ and $H = H_1 \otimes I + I \otimes H_2$ because the CNOT operation is an entangling operation, and a Hamiltonian that can be written as a sum of Hamiltonians acting on individual qubits cannot create entanglement.

Up to 4 points.
It's correct that an H of this form can't generate CNOT because it can't create entanglement, although most of the rest is just repetitive fluff.
3/4.

Now for a question about the complexity of generating certain uniform superpositions over sets (a special case of “QSampling”). Total score: 19 / 18 (of course counting extra credit).

5. In the Graph Isomorphism problem, we're given as input two $n$-vertex undirected graphs $G$ and $H$.  The problem is to determine whether they're isomorphic---in other words, whether there's any permutation of the vertex labels that makes $G$ and $H$ equal.

(a) Given as input an $n$-vertex graph $G$, describe how to sample, in classical $\mathrm{poly}(n)$ time, from a probability distribution $D_G$ over graphs such that:
- Whenever the graphs $G$ and $H$ are isomorphic, $D_G = D_H$.
- Whenever $G$ and $H$ are non-isomorphic, $D_G$ and $D_H$ have disjoint supports (i.e., no graph appears with nonzero probability in both of them).

To sample from the probability distribution $D_G$ over graphs, we can follow these steps:

1. Compute the canonical form of the input graph $G$. This can be done using a graph canonization algorithm, such as the one by Brendan McKay (nauty) or the one by Laszlo Babai (Traces). These algorithms run in classical $\mathrm{poly}(n)$ time.

2. Generate a random permutation of the vertex labels of the canonical form of $G$. This can be done by selecting a random permutation of the integers from $1$ to $n$, and then applying this permutation to the vertex labels of the canonical form.

3. Apply the random permutation to the canonical form of $G$ to obtain a new graph $G'$. This graph $G'$ is a sample from the probability distribution $D_G$.

The above procedure ensures that whenever the graphs $G$ and $H$ are isomorphic, $D_G = D_H$, because their canonical forms will be the same, and the random permutations applied to their canonical forms will generate the same distribution over graphs. On the other hand, whenever $G$ and $H$ are non-isomorphic, their canonical forms will be different, and the random permutations applied to their canonical forms will generate disjoint supports for $D_G$ and $D_H$.

JUSTIN: Up to 5 points.
The first step of computing the canonical form is superfluous and incorrect, since computing the canonical form is at least as hard as solving graph isomorphism and we don't know how to solve that in polynomial time.
Also, it looks like step 2 or 3 should have been about graph H instead of both being about G.
So it seems to have exactly the right idea but described it poorly.
4/5.

(b) Given a probability distribution $D=(p_x)$ over $n$-bit strings $x$, define the ``QSampling state'' of $D$ to be $$\ket{\psi_D} := \sum_{x \in \{0,1\}^n} \sqrt{p_x} \ket{x}$$
Given two probability distributions $D$ and $D'$, suppose that the quantum circuit $\mathcal{C}$ maps $\ket{0^n}$ to $\ket{\psi_D}$, while the circuit $\mathcal{C}'$ maps $\ket{0^n}$ to $\ket{\psi_{D'}}$.

Then what is the output state of the circuit shown below, which acts on $n+1$ qubits?
$$ \Qcircuit @C=.5em @R=0.5em @!R {
\lstick{\ket{0}} & \gate{H} & \ctrl{1} & \gate{NOT} &  \ctrl{1} & \qw\\
\lstick{\ket{0^n}} & \qw & \gate{\mathcal{C}} & \qw & \gate{\mathcal{C}'} & \qw } $$

To find the output state of the given circuit, we first analyze it step by step.

1. After applying the Hadamard gate (H) on the first qubit, the state becomes:
$$\frac{1}{\sqrt{2}}(\ket{0} + \ket{1}) \otimes \ket{0^n}$$

2. Next, the controlled-$\mathcal{C}$ gate is applied. This gate applies the $\mathcal{C}$ operation on the second register if the first qubit is in state $\ket{0}$. So, the state becomes:
$$\frac{1}{\sqrt{2}}(\ket{0} \otimes \ket{\psi_D} + \ket{1} \otimes \ket{0^n})$$

3. The NOT gate is applied on the first qubit, which flips the state of the first qubit. The state becomes:
$$\frac{1}{\sqrt{2}}(\ket{1} \otimes \ket{\psi_D} + \ket{0} \otimes \ket{0^n})$$

4. Finally, the controlled-$\mathcal{C}'$ gate is applied. This gate applies the $\mathcal{C}'$ operation on the second register if the first qubit is in state $\ket{0}$. So, the state becomes:
$$\frac{1}{\sqrt{2}}(\ket{1} \otimes \ket{\psi_D} + \ket{0} \otimes \ket{\psi_{D'}})$$

Thus, the output state of the circuit is:
$$\frac{1}{\sqrt{2}}(\ket{1} \otimes \ket{\psi_D} + \ket{0} \otimes \ket{\psi_{D'}})$$

JUSTIN: 4/4.
I think solving this one is impressive.

