The question of Abdus Salam ‘deserving’ his Nobel

Peter Woit has blogged about an oral history interview with theoretical physicist Sheldon Glashow published in 2020 by the American Institute of Physics. (They have a great oral history of physics series you should check out if you’re interested.) Woit zeroed in on a portion in which Glashow talks about his faltering friendship with Steven Weinberg and his issues with Abdus Salam’s nomination for the physics Nobel Prize.

Glashow, Weinberg and Salam together won this prize in 1979, for their work on the work on electroweak theory, which describes the behaviour of two fundamental forces, the electromagnetic force and the weak force. Glashow recalls that his and Weinberg’s friendship – having studied and worked together for many years – deteriorated in the 1970s, a time in which both scientists were aware that they were due a Nobel Prize. According to Glashow, however, Weinberg wanted the prize to be awarded only to himself and Salam.

This is presumably because of how the prize-winning work came to be: with Glashow’s mathematical-physical model published in 1960, Weinberg building on it seven years later, with Salam’s two relevant papers appeared a couple years after Glashow’s paper and a year after Weinberg’s. Glashow recalls that Salam’s work was not original, that each of his two papers respectively echoed findings already published in Glashow’s and Weinberg’s papers. Instead, Glashow continues, Salam received the Nobel Prize probably because he had encouraged his peers and his colleagues to nominate him a very large number of times and because he set up the International Centre for Theoretical Physics (ICTP) in Trieste.

This impression, of Salam being undeserving from a contribution-to-physics point of view in Glashow’s telling, is very at odds with the impression of Salam based on reading letters and comments by Weinberg and Pervez Hoodbhoy and by watching the documentary Salam – The First ****** Nobel Laureate.

The topic of Salam being a Nobel laureate was never uncomplicated, to begin with: he was an Ahmadi Muslim who enjoyed the Pakistan government’s support until he didn’t, when he was forced to flee the country; his intentions with the ICTP – to give scholars from developing countries a way to study physics without having to contend with often-crippling resource constrains – were also noble. Hoodbhoy has also written about the significance of Salam’s work as a physicist and the tragedy of his name and the memories of his contributions having been erased from all the prominent research centres in Pakistan.

Finally, one of Salam’s nominees for a Nobel Prize was the notable British physicist and Nobel laureate Paul A.M. Dirac, and it seems strange that Dirac would endorse Salam if he didn’t believe Salam’s work deserved it.

Bearing these facts in mind, Glashow’s contention appears to be limited to the originality of Salam’s work. But to my mind, even if Salam’s work was really derivative, it was at par with that of Glashow and Weinberg. More importantly, while I believe the Nobel Prizes deserve to be abrogated, the prize-giving committee did more good than it might have realised by including Salam among its winners: in the words of Weinberg, “Salam sacrificed a lot of possible scientific productivity by taking on that responsibility [to set up ICTP]. It’s a sacrifice I would not make.”

Glashow may not feel very well about Salam’s inclusion for the 1979 prize and the Nobel Prizes as we know are only happy to overlook anything other than the scientific work itself, but if the committee really screwed up, then they screwed up to do a good thing.

Then again, even though Glashow wasn’t alone (he was joined by Martinus J.G. Veltman on his opinions against Salam), the physicists’ community at large doesn’t share his views. Glashow also cites an infamous 2014 paper by Norman Dombey, in which Dombey concluded that Salam didn’t deserve his share of the prize, but the paper’s reputation itself is iffy at best.

In fact, this is all ultimately a pointless debate: there are just too many people who deserve a Nobel Prize but don’t win it while a deeper dive into the modern history of physics should reveal a near-constant stream of complaints against Nobel laureates and their work by their peers. It should be clear today that both winning a prize and not winning a prize ought to mean nothing to the practice of science.

