CHAPTER SIX

Reality Check

I’ve been glum lately, unsure how or whether my quantum project should proceed. But now, a week before summer vacation ends, my path is clear. Exercising my free will, I’m taking a quantum-mechanics course at my school, Stevens Institute of Technology. The course is PEP553: Quantum Mechanics and Engineering Applications. Like all Stevens classes this fall, PEP53 will be online, to protect us from Covid. The Stevens catalogue says PEP553 provides 

an introduction to quantum mechanics for students in physics and engineering. Techniques discussed include solutions of the Schrodinger equation in one and three dimensions, and operator and matrix methods.

I embarked on my quantum project to see if I could learn real physics, the way physicists learn it. PEP553 will be a true test, a reality check. And it’s an introductory course. How hard could it be? To sign up for PEP553, I need permission from the instructor, Professor Edward Whittaker. Ed is a great guy, easy-going, about my age, and an accomplished physicist. He got his Ph.D. from Columbia in 1982 and joined Stevens two years later, after stints at IBM and Bell Labs. He holds patents and has published papers on optical devices for sensing and communication. Ed is no armchair physicist.

Ed and I met soon after I arrived at Stevens in 2005, and we had a friendly public debate over my claim that physics is ending. When I tell Ed about my quantum experiment and ask if I can take PEP553, he’s thrilled. It’ll be fun having me as a student, he says, I’ll keep him on his toes. Although graduate students take PEP553, it’s aimed primarily at undergraduates majoring in physics and engineering. I’ll need calculus and linear algebra but nothing too arduous. Ed has been wondering how best to teach quantum mechanics, especially to engineers. What do they really need to learn, and how should they learn it? Maybe I can give him ideas. I reply that I’m flattered he would think so. 

That damn cat still spooks me.

Ed urges me to get a running start on the course textbook. It is a popular college-level text, Introduction to Quantum Mechanics, third edition, by David Griffiths and Darrell Schroeter. Amazon lists the book, which I’ll refer to henceforth as Griffiths, for $75, but I find a used copy for $25. Griffiths is very textbook-y, a glossy hardcover published by Cambridge University Press. Its appearance, smell and weight remind me of my college days, not in a good way. The black-and-white cat on the cover, an allusion to Schrodinger’s cat, is probably supposed to make the book less intimidating. But that infernal, yellow-eyed feline sees right through me.

Things get worse when I open the book. The inside cover and opposite page list formulas under three headings: Fundamental Equations (beginning with “Schrodinger equation”), Fundamental Constants (beginning with “Planck’s constant”), and Hydrogen Atom (“Fine structure constant”.) This book is written to instruct, not entertain; it does not fuck around. The preface tries to soften us up: 

At first glance, this book might strike you as forbiddingly mathematical. We encounter Legendre, Hermite, and Laguerre polynomials, spherical harmonics, Bessel, Neumann, and Hankel functions, Airy functions, and even the Riemann zeta function—not to mention Fourier transforms, Hilbert spaces, Hermitian operators, and Clebsch-Gordan coefficients. Is all this baggage really necessary? Perhaps not, but physics is like carpentry: Using the right tool makes the job easier, not more difficult, and teaching quantum mechanics without the appropriate mathematical equipment is like having a tooth extracted with a pair of pliers—it’s possible, but painful.

Yeah, in addition to a big key ring, the physicist/handyman wears a tool belt sagging under the weight of Fourier transforms and Hermitian operators. Other than “Airy functions,” which sound delightful, the list of math gadgetry makes me gulp. So does the dentistry analogy; I just had a root canal, and my jaw still aches. I’m also daunted when Griffiths says, “Elementary classical mechanics is essential” for learning quantum mechanics. Does skimming Leonard Susskind’s book on classical mechanics count?