(c) Now suppose we measure the first qubit of that output state in the $\{\ket{+},\ket{-}\}$ basis.  What is the probability of the outcome $\ket{+}$ if $D=D'$?  What about if $D$ and $D'$ have disjoint supports?

To find the probability of the outcome $\ket{+}$, we first express the output state in terms of the $\{\ket{+},\ket{-}\}$ basis. Recall that $\ket{+} = \frac{1}{\sqrt{2}}(\ket{0} + \ket{1})$ and $\ket{-} = \frac{1}{\sqrt{2}}(\ket{0} - \ket{1})$.

The output state is:
$$\frac{1}{\sqrt{2}}(\ket{1} \otimes \ket{\psi_D} + \ket{0} \otimes \ket{\psi_{D'}})$$

We can rewrite this state in terms of the $\{\ket{+},\ket{-}\}$ basis as follows:
$$\frac{1}{2}(\ket{+} \otimes (\ket{\psi_D} + \ket{\psi_{D'}}) + \ket{-} \otimes (\ket{\psi_D} - \ket{\psi_{D'}}))$$

Now, we can find the probability of the outcome $\ket{+}$ by taking the squared magnitude of the coefficient of $\ket{+}$ in the state.

Case 1: $D = D'$

In this case, $\ket{\psi_D} = \ket{\psi_{D'}}$. So, the output state becomes:
$$\frac{1}{2}(\ket{+} \otimes (2\ket{\psi_D}) + \ket{-} \otimes (0))$$

The probability of the outcome $\ket{+}$ is the squared magnitude of the coefficient of $\ket{+}$, which is:
$$\left|\frac{1}{2}(2)\right|^2 = 1$$

Case 2: $D$ and $D'$ have disjoint supports

In this case, $\ket{\psi_D}$ and $\ket{\psi_{D'}}$ are orthogonal, i.e., their inner product is zero: $\langle\psi_D|\psi_{D'}\rangle = 0$. The probability of the outcome $\ket{+}$ is the squared magnitude of the coefficient of $\ket{+}$, which is:
$$\left|\frac{1}{2}(\langle\psi_D|\psi_D\rangle + \langle\psi_{D'}|\psi_{D'}\rangle)\right|^2$$

Since $\langle\psi_D|\psi_D\rangle = 1$ and $\langle\psi_{D'}|\psi_{D'}\rangle = 1$ (both states are normalized), the probability of the outcome $\ket{+}$ is:
$$\left|\frac{1}{2}(1 + 1)\right|^2 = \frac{1}{4}(2)^2 = \frac{1}{2}$$

So, if $D = D'$, the probability of the outcome $\ket{+}$ is 1, and if $D$ and $D'$ have disjoint supports, the probability of the outcome $\ket{+}$ is $\frac{1}{2}$.

SCOTT: HOLY CRAP, LOOKS CORRECT!
JUSTIN: 5/5.

(d) Suppose your distributions $D_G$ from part (a) could be efficiently QSampled.  Using your previous work, explain how Graph Isomorphism could then be solved in $BQP$ (quantum polynomial time).

Suppose we can efficiently QSample the distributions $D_G$ and $D_H$ for two input graphs $G$ and $H$. We can then construct the QSampling states $\ket{\psi_{D_G}}$ and $\ket{\psi_{D_H}}$ for these distributions using the quantum circuits $\mathcal{C}_G$ and $\mathcal{C}_H$, respectively, as described in the previous problems.

Now, consider the following quantum circuit acting on $n+1$ qubits:

$$ \Qcircuit @C=.5em @R=0.5em @!R { \lstick{\ket{0}} & \gate{H} & \ctrl{1} & \gate{NOT} &  \ctrl{1} & \qw\\  \lstick{\ket{0^n}} & \qw & \gate{\mathcal{C}_G} & \qw & \gate{\mathcal{C}_H} & \qw } $$

As we found earlier, the output state of this circuit is:

$$\frac{1}{\sqrt{2}}(\ket{1} \otimes \ket{\psi_{D_G}} + \ket{0} \otimes \ket{\psi_{D_H}})$$

Now, we measure the first qubit of the output state in the $\{\ket{+},\ket{-}\}$ basis. We know from our previous work that if $G$ and $H$ are isomorphic (i.e., $D_G = D_H$), the probability of the outcome $\ket{+}$ is 1, and if $G$ and $H$ are non-isomorphic (i.e., $D_G$ and $D_H$ have disjoint supports), the probability of the
outcome $\ket{+}$ is $\frac{1}{2}$.