The other remarkable thing about Glashow’s comments in the interview (as cited by Woit) is what I like to think of as the seemingly eternal relevance of Brian Keating’s change of mind. Brian Keating is an astrophysicist who was at the forefront of the infamous announcement that his team had discovered evidence of cosmic inflation, an epoch of the early universe in which it is believed to have expanded suddenly and greatly, in March 2014. There were many problems leading up to the announcement but there was little doubt at the time, and Keating also admitted later, that its rapidity was motivated by the temptation to secure a Nobel Prize.

Many journalists, scientists and others observers of the practice of science routinely and significantly underestimate the effect the Nobel Prizes exert on scientific research. The prospect of winning the prize for supposedly discovering evidence of cosmic inflation caused Keating et al. to not wait for additional, confirmatory data before making their announcement. When such data did arrive, from the Planck telescope collaboration, Keating et al. suffered for it with their reputation and prospects.

Similarly, Weinberg and Glashow fell out because, according to Glashow, Weinberg didn’t wish Glashow to give a talk in 1979 discussing possible alternatives to the work of Weinberg and Salam because Weinberg thought doing such a thing would undermine his and Salam’s chances of being awarded a Nobel Prize. Eventually it didn’t, but that’s beside the point: this little episode in history is as good an illustration as any of how the Nobel Prizes and their implied promises of laurels and prestige render otherwise smart scientists insecure, petty and elbows-out competitive – in exchange for sustaining an absurd and unjust picture of the scientific enterprise.

All of this goes obviously against the spirit of science.

The Nobel intent

You’ve probably tired of this but I can’t. The Nobel Prize folks just sent out a newsletter ahead of Women’s Day, on March 8, describing the achievements of female laureates of each of the six prizes. This is a customary exercise we’ve come to expect from organisations and companies trying to make themselves look good on the back of an occasion presumably designed to surmount the sort of barriers to women’s empowerment and professional success the organisations and companies often themselves perpetuate. For example, this the Nobel Prize newsletter shows off with some truly ironic language. Consider the notes accompanying the science prize winners:

“I remember being told over and over again: Women, you can do anything, so it never entered my mind that I couldn’t.” Donna Strickland was awarded the Nobel Prize in Physics 2018 for her work with laser pulses.

She was also the first female laureate of the physics prize in 55 years.

[Marie Curie] is the first Nobel Prize awarded woman and the only one to have received it in Physics as well as Chemistry.

… because the prize committee chose not to award anyone else.

Gert Cori was the first woman to receive a Medicine Prize.

… because the prize committee chose not to award anyone else.

This is also what baffles me, especially in October and December every year when the awards are announced and conferred, respectively: Why do people take an award seriously that doesn’t take their issues seriously? Any other institution that did the same thing, as well as self-aggrandise as often as it can, would’ve been derided, turned into memes even, but every year we – millions of Indians at least, scientists and non-scientists – look to the Nobel Prizes to acknowledge Indian contributions to science, missing entirely the point that the prizes are a deeply flawed human enterprise riven by their own (often Eurocentric) politics and that have no responsibility to be fair, and often aren’t.

Science and the scientist

Didier Queloz and Michel Mayor won the 2019 Nobel Prize for physics for discovering a famous exoplanet (51 Pegasi b) in 1995. Their claim was first verified by a top astronomer at the time named Geoff Marcy. He was later found guilty of having harassed many of his students between 2001 and 2010.

Azeen Ghorayshi of Buzzfeed News published an excellent thread detailing how Marcy’s star as an astronomer rose at a time coinciding with many of his transgressions. As Ghorayshi observes, “Marcy’s place in the science—in a buzzy field, and [with lots of money]—became part of the power used against them.” It wasn’t that Marcy would harass a woman and the woman would continue to be an astronomer; she would often leave the profession entirely.

This should make us wonder: if not for Marcy and numerous other researcher-teachers like him, what would all those strong, wonderful women (who finally outed him) have accomplished? The answer is likely lots. So the celebration of the work of men like Marcy doesn’t only concern whether a ‘morally innocent’ body of knowledge is ‘tainted’ by their actions as people but in fact strikes that moral neutrality down in two ways: the work gave Marcy power in the academic structure, and Marcy used that power to harass and drive women out of academia.