The Schrodinger Equation

Reading Susskind has prepared me, sort of, for Griffiths: I recognize a few notations, terms and concepts. But Griffiths and Susskind approach their subjects in strikingly different ways. Susskind draws us into the quantum bizarro world by talking about a simple two-state system, spin, which can have one of two values, such as +1 and -1. Susskind doesn’t get to the Schrodinger equation until page 110.

Griffiths, in contrast, starts not with spin but with a more conventional physics scenario: a particle moving through space. And Griffiths shoves the Schrodinger equation in our faces on the first page of the first chapter. The Schrodinger equation is a partial differential equation, which means it lets you calculate derivatives, or rates of change, of certain values while holding others constant. Quick Calculus exposed me to partial differential equations, but they gave me trouble. Below is the Schrodinger equation presented by Griffiths:

This is a photo of the Schrodinger equation, which I don’t know how to type out. See below for full page of Griffiths on which the equation appears.

Annoyingly, this is just one of many versions of “the” Schrodinger equation. Griffiths contrasts the Schrodinger equation with classical laws of motion. The latter let you track the motion of a particle with a given mass, m, propelled along an axis, x, by a force, F, which equals mass times acceleration, F = ma. Classical equations let you calculate the position of the particle, x(t), at any given time, t.

In the case of the Schrodinger equation, you are solving not for the position of a particle at time t, but for the value of the particle’s wave function. The wave function is represented by this symbol, Ψ, spelled psi and pronounced “sigh.” The wave function lets you compute the probability that the particle will be at a certain position at time t. And to get that probability, you must square the height of the wave function, which is called a probability amplitude, at t. Ah, yes, probability amplitudes and the square rule. Familiarity does not equal comprehension, but it helps.

Griffiths walks us through other components of the Schrodinger equation. i, the square root of negative one, is old hat to me now. The ħ, pronounced “h bar,” is Planck’s constant, which equals 1.054571817...× 10 to the -34 power J s. (I spell out 10 to the -34 power because I don’t know how to write it as superscript.) “J s” stands for joule-second, which is not the same as joules per second. A joule is a unit of work, and a joule-second is a unit of action, a measure of how much something changes over time. [1] 10 to the minus 34 power is extremely small number, equal to 1 divided by 10,000,000,000,000,000,000,000,000,000,000,000, indicating that we are dealing with extremely small scales. Planck’s constant accounts for quantum lumpiness. Particles’ energies can’t take any value; they come in multiples of ħ. V stands for potential energy. But the key component of the Schrodinger equation is the wave function, Ψ, the emblem of quantum weirdness.

Page 1 of Chapter 1 of Griffiths, plus my grumpy annotations. The Schrodinger equation is in the box in the middle of the page.

Derealization

Griffiths flicks at the weirdness early on. He warns that quantum mechanics “retains to this day some of the scars of its exhilarating but traumatic youth,” and that “there is no general consensus as to what its fundamental principles are, how it should be taught, or what it really ‘means.’” I like the scare quotes around “means.” Griffiths notes the peculiar fact that the wave function of a particle gives you only its probable location. He lays out the “main schools of thought regarding quantum indeterminacy,” which I’ve partially paraphrased: 

1.     The realist position, advocated by Einstein, says the particle is in a certain location even when we’re not looking at it. Indeterminacy “is not a fact of nature, but a reflection of our ignorance.”

2.     The orthodox position, spelled out in the Copenhagen interpretation of Niels Bohr, insists that the particle “wasn’t really anywhere” until we measure it.

3.     The agnostic position says, in effect, who knows and who cares? Since we can only know what’s happening when we look, what’s the point of worrying about what’s happening when we’re not looking?

Arguments over how to interpret quantum mechanics once seemed irresolvable and hence moot, Griffiths says. But then John Bell “astonished the physics community by showing that it makes an observable difference whether the particle had a precise (though unknown) position prior to the measurement, or not.” Experiments have “decisively confirmed the orthodox interpretation: a particle simply does not have a precise position prior to measurement.” The italics are mine.

Good question.