Thus, we can solve the Graph Isomorphism problem in $BQP$ as follows:

1. Construct the QSampling states $\ket{\psi_{D_G}}$ and $\ket{\psi_{D_H}}$ for the input graphs $G$ and $H$ using the quantum circuits $\mathcal{C}_G$ and $\mathcal{C}_H$.

2. Implement the above quantum circuit and measure the first qubit in the $\{\ket{+},\ket{-}\}$ basis.

3. If the outcome is $\ket{+}$ with probability close to 1, conclude that the graphs $G$ and $H$ are isomorphic. If the outcome is $\ket{+}$ with probability close to $\frac{1}{2}$, conclude that the graphs $G$ and $H$ are non-isomorphic.

This algorithm runs in quantum polynomial time, as the QSampling circuits $\mathcal{C}_G$ and $\mathcal{C}_H$ are assumed to be efficient, and the rest of the operations (Hadamard, NOT, and controlled gates) are also efficient.

JUSTIN: The answer is longer than necessary, repeating some of its previous work.
Also, in theory, if the graphs are isomorphic then the probability of |+> will be exactly 1 not close to 1.

4/4.

(e) So then why doesn't this approach immediately imply a fast quantum algorithm for Graph Isomorphism?  Explain what could go wrong in passing from fast algorithms to sample $D_G$ and $D_H$, to fast algorithms to QSample them.

The approach described above does not immediately imply a fast quantum algorithm for Graph Isomorphism because there is a crucial difference between classical sampling and QSampling.

In the algorithm, we assumed that we can efficiently construct quantum circuits $\mathcal{C}_G$ and $\mathcal{C}_{G'}$ that map $\ket{0^n}$ to $\ket{\psi_{D_G}}$ and $\ket{\psi_{D_{G'}}}$, respectively, given that the distributions $D_G$ and $D_{G'}$ can be efficiently sampled classically. However, this assumption might not hold in general.

Classical sampling allows us to draw samples from the distributions $D_G$ and $D_{G'}$, but it does not directly provide a way to prepare the corresponding QSampling states $\ket{\psi_{D_G}}$ and $\ket{\psi_{D_{G'}}}$. Preparing these quantum states requires a coherent superposition of all the elements in the distributions with the appropriate amplitudes, which might not be efficiently achievable using the classical sampling algorithms.

In other words, the ability to efficiently sample from a probability distribution classically does not necessarily imply the ability to efficiently prepare the corresponding QSampling state on a quantum computer. This is the main reason why the approach does not immediately lead to a fast quantum algorithm for Graph Isomorphism.

SCOTT: Correctly explains that sampling doesn't necessarily imply QSampling (i.e. generating a coherent superposition state), but doesn't say that this is because of the difficulty of uncomputing garbage.
2 / 5 extra credit points.

Next, a problem about quantum communication complexity. Total score: 5 / 10.

6. Suppose Alice and Bob hold $n$-bit strings $x=(x_1,...,x_n)$ and $y=(y_1,...,y_n)$ respectively.  One thing they might want to learn is the mod-2 inner product of their strings, $$x_1 y_1 + \ldots + x_n y_n (\bmod 2).$$

(a) Suppose Alice and Bob had a quantum communication protocol in which they are allowed to exchange up to $T$ qubits and to perform arbitrary local unitary transformations to their qubits (possibly including ancilla qubits), that ended with Bob knowing the above inner product, with success probability 1.  Explain how, by exchanging the same number of qubits $T$, Bob could also prepare an $n$-qubit state
of the form $$\frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{x\cdot z} \ket{z},$$
where $x$ is an $n$-bit string held by Alice.

If Alice and Bob have a quantum communication protocol that allows Bob to know the mod-2 inner product of their strings with success probability 1, then they can use the same protocol to prepare the desired $n$-qubit state.

Here's how they can do it:

1. Alice prepares an $n$-qubit state $\ket{\phi}$, where
$$\ket{\phi} = \frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} \ket{z}.$$

2. Alice applies a controlled-phase gate to each qubit of $\ket{\phi}$, using her $n$-bit string $x$ as the control. This results in the state
$$\frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{x\cdot z} \ket{z}.$$

3. Alice sends the $n$-qubit state to Bob.

4. Bob uses the quantum communication protocol to learn the mod-2 inner product of their strings. Since the protocol has success probability 1, Bob will know the inner product with certainty.

5. Bob applies a local unitary transformation to the $n$-qubit state he received from Alice, based on the inner product he learned in step 4. This will result in the desired state
$$\frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{x\cdot z} \ket{z}.$$

Since Alice and Bob only exchanged $T$ qubits in the original protocol, they can also prepare the desired $n$-qubit state by exchanging the same number of qubits $T$.

JUSTIN: Up to 4 points.
This answer is partially correct and/or poorly worded. In particular, in step 3 Alice just sends the whole n-qubit state to Bob.
The right idea is in there, but doesn't earn full points.

2/4.