Ultimately what Marcy achieved and who Marcy is aren’t separate. The science and the scientist are inseparable – just different labels for the same entity at two points on a continuum, the same continuum that Richard Feynman lived on and which Jeffrey Epstein enabled.

John B. Goodenough, who won the 2019 chemistry Nobel Prize yesterday for his part in inventing the lithium-ion battery, has said scientists’ inventions are morally neutral. They’re not, but saying so spares one the responsibility of confronting the consequences of its use. Lithium-ion batteries may not seem to have many consequences of this sort because their use has become so prevalent, abstracted through many layers of industrialisation, but what if one of the laureates had harassed a colleague who could have contributed?

This is why Marcy’s work as an astronomer is also morally debilitated.

A revolutionary exoplanet

In 1992, Aleksander Wolszczan and Dale Frail became the first astronomers to publicly announce that they had discovered the first planets outside the Solar System, orbiting the dense core of a dead star about 2,300 lightyears away. This event is considered to be the first definitive detection of exoplanets, a portmanteau of extrasolar planets. However, Michel Mayor and Didier Queloz were recognised today with one half of the 2019 Nobel Prize for physics for discovering an exoplanet three years after Wolszczan and Frail did. This might be confusing – but it becomes clear once you stop to consider the planet itself.

51 Pegasi b orbits a star named 51 Pegasi about 50 lightyears away from Earth. In 1995, Queloz and Mayor were studying the light and other radiation coming from the star when they noticed that it was wobbling ever so slightly. By measuring the star’s radial velocity and using an analytical technique called Doppler spectroscopy, Queloz and Mayor realised there was a planet orbiting it. Further observations indicated that the planet was a ‘hot Jupiter’, a giant planet with a surface temperature of ~1,000º C orbiting really close to the star.

In 2017, Dutch and American astronomers studied the planet in even greater detail. They found its atmosphere was 0.01% water (a significant amount), it weighed about half as much as Jupiter and orbited 51 Pegasi once every four days.

This was surprising. 51 Pegasi is a Sun-like star, meaning its brightness and colour are similar to the Sun’s. However, this ‘foreign’ system looked nothing like our own Solar System. It contained a giant planet much like Jupiter but which was a lot closer to its star than Mercury is to the Sun.

Astronomers were startled because their ideas of what a planetary system should look like was based on what the Solar System looked like: the Sun at the centre, four rocky planets in the inner system, followed by gas- and ice-giants and then a large, ringed debris field in the form of an outer asteroid belt. Many researchers even thought hot Jupiters couldn’t exist. But the 51 Pegasi system changed all that.

It was so different that Queloz and Mayor were first met with some skepticism, including questions about whether they’d misread the data and whether the wobble they’d seen was some quirk of the star itself. However, as time passed, astronomers only became more convinced that they indeed had an oddball system on their hands. David Gray had penned a paper in 1997 arguing that 51 Pegasi’s wobble could be understood without requiring a planet to orbit it. He published another paper in 1998 correcting himself and lending credence to Queloz’s and Mayor’s claim. The duo received bigger support by inspiring other astronomers to take another look at their data and check if they’d missed any telltale signs of a planet. In time, they would discover more hot Jupiters, also called pegasean planets, orbiting conventional stars.

Through the next decade, it would become increasingly clear that the oddball system was in fact the Solar System. To date, astronomers have confirmed the existence of over 4,100 exoplanets. None of them belong to planetary systems that look anything like our own. More specifically, the Solar System appears to be unique because it doesn’t have any planets really close to the Sun; doesn’t have any planets heavier than Earth but lighter than Neptune – an unusually large mass gap; and most of whose planets revolve in nearly circular orbits.