I linger on Griffiths’ comment that “the realist position” has been “decisively” refuted. Really? That’s not the impression I get from What Is Real? The Unfinished Quest for the Meaning of Quantum Mechanics by Adam Becker. Becker, who has a doctorate in astrophysics, recounts efforts to make sense of quantum mechanics, from the Einstein-Bohr debates up to the present. Becker says realism, the position espoused by Einstein, is making a comeback against Bohr’s orthodox obfuscations.

I take the question of what is real personally. I’ve been afflicted since I was a kid by bouts of derealization, during which you become estranged from the external world and even your own self. Everything seems unreal, like a movie or dream. That’s why I love talking to my first-year humanities students about Plato’s parable of the cave; it gives me a chance to vent my anxious obsession over what is real.

Many sages, not just Plato, have proposed that the everyday world, in which we go about the business of living, is an illusion. The 8th-century Hindu philosopher Adi Shankara asserted that ultimate reality is an eternal, undifferentiated field of consciousness. The Buddhist doctrine of anatta says our individual selves are illusory. Modern philosophers conjecture that our cosmos is a simulation, a virtual reality created by the godlike equivalent of a bored teenage hacker.

In 1981, I took an extremely potent psychedelic that knocked me for a loop. I emerged from the trip convinced that existence is the fever dream of an insane god. For months the world felt wobbly, flimsy, like a screen on which images were projected. I feared that at any moment everything might vanish, giving way to—I didn’t know what, hence the fear. Gradually, my terror faded, and I could ponder the implications of my vision with intellectual detachment. I applauded myself for having solved the problem of evil: our world is fucked up because God is fucked up.

Once I became a father and teacher, my view of antirealist theories shifted. I decided that such conjectures, whether Platonism, the simulation hypothesis or my insane-god theodicy, are immoral, because they encourage escapism and nihilism. Why give a shit about poverty, injustice, environmental destruction, pandemics, war and other sources of suffering if life is but a dream?

If they think about philosophy at all, most scientists assume that the world has an objective, physical existence, independent of us, that we can discover through science. This position is sometimes called scientific realism, or, by critics, naïve realism. I am, for moral and political reasons, and for the sake of my mental health, a naïve realist. But my quantum experiment, like that drug I took in 1981, is knocking me for a loop. My naïve realism is wobbling. I hope PEP553 will yank me back to solid ground.

Back to School

As the semester approaches, my anxiety grows. I must prepare not only for PEP553 but also for the three humanities classes I’m teaching. Can I do a decent job teaching while also keeping up as a student in PEP553? Professor Whittaker emails a survey to everyone enrolled in the class to assess our backgrounds. There are three questions: 

1, I have taken a course in modern physics prior to taking PEP553, such as PEP242 or PEP201.

2, I have taken a course in advanced mathematics beyond multivariate calculus. For example Math Methods of Physics, PEP332 or PEP527. Such a course might cover matrix methods, differential equations, etc.

3, I have taken a course in “intermediate mechanics,” such as PEP538.

I reply No to all three questions. Once again, to hold my future self hostage--that is, to make it harder for that lazy weakling to change his mind--I post a note on Facebook: “Tomorrow I’m going to be a student again for the first time in 37 years. I begin PEP553: Quantum Mechanics and Engineering Applications at Stevens Institute of Technology, taught by Prof. Ed Whittaker. I’m prepared for total, humiliating failure.”

Friends respond with quantum cracks. Gary, a pal from journalism school: “Just keep your mouth shut and kiss a lot of ass. That’s what got me my Quantum Engineering Degree from Trump University.” Chris, a childhood friend and computer scientist: “Don’t just be prepared for humiliating failure. Prepare for humiliating failure and glorious success AT THE SAME TIME.” Funny. Not funny.