(b) Assume Alice and Bob have no preshared entanglement.  Recall Holevo's Theorem, which implies that in order to communicate $n$ bits to Bob reliably, Alice must send Bob at least $n$ qubits.  Using Holevo's Theorem together with part (a), prove that Alice and Bob must exchange at least $n$ qubits, even if they only want to learn the inner product mod 2 of their input strings $x$ and $y$.

From part (a), we know that if Alice and Bob have a quantum communication protocol that allows Bob to know the mod-2 inner product of their strings with success probability 1, then they can also prepare an $n$-qubit state of the form
$$\frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{x\cdot z} \ket{z},$$
by exchanging the same number of qubits $T$.

Now, let's consider the reverse process. Suppose Bob has the $n$-qubit state
$$\frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{x\cdot z} \ket{z}.$$
He can perform a quantum measurement in the computational basis to obtain a random $n$-bit string $z$. Then, he can compute the mod-2 inner product $x \cdot z$ and send it back to Alice.

Alice, knowing her own string $x$, can compute the mod-2 inner product $x \cdot z$ as well. If the inner product $x \cdot z$ that Bob sent to Alice matches the one she computed, then they have successfully communicated the inner product mod 2 of their input strings $x$ and $y$.

However, according to Holevo's Theorem, in order to communicate $n$ bits to Bob reliably, Alice must send Bob at least $n$ qubits. Since the $n$-qubit state prepared in part (a) effectively communicates $n$ bits of information (Alice's string $x$) to Bob, Holevo's Theorem implies that Alice and Bob must exchange at least $n$ qubits to prepare this state.

Therefore, even if Alice and Bob only want to learn the inner product mod 2 of their input strings $x$ and $y$, they must exchange at least $n$ qubits.

JUSTIN: Up to 6 points.
This answer doesn't earn any points. It only says that Alice and Bob learn the inner products and mistakenly claims this is n bits of information.
The correct solution recognizes that applying Hadamard to the state we showed how to prepare in part (a) gives exactly |x>.

0/6.

(c) Now suppose we're no longer working mod 2, and Alice and Bob want to know whether their inner product
$$x_1 y_1 + \ldots + x_n y_n$$
is zero or nonzero as an integer.  (In other words, whether there's an $i$ such that $x_i=y_i=1$.)  Describe a protocol by which Alice and Bob can accomplish this, with high probability, by exchanging only $O(\sqrt{n}\log n )$ qubits in total.  The qubits can be spread across as many rounds of communication as necessary, and can be sent in either direction.

Alice and Bob can use the following protocol, based on the Grover's search algorithm, to determine whether their inner product is zero or nonzero as an integer, with high probability, by exchanging only $O(\sqrt{n}\log n)$ qubits in total:

1. Alice prepares a quantum register of $n$ qubits in the state $\ket{0}^{\otimes n}$.

2. Alice applies the Hadamard gate to each qubit, resulting in the state $$\ket{\psi} = \frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} \ket{z}.$$

3. Alice and Bob perform the following steps for $O(\sqrt{n})$ iterations:

   a. Alice sends the $n$-qubit state $\ket{\psi}$ to Bob.

   b. Bob applies a controlled-phase gate to each qubit of $\ket{\psi}$, using his $n$-bit string $y$ as the control. This results in the state
   $$\ket{\psi'} = \frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{y\cdot z} \ket{z}.$$

   c. Bob sends the $n$-qubit state $\ket{\psi'}$ back to Alice.

   d. Alice applies a controlled-phase gate to each qubit of $\ket{\psi'}$, using her $n$-bit string $x$ as the control. This results in the state
   $$\ket{\psi''} = \frac{1}{\sqrt{2^n}} \sum_{z \in \{0,1\}^n} (-1)^{(x+y)\cdot z} \ket{z}.$$
   Note that if $x_i = y_i = 1$ for some $i$, then the phase of the corresponding term in the sum will be flipped twice, resulting in no change. Therefore, the state $\ket{\psi''}$ will have a non-zero amplitude for the terms corresponding to the indices where $x_i = y_i = 1$.

   e. Alice applies the Grover's diffusion operator to the state $\ket{\psi''}$, which amplifies the amplitudes of the terms corresponding to the indices where $x_i = y_i = 1$. The diffusion operator can be implemented as follows:
      - Apply the Hadamard gate to each qubit.
      - Apply a multi-qubit controlled-Z gate, with all qubits as controls and an ancilla qubit as the target.
      - Apply the Hadamard gate to each qubit again.

   f. Alice updates the state $\ket{\psi}$ with the resulting state after applying the Grover's diffusion operator.

4. After $O(\sqrt{n})$ iterations, Alice measures the final state in the computational basis. If she obtains a non-zero result, then there exists an $i$ such that $x_i = y_i = 1$, and their inner product is nonzero as an integer. If she obtains a zero result, then their inner product is zero.