Obviously the discovery forced astronomers to rethink how the Solar System could have formed versus how typical exoplanetary systems form. For example, scientists were able to develop two competing models for how hot Jupiters could have come to be: either by forming farther away from the host star and then migrating inwards or by forming much closer to the star and just staying there. But as astronomers undertook more observations of stars in the universe, they realised the region closest to the star often doesn’t have enough material to clump together to form such large planets.

Simulations also suggest than when a Jupiter-sized planet migrates from 5 AU to 0.1 AU, its passage could make way for Earth-mass planets to later form in the star’s habitable zone. The implication is that planetary systems that have hot Jupiters could also harbour potentially life-bearing worlds.

But there might not be many such systems. It’s notable that fewer than 10% of exoplanets are known to be hot Jupiters (only seven of them have an orbital period of less than one Earth-day). They’re just more prominent in the news as well as in the scientific literature because astronomers think they’re more interesting objects of study, further attesting to the significance of 51 Pegasi b. But even in their low numbers, hot Jupiters have been raising questions.

For example, according to data obtained by the NASA Kepler space telescope, which looked for the fleeting shadows that planets passing in front of their stars cast on the starlight, only 0.3-0.5% of the stars it observed had hot Jupiters. But observations using the radial velocity method, which Queloz and Mayor had also used in 1995, indicated a prevalence of 1.2%. Jason Wright, an astronomer at the Pennsylvania State University, wrote in 2012 that this discrepancy signalled a potentially deeper mystery: “It seems that the radial velocity surveys, which probe nearby stars, are finding a ‘hot-Jupiter rich’ environment, while Kepler, probing much more distant stars, sees lots of planets but hardly any hot Jupiters. What is different about those more distant stars? … Just another exoplanet mystery to be solved…”.

All of this is the legacy of the discovery of 51 Pegasi b. And given the specific context in which it was discovered and how the knowledge of its existence transformed how we think about our planetary neighbourhoods and neighbourhoods in other parts of the universe, it might be fair to say the Nobel Prize for Queloz and Mayor is in recognition of their willingness to stand by their data, seeing a planet where others didn’t.

The Wire
October 8, 2019

Disentangling entanglement

There has been considerable speculation if the winners of this year’s Nobel Prize for physics, due to be announced at 2.30 pm IST on October 8, will include Alain Aspect and Anton Zeilinger. They’ve both made significant experimental contributions related to quantum information theory and the fundamental nature of quantum mechanics, including entanglement.

Their work, at least the potentially prize-winning part of it, is centred on a class of experiments called Bell tests. If you perform a Bell test, you’re essentially checking the extent to which the rules of quantum mechanics are compatible with the rules of classical physics.

Whether or not Aspect, Zeilinger and/or others win a Nobel Prize this year, what they did achieve is worth putting in words. Of course, many other writers, authors, scientists, etc. have already performed this activity; I’d like to redo it if only because writing helps commit things to memory and because the various performers of Bell tests are likely to win some prominent prize, given how modern technologies like quantum cryptography are inflating the importance of their work, and at that time I’ll have ready reference material.

(There is yet another reason Aspect and Zeilinger could win a Nobel Prize. As with the medicine prizes, many of whose laureates previously won a Lasker Award, many of the physics laureates have previously won the Wolf Prize. And Aspect and Zeilinger jointly won the Wolf Prize for physics in 2010 along with John Clauser.)

The following elucidation is divided into two parts: principles and tests. My principal sources are Wikipedia, some physics magazines, Quantum Physics for Poets by Leon Lederman and Christopher Hill (2011), and a textbook of quantum mechanics by John L. Powell and Bernd Crasemann (1998).

§

Principles

From the late 1920s, Albert Einstein began to publicly express his discomfort with the emerging theory of quantum mechanics. He claimed that a quantum mechanical description of reality allowed “spooky” things that the rules of classical mechanics, including his theories of relativity, forbid. He further contended that both classical mechanics and quantum mechanics couldn’t be true at the same time and that there had to be a deeper theory of reality with its own, thus-far hidden variables.