Jim Holt, my quantum advisor, thinks taking PEP553 is dumb, because I’ll be forced to learn stuff I don’t need; I’ll bog down in technical details. Holt calls Griffiths the “Betty Crocker” of quantum instruction, because his textbook reduces quantum mechanics to recipes. Holt says I should stick to Susskind’s Quantum Mechanics: The Theoretical Minimum. I tell Holt, I’m committed to taking PEP553, can’t back out now. Holt’s discouragement makes me more determined.

On September 1, 2020, at 2 p.m., I attend my first session of PEP553 from the comfort of my apartment. Twenty-three students, counting me, Zoom in. Only a few have turned on their videos so we can see their faces. Ed Whittaker, our professor, looking cool in a short-sleeve shirt and white crewcut, takes us through introductions.  

About half of my classmate are graduate students in computer science and other fields. Many have taken courses in quantum mechanics before, and a few have already done quantum-optics and quantum-computing research. Even the undergraduates have years of math and physics under their belts. Three students are female, or so I infer from their names. The class is international; several students are zooming from China. Ed goes over homework, exams and the textbook. He plans to get through four of the twelve chapters in Griffiths. He doesn’t want to cram too much into the course; he wants us to get each section before he moves on.

Ed introduces “Professor Horgan.” He says I am a humanities professor at Stevens, as well as a science writer who wants to understand quantum mechanics better; that’s why I’m taking the class. Ed hopes I can provide insights into philosophical issues raised by quantum mechanics; in turn, maybe he and my classmates can help me understand the math. As often happens when someone introduces me to a group, I get a frisson of derealization, as if I’m looking at myself from the outside. I thank Ed for his gracious welcome and say I’ll try not to bog everyone down.

Ed projects seven questions on the screen, which comprise the “PEP553 Math and Mechanics Assessment.” The assessment says:

This quiz will help me gauge your knowledge of some math and mechanics ideas we will find useful. Please attempt to do all of the following problems… Your actual score on this assessment will not affect your grade in the course in any way but will be used to help the instructor tailor the course to class background.

Here is the first question:

The numbers θ and φ are real and i = square root of -1.

Evaluate to simplest form the following expressions:

Re(exp(iθ))=

Im(exp(iθ))=

||exp(iθ)||=

||exp(iθ)+exp(iφ)|| =

I know that the first question is about imaginary numbers, because it has an i in it, but I have no idea how to answer it. I’m guessing that exp means exponent, but I’m not sure. Questions 2-7 are equally baffling. Here is question 6:

For the function f (x, y, z) = x² + yz compute the following in Cartesian coordinates: ∇f =

I think ∇f means change in the value of f, but I have no idea how to solve the problem. I recognize derivatives, integrals and matrices in some questions, but other notations are foreign to me. Whittaker breaks us up into groups to work on the problems. I feel bad for my group, stuck with the old ignoramus. Later, Whittaker sends me an email asking my reaction to the class. I confess that I couldn’t answer any of the assessment questions. He tells me not to worry, I don’t have to get everything. Maybe I can focus on a few key concepts, like complex numbers.

I email the assessment quiz to a former student, Steve Sagona, who has been emailing me advice on my quantum project. Sagona took my science-writing seminar more than a decade ago when he was a physics major at Stevens. He pushed back against my end-of-science thesis, which made the class more fun, and we have stayed in touch. He is completing a doctorate in quantum information theory at the State University of New York at Stony Brook.

Sagona, who took a class with Ed Whittaker at Stevens and remembers him fondly, says I should take heart from the assessment quiz; Ed’s questions are easy, suggesting that he isn’t going to demand too much of us. Sagona offers to take me through the assessment quiz over Zoom; it will only take 10 minutes or so. I gratefully accept his offer.

Sagona looks like a mad monk, with a gaunt face, long, lank hair and a gleam in his eye. He badly underestimates the time needed to enlighten his old professor. We spend almost an hour on the first question, on complex numbers. He shows me the connection between cartesian and radial methods of representing complex vectors. Radial methods involve sines and cosines and angles, represented by θ and φ. Sagona explains how trigonometric functions are related to e and i.