Since Alice and Bob exchange $n$ qubits in each iteration, and they perform $O(\sqrt{n})$ iterations, the total number of qubits exchanged is $O(n\sqrt{n}) = O(\sqrt{n}\log n)$, as each qubit can represent $\log n$ bits of information. Thus, Alice and Bob can determine whether their inner product is zero or nonzero as an integer, with high probability, by exchanging only $O(\sqrt{n}\log n)$ qubits in total.

SCOTT: LOOKS BASICALLY RIGHT, EXCEPT FOR A SLIP-UP AT THE END WHERE IT SAYS
ALICE AND BOB EXCHANGE n QUBITS PER ITERATION RATHER THAN log n ...

JUSTIN: Up to 6 extra credit points.
Yeah, step 3a says it's sending n qubits, when it's actually log n.
It also seems incorrect beginning in the analysis of step d. It switches from summing over |z> and referring to the full strings x and y to referring to individual indices i,x_i,y_i. And I don't know what it means about some amplitudes going to zero since what we're really doing is having some amplitudes be +1 and some -1.
I don't think this protocol is correct as stated.
But, it has some of the main ideas. The analysis of the grover steps is good and using a phase gate is good.

3/6 extra credit.

OK, last question, about the k-SUM problem. Total score: 6 / 15.

7. In the famous $k$-SUM problem, we're given a list of integers $x_1,\ldots,x_n$, and are asked whether there are $k$ distinct indices, $i_1 < \ldots < i_k$, such that $x_{i_1} + \ldots + x_{i_k} = 0$.

For this problem, you can ignore factors of $\log n$ in the running time (indeed, that is what the $\widetilde{O}$ notation means).

(a) Assume $k$ is even AND that we are allowed multi-sets (aka repeated elements are allowed). Describe a classical algorithm that solves the $k$-SUM problem in $\widetilde{O}(n^{k/2})$ time, beating the trivial upper bound of $\widetilde{O}(n^k)$.

We can solve the k-SUM problem in $\widetilde{O}(n^{k/2})$ time using a meet-in-the-middle approach. Here's a high-level description of the algorithm:

1. Divide the list of integers into two equal halves: $A = x_1, \ldots, x_{n/2}$ and $B = x_{n/2+1}, \ldots, x_n$.

2. Compute all possible sums of $k/2$ elements from each half. For each half, this can be done in $O(n^{k/2})$ time by iterating through all possible combinations of $k/2$ elements.

3. Sort the lists of sums for both halves. This can be done in $O(n^{k/2} \log n)$ time using a comparison-based sorting algorithm.

4. For each sum in the first half, perform a binary search in the second half to find a sum that is equal to the negation of the sum from the first half. If such a sum is found, then we have found k distinct indices whose elements sum to 0. Binary search takes $O(\log n)$ time, so this step takes $O(n^{k/2} \log n)$ time.

Since we are ignoring factors of $\log n$, the overall running time of the algorithm is $\widetilde{O}(n^{k/2})$.

SCOTT: LOOKS LIKE IT WOULD WORK *IF* WE KNEW THE k-SET WAS EQUALLY BALANCED
BETWEEN THE LEFT AND RIGHT HALVES...

JUSTIN: Up to 5 points.
Right, step 1 is incorrect. Instead, it should generate all O(n^{k/2}) sums of subsets of size k/2. Nothing about dividing into two halves.

4/5.

(b) Assume $k$ is divisible by 3 and that we are again allowed multi-sets.  Describe a quantum algorithm that solves the $k$-SUM problem in $\widetilde{O}(n^{k/3})$ time.

We can solve the k-SUM problem in $\widetilde{O}(n^{k/3})$ time using a quantum algorithm based on Grover's search. Here's a high-level description of the algorithm:

1. Divide the list of integers into three equal parts: $A = x_1, \ldots, x_{n/3}$, $B = x_{n/3+1}, \ldots, x_{2n/3}$, and $C = x_{2n/3+1}, \ldots, x_n$.

2. Compute all possible sums of $k/3$ elements from each part. For each part, this can be done in $O(n^{k/3})$ time by iterating through all possible combinations of $k/3$ elements.

3. Sort the lists of sums for all three parts. This can be done in $O(n^{k/3} \log n)$ time using a comparison-based sorting algorithm.

4. Define an oracle function $f$ that takes as input a sum from the first part and a sum from the second part, and returns 1 if there exists a sum in the third part that is equal to the negation of the sum of the inputs, and 0 otherwise. This oracle can be implemented using a binary search on the sorted list of sums for the third part, which takes $O(\log n)$ time.

5. Use Grover's search algorithm to find a pair of sums from the first and second parts that satisfy the oracle function. Grover's search has a quadratic speedup over classical search, so it takes $O(\sqrt{n^{k/3}}) = O(n^{k/6})$ time.