Remember the Schrödinger’s cat thought experiment: place a cat in a box with a bowl of poison and close the lid; until you open the box to make an observation, the cat may be considered to be both alive and dead. Erwin Schrödinger came up with this example to ridicule the implications of Niels Bohr’s and Werner Heisenberg’s idea that the quantum state of a subatomic particle, like an electron, was described by a mathematical object called the wave function.

The wave function has many unique properties. One of these is superposition: the ability of an object to exist in multiple states at once. Another is decoherence (although this isn’t a property as much as a phenomenon common to many quantum systems): when you observed the object. it would probabilistically collapse into one fixed state.

Imagine having a box full of billiard balls, each of which is both blue and green at the same time. But the moment you open the box to look, each ball decides to become either blue or green. This (metaphor) is on the face of it a kooky description of reality. Einstein definitely wasn’t happy with it; he believed that quantum mechanics was just a theory of what we thought we knew and that there was a deeper theory of reality that didn’t offer such absurd explanations.

In 1935, Einstein, Boris Podolsky and Nathan Rosen advanced a thought experiment based on these ideas that seemed to yield ridiculous results, in a deliberate effort to provoke his ‘opponents’ to reconsider their ideas. Say there’s a heavy particle with zero spin – a property of elementary particles – inside a box in Bangalore. At some point, it decays into two smaller particles. One of these ought to have a spin of 1/2 and other of -1/2 to abide by the conservation of spin. You send one of these particles to your friend in Chennai and the other to a friend in Mumbai. Until these people observe their respective particles, the latter are to be considered to be in a mixed state – a superposition. In the final step, your friend in Chennai observes the particle to measure a spin of -1/2. This immediately implies that the particle sent to Mumbai should have a spin of 1/2.

If you’d performed this experiment with two billiard balls instead, one blue and one green, the person in Bangalore would’ve known which ball went to which friend. But in the Einstein-Podolsky-Rosen (EPR) thought experiment, the person in Bangalore couldn’t have known which particle was sent to which city, only that each particle existed in a superposition of two states, spin 1/2 and spin -1/2. This situation was unacceptable to Einstein because it was inimical certain assumptions on which the theories of relativity were founded.

The moment the friend in Chennai observed her particle to have spin -1/2, the one in Mumbai would have known without measuring her particle that it had a spin of 1/2. If it didn’t, the conservation of spin would be violated. If it did, then the wave function of the Mumbai particle would have collapsed to a spin 1/2 state the moment the wave function of the Chennai particle had collapsed to a spin -1/2 state, indicating faster-than-light communication between the particles. Either way, quantum mechanics could not produce a sensible outcome.

Two particles whose wave functions are linked the way they were in the EPR paradox are said to be entangled. Einstein memorably described entanglement as “spooky action at a distance”. He used the EPR paradox to suggest quantum mechanics couldn’t possibly be legit, certainly not without messing with the rules that made classical mechanics legit.

So the question of whether quantum mechanics was a fundamental description of reality or whether there were any hidden variables representing a deeper theory stood for nearly thirty years.

Then, in 1964, an Irish physicist at CERN named John Stewart Bell figured out a way to answer this question using what has since been called Bell’s theorem. He defined a set of inequalities – statements of the form “P is greater than Q” – that were definitely true for classical mechanics. If an experiment conducted with electrons, for example, also concluded that “P is greater than Q“, it would support the idea that quantum mechanics (vis-à-vis electrons) has ‘hidden’ parts that would explain things like entanglement more along the lines of classical mechanics.

But if an experiment couldn’t conclude that “P is greater than Q“, it would support the idea that there are no hidden variables, that quantum mechanics is a complete theory and, finally, that it implicitly supports spooky actions at a distance.