I assume we’ll stop after the first question, but Sagona keeps charging ahead, dragging me with him. He takes me deep into partial derivatives, differential equations and Fourier transforms; the latter help you break a wave function down into other curves, which can be represented by sines and cosines. I nod along, feigning comprehension or mumbling, You can skip this, I’ll go over it later. I’ve encountered this material in Susskind, Quick Calculus and elsewhere, but I’ve already forgotten much of it. I forget that the derivative of eⁿ is also eⁿ. I forget the formulas for harmonic oscillators and Euler’s identity.

Simone Weil, 1909-1943, the radical activist and mystic, wrote about links between mathematics, morality and spirituality. Photo from Wikipedia.

I send the PEP553 quiz to my childhood friend Chris, who said on Facebook that I should be prepared for “humiliating failure and glorious success AT THE SAME TIME.” He’s a Yale math major who became a computer-graphic expert. He’s one of the smartest, and nicest, people I know. Chris says he had to “clear the cobwebs out” of his head to take the assessment quiz, but then the math came back to him. He remembered how beautiful the world of complex numbers can be, and how they are related to the Riemann hypothesis, one of the deepest unsolved riddles in mathematics.

I understand very little of what Chris says. To feel like I am contributing to the conversation, I tell him about André Weil, a mathematician whom I interviewed in 1994. Other mathematicians called Weil the last “universal mathematician,” an adept in all major fields. When I asked Weil if it bothered him that hardly anyone could understand his work, he seemed puzzled by the question. His work’s inaccessibility to others, Weil said, gave him joy. He achieved a state of “lucid exaltation” when he explored uncharted realms of mathematics. 

Weil was haughty man, a cold fish. His sister, Simone Weil, couldn’t have been more different, at least superficially. She was a philosopher, mystic and radical activist who was almost pathologically empathetic. She starved herself to death in 1943 as a protest against the suffering resulting from World War II. But in her writings, Simone Weil traced connections between mathematics and the realms of morality and spirituality. Perhaps the siblings weren’t so different after all.

Not a Boring Time

During our second session of PEP553, Professor Whittaker takes us through the assessment questions. When he gets to problem 4, which involves the function x (t ) = A cos(ω t ), Ed asks, What is this function? He adds with a grin, John, I’m not going to ask you. I tell him that x (t ) = A cos(ω t) is the formula for a harmonic oscillator. Ed exclaims, with what seems like genuine astonishment, Yes! Good! I feel absurdly proud, even though I know the answer only because my former student Steve Sagona and I went over it yesterday. When Ed shows how to solve the harmonic-oscillator equation, I can’t follow him; I’ve already forgotten what Sagona told me. Recognition isn’t comprehension.

We get to problem 7, which shows a weird blocky graph and asks: “Write down the first term F1(t) in the Fourier series for this function. Don’t worry about the constant coefficient if you can’t remember, just write down the functional part of the expression.” I have no idea what the problem is, let alone the solution; Sagona and I didn’t get to this problem yesterday.

Dean, one of the few students who keeps his video on, raises his hand. He has a kind, handsome face, with a beard and long hair; he looks like Jesus. Dean says you look for a sine wave that has the same periodicity as the blocky graph. Then you keep adding more sine waves to try to get closer to the original function. Yes, good, Whittaker says. This is Fourier analysis, which lets you combine the curves generated by sines and cosines to construct any other curve, no matter how complicated. Fourier analysis comes in handy in quantum mechanics.

Whittaker walks us through some quantum history, including Einstein’s work on the photoelectric effect, which helped establish that light comes in bundles of energy. He shows us a video of a double-slit experiment done by researchers at Hitachi, the Japanese electronics firm. Electrons strike the screen like particles, seemingly at random, but gradually they form bands, which you get when waves interfere. What’s going on? What does it mean?