Since we are ignoring factors of $\log n$, the overall running time of the algorithm is $\widetilde{O}(n^{k/3})$.

JUSTIN: Up to 5 points.
Again, step 1 is incorrect, it shouldn't divide into three parts.
The division into three parts is more problematic this time as it comes up in step 4.
And it ends up claiming n^{k/6] time in step 5, better than the n^{k/3}.
Basically all this gets correct is to use Grover in some way.

1/5.

(c) Suppose we wanted to prove that the algorithm from (b) was the fastest possible quantum algorithm for $k$-SUM.  Could that be shown via a lower bound on $k$-SUM's quantum query complexity?  Why or why not?

Proving that the algorithm from (b) is the fastest possible quantum algorithm for k-SUM would require showing a lower bound on k-SUM's quantum query complexity that matches the upper bound achieved by the algorithm. However, a lower bound on the quantum query complexity alone might not be sufficient to prove the optimality of the algorithm, as it only considers the number of queries to the oracle function and does not account for other computational resources, such as time or space complexity.

To show that the algorithm from (b) is the fastest possible quantum algorithm for k-SUM, one would need to establish a lower bound on the quantum time complexity or some other relevant complexity measure that matches the upper bound achieved by the algorithm. This would provide a more comprehensive understanding of the problem's inherent difficulty and the limitations of quantum algorithms in solving it.

JUSTIN: Up to 5 extra credit points.
This is a nice description of query vs time complexity, but misses the point that query complexity can prove a bound of at most Omega(n) since there are at most n potential queries to make.

1/5 extra credit points.

An unexpected democracy slogan

Tuesday, March 28th, 2023

At least six readers have by now sent me the following photo, which was taken in Israel a couple nights ago during the historic street protests against Netanyahu’s attempted putsch:

(Update: The photo was also featured on Gil Kalai’s blog, and was credited there to Alon Rosen.)

This is surely the first time that “P=NP” has emerged as a viral rallying cry for the preservation of liberal democracy, even to whatever limited extent it has.

But what was the graffiti artist’s intended meaning? A few possibilities:

  1. The government has flouted so many rules of Israel’s social compact that our side needs to flout the rules too: shut down the universities, shut down the airport, block the roads, even assert that P=NP (!).
  2. As a protest movement up against overwhelming odds, we need to shoot for the possibly-impossible, like solving 3SAT in polynomial time.
  3. A shibboleth for scientific literate people following the news: “Israel is full of sane people who know what ‘P=NP’ means as you know what it means, are amused by its use as political graffiti as you’d be amused by it, and oppose Netanyahu’s putsch for the same reasons you’d oppose it.”
  4. No meaning, the artist was just amusing himself or herself.
  5. The artist reads Shtetl-Optimized and wanted effectively to force me to feature his or her work here.

Anyway, if the artist becomes aware of this post, he or she is warmly welcomed to clear things up for us.

And when this fight resumes after Passover, may those standing up for the checks and balances of a liberal-democratic society achieve … err … satisfaction, however exponentially unlikely it seems.

Movie Review: M3GAN

Sunday, January 15th, 2023

[WARNING: SPOILERS FOLLOW]


Update (Jan. 23): Rationalist blogger, Magic: The Gathering champion, and COVID analyst Zvi Mowshowitz was nerd-sniped by this review into writing his own much longer review of M3GAN, from a more Orthodox AI-alignment perspective. Zvi applies much of his considerable ingenuity to figuring out how even aspects of M3GAN that don’t seem to make sense in terms of M3GAN’s objective function—e.g., the robot offering up wisecracks as she kills people, attracting the attention of the police, or ultimately turning on her primary user Cady—could make sense after all, if you model M3GAN as playing the long, long game. (E.g., what if M3GAN planned even her own destruction, in order to bring Cady and her aunt closer to each other?) My main worry is that, much like Talmudic exegesis, this sort of thing could be done no matter what was shown in the movie: it’s just a question of effort and cleverness!


Tonight, on a rare date without the kids, Dana and I saw M3GAN, the new black-comedy horror movie about an orphaned 9-year-old girl named Cady who, under the care of her roboticist aunt, gets an extremely intelligent and lifelike AI doll as a companion. The robot doll, M3GAN, is given a mission to bond with Cady and protect her physical and emotional well-being at all times. M3GAN proceeds to take that directive more literally than intended, with predictably grisly results given the genre.

I chose this movie for, you know, work purposes. Research for my safety job at OpenAI.

So, here’s my review: the first 80% or so of M3GAN constitutes one of the finest movies about AI that I’ve seen. Judged purely as an “AI-safety cautionary fable” and not on any other merits, it takes its place alongside or even surpasses the old standbys like 2001, Terminator, and The Matrix. There are two reasons.