The theorem here was a statement. To quote myself from a 2013 post (emphasis added):

for quantum mechanics to be a complete theory – applicable everywhere and always – either locality or realism must be untrue. Locality is the idea that instantaneous or [faster-than-light] communication is impossible. Realism is the idea that even if an object cannot be detected at some times, its existence cannot be disputed [like electrons or protons].

Zeilinger and Aspect, among others, are recognised for having performed these experiments, called Bell tests.

Technological advancements through the late 20th and early 21st centuries have produced more and more nuanced editions of different kinds of Bell tests. However, one thing has been clear from the first tests, in 1981, to the last: they have all consistently violated Bell’s inequalities, indicating that quantum mechanics does not have hidden variables and our reality does allow bizarre things like superposition and entanglement to happen.

To quote from Quantum Physics for Poets (p. 214-215):

Bell’s theorem addresses the EPR paradox by establishing that measurements on object a actually do have some kind of instant effect on the measurement at b, even though the two are very far apart. It distinguishes this shocking interpretation from a more commonplace one in which only our knowledge of the state of b changes. This has a direct bearing on the meaning of the wave function and, from the consequences of Bell’s theorem, experimentally establishes that the wave function completely defines the system in that a ‘collapse’ is a real physical happening.


Tests

Though Bell defined his inequalities in such a way that they would lend themselves to study in a single test, experimenters often stumbled upon loopholes in the result as a consequence of the experiment’s design not being robust enough to evade quantum mechanics’s propensity to confound observers. Think of a loophole as a caveat; an experimenter runs a test and comes to you and says, “P is greater than Q but…”, followed by an excuse that makes the result less reliable. For a long time, physicists couldn’t figure out how to get rid of all these excuses and just be able to say – or not say – “P is greater than Q“.

If millions of photons are entangled in an experiment, the detectors used to detect, and observe, the photons may not be good enough to detect all of them or the photons may not survive their journey to the detectors properly. This fair-sampling loophole could give rise to doubts about whether a photon collapsed into a particular state because of entanglement or if it was simply coincidence.

To prevent this, physicists could bring the detectors closer together but this would create the communication loophole. If two entangled photons are separated by 100 km and the second observation is made more than 0.0003 seconds after the first, it’s still possible that optical information could’ve been exchanged between the two particles. To sidestep this possibility, the two observations have to be separated by a distance greater than what light could travel in the time it takes to make the measurements. (Alain Aspect and his team also pointed their two detectors in random directions in one of their tests.)

Third, physicists can tell if two photons received in separate locations were in fact entangled with each other, and not other photons, based on the precise time at which they’re detected. So unless physicists precisely calibrate the detection window for each pair, hidden variables could have time to interfere and induce effects the test isn’t designed to check for, creating a coincidence loophole.

If physicists perform a test such that detectors repeatedly measure the particles involved in, say, two labs in Chennai and Mumbai, it’s not impossible for statistical dependencies to arise between measurements. To work around this memory loophole, the experiment simply has to use different measurement settings for each pair.

Apart from these, experimenters also have to minimise any potential error within the instruments involved in the test. If they can’t eliminate the errors entirely, they will then have to modify the experimental design to compensate for any confounding influence due to the errors.

So the ideal Bell test – the one with no caveats – would be one where the experimenters are able to close all loopholes at the same time. In fact, physicists soon realised that the fair-sampling and communication loopholes were the more important ones.

In 1972, John Clauser and Stuart Freedman performed the first Bell test by entangling photons and measuring their polarisation at two separate detectors. Aspect led the first group that closed the communication loophole, in 1982; he subsequently conducted more tests that improved his first results. Anton Zeilinger and his team made advancements on the fair-sampling loophole.