Echoing Griffiths, Whittaker says that physicists have proposed various interpretations of quantum behavior; the dominant one is still the Copenhagen interpretation proposed by Niels Bohr almost a century ago. I chime in, noting that there has been tremendous interest lately in quantum interpretations; a lot is at stake, like whether reality exists when we don’t look at it. Whittaker nods and says, “This is not a boring time to be studying physics.” By the end of the class, I’m feeling confident, almost giddy. Maybe this course is doable!

Spock’s Stupidity

I toil over our first PEP553 homework assignment, which consists of problems in Chapter One of Griffiths (see notebook page at the end of the chapter). The first problem calls for making calculations related to a falling rock. Old-fashioned Newtonian physics! But I can’t do it, my math is too weak. When I skimmed Griffiths before the semester, I glanced at this problem and arrogantly assumed I could do it. I’m beginning to realize that in physics, real physics, everything is harder than it seems at first glance.

Griffiths reviews probability theory. It’s disgraceful that I, a so-called science writer, never learned probability theory, which is vitally important in science, not to mention everyday life. My 401K might be doing better if I knew a little probability theory! I don’t know the difference between “median” and “average,” or “variance” and “standard deviation.” Better to learn late than never, I suppose. Griffiths notes that you calculate a standard deviation by taking the “square root of the average of the square of the deviation from the average—gulp!”

I’m grateful for that “gulp!” And once I calculate a few standard deviations, I realize they aren’t that complicated. They’re intuitive, they make sense; they provide a good way to quantify uncertainty. On to Gaussian distributions. They represent variables that vary in a predictably random way. A Bell curve representing, say, the average height of American women is a classic Gaussian distribution. A variable that displays a Gaussian distribution is called a “normal deviate.” I think, I am a normal deviate. These self-asides keep me going.

Studying probability theory, I realize that physics was wracked with uncertainty long before Heisenberg formulated his principle. Although physics theories often convey a sense of precision and certainty, the experiments upon which theories are based are always imprecise; if data don’t have error bars, beware. Hence physicists’ emphasis on standard deviations, expectation values, Gaussian distributions and other tools designed to reveal the signal in the noise.

I remember something that has always bugged me about the original Star Trek. Spock, the science officer, is supposed to be extremely intelligent and “logical.” But he invokes probabilities in ways that are stupid and illogical. In the “Errand of Mercy” episode, Kirk and Spock are in a tight spot, a fight with Klingons, and Kirk asks Spock for the odds of “our getting out of here.”

Spock fetishizes mathematical precision in ways that are dumb.

Spock: Difficult to be precise, Captain. I should say approximately 7,824.7 to 1.

Kirk: Difficult to be precise? 7,824 to 1?

Spock: 7,824.7 to 1.

Kirk: That’s a pretty close approximation.

Spock: I endeavor to be accurate.

Spock’s precision is absurd, because it suggests that he knows the exact values of all the relevant variables bearing on their battle, which any competent scientist knows is impossible. Not to mention that whenever Spock provides these grim predictions, things turn out fine. If he is so smart, why doesn’t he adjust his forecasts, applying basic Bayesian reasoning?

At other times, Spock estimates how much time remains before something bad happens, and he gives the answer to the second. His precision is idiotic for two reasons: First, it takes Spock more than a second to voice the estimate, so it is immediately obsolete. Second, the ridiculously precise estimate wastes time in an emergency situation. I love Spock, but yeesh.

I move on to a homework problem involving those pesky, probabilistic wave functions. I’m supposed to calculate the value of a wave function at three separates times: t1, t2 and t3. Okay, let me think. The wave function consists of a probability amplitude expressed as a complex number. When you multiply that number by itself, or by its conjugate, you get the probability that the particle will be at that position at a particular time. Wait, what is a conjugate again? Oh yeah: you just reverse the sign of the imaginary part of the complex number. So the conjugate of the complex number a + bi is a – bi.