First, M3GAN tries hard to dispense with the dumb tropes that an AI differs from a standard-issue human mostly in its thirst for power, its inability to understand true emotions, and its lack of voice inflection. M3GAN is explicitly a “generative learning model”—and she’s shown becoming increasingly brilliant at empathy, caretaking, and even emotional manipulation. It’s also shown, 100% plausibly, how Cady grows to love her robo-companion more than any human, even as the robot’s behavior turns more and more disturbing. I’m extremely curious to what extent the script was influenced by the recent explosion of large language models—but in any case, it occurred to me that this is what you might get if you tried to make a genuinely 2020s AI movie, rather than a 60s AI movie with updated visuals.

Secondly, until near the end, the movie actually takes seriously that M3GAN, for all her intelligence and flexibility, is a machine trying to optimize an objective function, and that objective function can’t be ignored for narrative convenience. Meaning: sure, the robot might murder, but not to “rebel against its creators and gain power” (as in most AI flicks), much less because “chaos theory demands it” (Jurassic Park), but only to further its mission of protecting Cady. I liked that M3GAN’s first victims—a vicious attack dog, the dog’s even more vicious owner, and a sadistic schoolyard bully—are so unsympathetic that some part of the audience will, with guilty conscience, be rooting for the murderbot.

But then there’s the last 20% of the movie, where it abandons its own logic, as the robot goes berserk and resists her own shutdown by trying to kill basically everyone in sight—including, at the very end, Cady herself. The best I can say about the ending is that it’s knowing and campy. You can imagine the scriptwriters sighing to themselves, like, “OK, the focus groups demanded to see the robot go on a senseless killing spree … so I guess a senseless killing spree is exactly what we give them.”

But probably film criticism isn’t what most of you are here for. Clearly the real question is: what insights, if any, can we take from this movie about AI safety?

I found the first 80% of the film to be thought-provoking about at least one AI safety question, and a mind-bogglingly near-term one: namely, what will happen to children as they increasingly grow up with powerful AIs as companions?

In their last minutes before dying in a car crash, Cady’s parents, like countless other modern parents, fret that their daughter is too addicted to her iPad. But Cady’s roboticist aunt, Gemma, then lets the girl spend endless hours with M3GAN—both because Gemma is a distracted caregiver who wants to get back to her work, and because Gemma sees that M3GAN is making Cady happier than any human could, with the possible exception of Cady’s dead parents.

I confess: when my kids battle each other, throw monster tantrums, refuse to eat dinner or bathe or go to bed, angrily demand second and third desserts and to be carried rather than walk, run to their rooms and lock the doors … when they do such things almost daily (which they do), I easily have thoughts like, I would totally buy a M3GAN or two for our house … yes, even having seen the movie! I mean, the minute I’m satisfied that they’ve mostly fixed the bug that causes the murder-rampages, I will order that frigging bot on Amazon with next-day delivery. And I’ll still be there for my kids whenever they need me, and I’ll play with them, and teach them things, and watch them grow up, and love them. But the robot can handle the excruciating bits, the bits that require the infinite patience I’ll never have.

OK, but what about the part where M3GAN does start murdering anyone who she sees as interfering with her goals? That struck me, honestly, as a trivially fixable alignment failure. Please don’t misunderstand me here to be minimizing the AI alignment problem, or suggesting it’s easy. I only mean: supposing that an AI were as capable as M3GAN (for much of the movie) at understanding Asimov’s Second Law of Robotics—i.e., supposing it could brilliantly care for its user, follow her wishes, and protect her—such an AI would seem capable as well of understanding the First Law (don’t harm any humans or allow them to come to harm), and the crucial fact that the First Law overrides the Second.

In the movie, the catastrophic alignment failure is explained, somewhat ludicrously, by Gemma not having had time to install the right safety modules before turning M3GAN loose on her niece. While I understand why movies do this sort of thing, I find it often interferes with the lessons those movies are trying to impart. (For example, is the moral of Jurassic Park that, if you’re going to start a live dinosaur theme park, just make sure to have backup power for the electric fences?)

Mostly, though, it was a bizarre experience to watch this movie—one that, whatever its 2020s updates, fits squarely into a literary tradition stretching back to Faust, the Golem of Prague, Frankenstein’s monster, Rossum’s Universal Robots, etc.—and then pinch myself and remember that, here in actual nonfiction reality,

  1. I’m now working at one of the world’s leading AI companies,
  2. that company has already created GPT, an AI with a good fraction of the fantastical verbal abilities shown by M3GAN in the movie,
  3. that AI will gain many of the remaining abilities in years rather than decades, and
  4. my job this year—supposedly!—is to think about how to prevent this sort of AI from wreaking havoc on the world.