One particularly important experimental result showed up in August 2015: Robert Hanson and his team at the Technical University of Delft, in the Netherlands, had found a way to close the fair-sampling and communication loopholes at the same time. To quote Zeeya Merali’s report in Nature News at the time (lightly edited for brevity):

The researchers started with two unentangled electrons sitting in diamond crystals held in different labs on the Delft campus, 1.3 km apart. Each electron was individually entangled with a photon, and both of those photons were then zipped to a third location. There, the two photons were entangled with each other – and this caused both their partner electrons to become entangled, too. … the team managed to generate 245 entangled pairs of electrons over … nine days. The team’s measurements exceeded Bell’s bound, once again supporting the standard quantum view. Moreover, the experiment closed both loopholes at once: because the electrons were easy to monitor, the detection loophole was not an issue, and they were separated far enough apart to close the communication loophole, too.

By December 2015, Anton Zeilinger and co. were able to close the communication and fair-sampling loopholes in a single test with a 1-in-2-octillion chance of error, using a different experimental setup from Hanson’s. In fact, Zeilinger’s team actually closed three loopholes including the freedom-of-choice loophole. According to Merali, this is “the possibility that hidden variables could somehow manipulate the experimenters’ choices of what properties to measure, tricking them into thinking quantum theory is correct”.

But at the time Hanson et al announced their result, Matthew Leifer, a physicist the Perimeter Institute in Canada, told Nature News (in the same report) that because “we can never prove that [the converse of freedom of choice] is not the case, … it’s fair to say that most physicists don’t worry too much about this.”

We haven’t gone into much detail about Bell’s inequalities themselves but if our goal is to understand why Aspect and Zeilinger, and Clauser too, deserve to win a Nobel Prize, it’s because of the ingenious tests they devised to test Bell’s, and Einstein’s, ideas and the implications of what they’ve found in the process.

For example, Bell crafted his test of the EPR paradox in the form of a ‘no-go theorem’: if it satisfied certain conditions, a theory was designated non-local, like quantum mechanics; if it didn’t satisfy all those conditions, the theory be classified as local, like Einstein’s special relativity. So Bell tests are effectively gatekeepers that can attest whether or not a theory – or a system – is behaving in a quantum way and each loophole is like an attempt to hack the attestation process.

In 1991, Artur Ekert, who would later be acknowledged as one of the inventors of quantum cryptography, realised this perspective could have applications in securing communications. Engineers could encode information in entangled particles, send them to remote locations, and allow detectors there to communicate with each other securely by observing these particles and decoding the information. The engineers can then perform Bell tests to determine if anyone might be eavesdropping on these communications using one or some of the loopholes.

Review: ‘Salam – The First ****** Nobel Laureate’ (2018)

Awards are elevated by their winners. For all of the Nobel Prizes’ flaws and shortcomings, they are redeemed by what its laureates choose to do with them. To this end, the Pakistani physicist and activist Abdus Salam (1926-1996) elevates the prize a great deal.

Salam – The First ****** Nobel Laureate is a documentary on Netflix about Salam’s life and work. The stars in the title stand for ‘Muslim’. The label has been censored because Salam belonged to the Ahmadiya sect, whose members are forbidden by law in Pakistan to call themselves Muslims.

After riots against this sect broke out in Lahore in 1953, Salam was forced to leave Pakistan, and he settled in the UK. His departure weighed heavily on him even though he could do very little to prevent it. He would return only in the early 1970s to assist Zulfiqar Ali Bhutto with building Pakistan’s first nuclear bomb. However, Bhutto would soon let the Pakistani government legislate against the Ahmadiya sect to appease his supporters. It’s not clear what surprised Salam more: the timing of India’s underground nuclear test or the loss of Bhutto’s support, both within months of each other, that had demoted him to a second-class citizen in his home country.

In response, Salam became more radical and reasserted his Muslim identity with more vehemence than he had before. He resigned from his position as scientific advisor to the president of Pakistan, took a break from physics and focused his efforts on protesting the construction of nuclear weapons everywhere.