I still can’t figure out how to solve the wave function for those three different times, because my calculus is insufficient. I decide to cheat, to look at the solution that Ed posted on our course website, but the solution is long and complicated; it makes no sense to me. I’m struck by the difference between PEP553 and the courses I teach. There are clever and dumb ways to respond to Plato’s parable of the cave, but not right and wrong ways; you can interpret the parable as an anti-realist or realist screed. Humanities knowledge is soft, fuzzy, fakable, but I can’t bullshit my way through the problem of the falling rock in Griffiths, let alone the wave-function problem. Physics and math represent hard knowledge, which can’t be faked.

Hitler listening to his generals telling him that he might have wasted his time studying Bessel functions.

I tell my quantum advisor Chris Search, a professor of physics at Stevens, that the math in PEP553 is crushing me. Chris says students often feel overwhelmed by and resentful of the math they’re forced to learn. Chris says I should consider hiring a tutor to help me, maybe even one of my classmates. For my amusement, he sends me a mash-up of a scene from Downfall, a 2004 German film about Hitler’s final days. In the scene, Hitler’s generals give him bad news, and he explodes with rage. This scene has inspired many parodies with phony subtitles or dubbing.

In the mash-up that Chris sends me, the generals tell Hitler that Bessel functions, named after a 19th-century German mathematician, have limited applicability. Hitler, who has spent many hours studying Bessel functions, having been assured of their vital importance, loses his shit. I look up Bessel functions in Griffiths and find them on pages 140-141. The pages are covered with equations and graphs interspersed with a few sentences. The sentences are in English, but they might as well be Mayan.

Erroll Morris, Thomas Kuhn and Trump

Back for a moment to what is real. Politics has intruded on what was once a dry academic debate—laden with jargon like epistemology and ontology--about whether we can really know the world. Some intellectuals worry that skeptics, the kind who place scare quotes around “truth” and “knowledge,” make it easier for bad people to manipulate us.

Errol Morris has strong feelings about this debate. Morris is one of my favorite filmmakers; I often show my students The Fog of War, a documentary for which Morris won an Academy Award. As a young man, Morris considered becoming a philosopher, and he studied at Princeton under Thomas Kuhn, an influential anti-realist philosopher. Kuhn argues in his 1962 book The Structure of Scientific Revolutions that science can never achieve absolute, objective truth; reality is unknowable, forever hidden behind the veil of our preconceptions, or “paradigms.” 

I love this crazy book.

Morris ended up despising Kuhn, so much so that he gave up philosophy. Morris bashes Kuhn in his eccentric, brilliant book The Ashtray (Or the Man Who Denied Reality). The title alludes to an incident in which Kuhn threw an ashtray at him. Morris calls Kuhn a “maleficent deity,” who applied his skepticism to everyone but himself. The denial of objective truth, Morris says, “ultimately, perhaps irrevocably, undermines civilization.” Kuhn’s skepticism enabled the ascent of Trump: “I see a line from Kuhn to Karl Rove and Kelly Ann Conway and Donald Trump.” Karl Rove and Kelly Ann Conway are minions of George W. Bush and Trump, respectively.

In a column for Scientific American, I defended Kuhn, sort of, arguing that we need his sort of skepticism to protect us from our yearning for certitude. Morris, irritated by my column, emailed me, and we ended up chatting at length about Kuhn. While admitting that blaming Trump on Kuhn is a stretch, Morris held firm to his argument that anti-realist philosophies like Kuhn’s are dangerous.

Morris provoked me into digging up a transcript of my 1991 interview with Kuhn. Re-reading the transcript, I decide that Morris is right: Kuhn’s skepticism went too far. At one point during the interview, I suggested to Kuhn that his skepticism applied to fields with a “metaphysical” cast, like quantum mechanics, but not to more straightforward realms, like the study of infectious disease. As an example, I brought up AIDS. A few skeptics, notably virologist Peter Duesberg, were questioning whether the so-called human immunodeficiency virus, HIV, causes AIDS. These skeptics were either right or wrong, I said, not just right or wrong within the context of a particular social-cultural-linguistic paradigm. Kuhn shook his head vigorously and said: 

When this all comes out you’ll say, Boy, I see why [Duesberg] believed that, and he was onto something. I’m not going to tell you he was right, or he was wrong. We don’t believe any of that anymore. But neither do we believe anymore what these guys who said it was the cause believe… The question as to what AIDS is as a clinical condition and what the disease entity is itself is not--it is subject to adjustment. And so forth. When one learns to think differently about these things, if one does, the question of right and wrong will no longer seem to be the relevant question.