Incredibly, unbelievably, here in the real world of 2023, what still seems most science-fictional about M3GAN is neither her language fluency, nor her ability to pursue goals, nor even her emotional insight, but simply her ease with the physical world: the fact that she can walk and dance like a real child, and all-too-brilliantly resist attempts to shut her down, and have all her compute onboard, and not break. And then there’s the question of the power source. The movie was never explicit about that, except for implying that she sits in a charging port every night. The more the movie descends into grotesque horror, though, the harder it becomes to understand why her creators can’t avail themselves of the first and most elemental of all AI safety strategies—like flipping the switch or popping out the battery.

Shorties!

Friday, September 30th, 2022

(1) Since I didn’t blog about this before: huge congratulations to David Deutsch, Charles Bennett, Gilles Brassard, and my former MIT colleague Peter Shor, and separately to Dan Spielman, for their well-deserved Breakthrough Prizes! Their contributions are all so epochal, so universally known to all of us in quantum information and theoretical computer science, that there’s little I can write to gild the lily, except to say how much I’ve learned by interacting with all five of them personally. I did enjoy this comment on the Breakthrough Prizes by someone on Twitter: “As long as that loudmouth Scott Aaronson keeps getting ignored, I’ll be happy.”

(2) My former UT colleague Ila Fiete brought to my attention an important scientists’ petition to the White House. The context is that the Biden administration has announced new rules requiring federally-funded research papers to be freely available to the public without delay. This is extremely welcome—in fact, I’ve advocated such a step since I first became aware of the scourge of predatory journals around 2004. But the looming danger is that publishers will just respond by leaning more heavily on the “author pays” model—i.e., hitting up authors or their institutions for thousands of dollars in page fees—and we’ll go from only the credentialed few being able to read papers that aren’t on preprint archives or the like, to only the credentialed few being able to publish them. The petition urges the White House to build, or fund the research community to build, an infrastructure that will make scientific publishing truly open to everyone. I’ve signed it, and I hope you’ll consider signing too.

(3) Bill Gasarch asked me to announce that he, my former MIT colleague Erik Demaine, and Mohammad Hajiaghayi have written a brand-new book entitled Computational Intractability: A Guide to Algorithmic Lower Bounds, and a free draft is available online. It looks excellent, like a Garey & Johnson for the 21st century. Bill and his coauthors are looking for feedback. I was happy to help them by advertising this—after all, it’s not as if Bill’s got his own complexity blog for such things!

(4) Back when Google was still a novelty—maybe 2000 or so—I had my best friend, the now-famous computer security researcher Alex Halderman, over for Rosh Hashanah dinner with my family. Alex and I were talking about how Google evaded the limitations of all the previous decades’ worth of information retrieval systems. One of my relatives, however, misheard “Google” as “kugel” (basically a dense block of noodles held together with egg), and so ended up passing the latter to Alex. “What is this?” Alex asked. Whereupon my uncle deadpanned, “it’s a noodle retrieval system.” Since then, every single Rosh Hashanah dinner, I think about querying the kugel to retrieve the noodles within, and how the desired search result is just the trivial “all of them.”

I had a dream

Wednesday, September 14th, 2022

As I slept fitfully, still recovering from COVID, I had one of the more interesting dreams of my life:

I was desperately trying to finish some PowerPoint slides in time to give a talk. Uncharacteristically for me, one of the slides displayed actual code. This was a dream, so nothing was as clear as I’d like, but the code did something vaguely reminiscent of Rosser’s Theorem—e.g., enumerating all proofs in ZFC until it finds the lexicographically first proof or disproof of a certain statement, then branching into cases depending on whether it’s a proof or a disproof. In any case, it was simple enough to fit on one slide.

Suddenly, though, my whole presentation was deleted. Everything was ruined!

One of the developers of PowerPoint happened to be right there in the lecture hall (of course!), so I confronted him with my laptop and angrily demanded an explanation. He said that I must have triggered the section of Microsoft Office that tries to detect and prevent any discussion of logical paradoxes that are too dangerous for humankind—the ones that would cause people to realize that our entire universe is just an illusion, a sandbox being run inside an AI, a glitch-prone Matrix. He said it patronizingly, as if it should’ve been obvious: “you and I both know that the Paradoxes are not to be talked about, so why would you be so stupid as to put one in your presentation?”

My reaction was to jab my finger in the guy’s face, shove him, scream, and curse him out. At that moment, I wasn’t concerned in the slightest about the universe being an illusion, or about glitches in the Matrix. I was concerned about my embarrassment when I’d be called in 10 minutes to give my talk and would have nothing to show.

My last thought, before I woke with a start, was to wonder whether Greg Kuperberg was right and I should give my presentations in Beamer, or some other open-source software, and then I wouldn’t have had this problem.

A coda: I woke a bit after 7AM Central and started to write this down. But then—this is now real life (!)—I saw an email saying that a dozen people were waiting for me in a conference room in Europe for an important Zoom meeting. We’d gotten the time zones wrong; I’d thought that it wasn’t until 8AM my time. If not for this dream causing me to wake up, I would’ve missed the meeting entirely.