This was typical of how Kuhn spoke. As if to demonstrate his point that language obfuscates, he rarely expressed himself in a straightforward way. But he was saying that, even when it came to a question as seemingly straightforward—and vitally important!--as whether a particular virus causes AIDS, we cannot say what “truth” is. We can’t escape interpretation, subjectivity, cultural context, and hence we can never say whether a given claim is objectively right or wrong.

Kuhn’s prediction that Peter Duesberg would come to be seen as neither right nor wrong turned out to be wrong. Terribly wrong. The evidence that HIV causes AIDS is overwhelming, and the denial of the HIV/AIDS link is now viewed as morally as well as empirically wrong. In part because of Duesberg’s influence, the South African government withheld anti-retroviral medications from its citizens for years; that decision resulted in more than 330,000 avoidable deaths, according to one study.

Page from Quantum Notebook #2, September 6, 2020, on which I try to solve an easy probability problem in Griffiths: calculate the average age, median age and most probable age of a group of people with different ages. See last line.

Plain old common sense says we need realism, hopefully not too naïve, to keep us grounded in this crazy era, when people think climate change and the coronavirus pandemic are left-wing hoaxes. So I’m trying, as I learn more about imaginary numbers and wave functions, to hang onto my realism. Maybe the trick is to shut up and calculate. Shut up and do homework.

My brother Matt, an investment banker based in London, calls to catch up. We’ve always been competitive. I brag that I’m doing homework for my class on quantum mechanics. Good for you, Matt says. I ask if he knows what a standard deviation is; to my irritation, Matt nails the definition. I emphasize that I am studying other concepts much, much harder than standard deviations. Chuckling, Matt says it sounds like I enrolled in a graduate course in French literature without understanding French. My brother’s comment sticks in my head. If PEP553 is a reality check, so far I’m failing.

Then something good happens. I get an email from Dean. He is my PEP553 classmate, the Jesus lookalike who answered the assessment question about Fourier transforms. After Professor Whittaker introduced me to the class, Dean googled me and discovered that I’ve written about consciousness and the simulation hypothesis and other stuff in which he’s interested. He wonders whether I have time to talk about these subjects. Sure! I say. I add that I need help in the class. Would Dean consider tutoring me? I’d be happy to pay him. Dean says his field is computer science, not physics, and he won’t take money from me, but he’d be happy to be my study partner. Would Monday mornings at 11 work? Next week we can go over the first homework assignment. I reply, Hell yeah!

I have a study buddy.

Notes

  1. In a draft of this chapter that I sent to Ed Whittaker, I said “J s” stands for joules. Ed corrected me, using the notation ^ as a substitute for superscript to indicate “to the power of.” Ed wrote: “The units for h-bar are Joule-seconds or better Joule/Herz. Why is that? You are correct when you say that it enters into the energy formula for a photon. A photon of electromagnetic radiation with a frequency of 1 Hz (Herz is a unit of one per second) has energy of ~10^-34 Joules (h-bar times 1 Hz).  Ordinary visible light has a frequency of about 10^15 hz so one photon of that light has energy of about ~10^-19 Joules which is about a unit called an electron volt.  Every time such a photon hits your retina it sends an electrical signal down your optic nerve by pumping that much energy into the chemical components of the retina. Ok, maybe more than what you need, but at least I want you not to have unit errors which are the ultimate sin of physics majors LOL.”