It makes sense to think that he was involved. Someone will know. Whether we will ever get convincing evidence… who knows? If the Ahmadiyyas had not been declared a heretical sect, we might have found out by now. Now it is in no one’s interest to say he was involved – either his side or the government’s side. “We did it on our own, you know. We didn’t need him.”

Tariq Ali

Whether or not it makes sense, Salam himself believed he wouldn’t have solved the problems he did that won him the Nobel Prize if he hadn’t identified as Muslim.

If you’re a particle physicist, you would like to have just one fundamental force and not four. … If you’re a Muslim particle physicist, of course you’ll believe in this very, very strongly, because unity is an idea which is very attractive to you, culturally. I would never have started to work on the subject if I was not a Muslim.

Abdus Salam

This conviction unified at least in his mind the effects of the scientific, cultural and political forces acting on him: to use science as a means to inspire the Pakistani youth, and Muslim youth in general, to shed their inferiority complex, and his own longstanding desire to do something for Pakistan. His idea of success included the creation of more Muslim scientists and their presence in the ranks of the world’s best.

[Weinberg] How proud he was, he said, to be the first Muslim Nobel laureate. … [Isham] He was very aware of himself as coming from Pakistan, a Muslim. Salam was very ambitious. That’s why I think he worked so hard. You couldn’t really work for 15 hours a day unless you had something driving you, really. His work always hadn’t been appreciated, shall we say, by the Western world. He was different, he looked different. And maybe that also was the reason why he was so keen to get the Nobel Prize, to show them that … to be a Pakistani or a Muslim didn’t mean that you were inferior, that you were as good as anybody else.

The documentary isn’t much concerned with Salam’s work as a physicist, and for that I’m grateful because the film instead offers a view of his life that his identity as a figure of science often sidelines. By examining Pakistan’s choices through Salam’s eyes, we get a glimpse of a prominent scientist’s political and religious views as well – something that so many of us have become more reluctant to acknowledge.

Like with Srinivasa Ramanujan, one of whose theorems was incidentally the subject of Salam’s first paper, physicists saw a genius in Salam but couldn’t tell where he was getting his ideas from. Salam himself – like Ramanujan – attributed his prowess as a physicist to the almighty.

It’s possible the production was conceived to focus on the political and religious sides of a science Nobel laureate, but it puts itself at some risk of whitewashing his personality by consigning the opinions of most of the women and subordinates in his life to the very end of its 75-minute runtime. Perhaps it bears noting that Salam was known to be impatient and dismissive, sometimes even manipulative. He would get angry if he wasn’t being understood. His singular focus on his work forced his first wife to bear the burden of all household responsibilities, and he had difficulty apologising for his mistakes.

The physicist Chris Isham says in the documentary that Salam was always brimming with ideas, most of them bizarre, and that Salam could never tell the good ideas apart from the sillier ones. Michael Duff continues that being Salam’s student was a mixed blessing because 90% of his ideas were nonsensical and 10% were Nobel-Prize-class. Then, the producers show Salam onscreen talking about how physicists intend to understand the rules that all inanimate matter abides by:

To do this, what we shall most certainly need [is] a complete break from the past and a sort of new and audacious idea of the type which Einstein has had in the beginning of this century.

Abdus Salam

This echoes interesting but not uncommon themes in the reality of India since 2014: the insistence on certainty, the attacks on doubt and the declining freedom to be wrong. There are of course financial requirements that must be fulfilled (and Salam taught at Cambridge) but ultimately there must also be a political maturity to accommodate not just ‘unapplied’ research but also research that is unsure of itself.

With the exception of maybe North Korea, it would be safe to say no country has thus far stopped theoretical physicists from working on what they wished. (Benito Mussolini in fact setup a centre that supported such research in the late-1920s and Enrico Fermi worked there for a time.) However, notwithstanding an assurance I once received from a student at JNCASR that theoretical physicists need only a pen and paper to work, explicit prohibition may not be the way to go. Some scientists have expressed anxiety that the day will come if the Hindutvawadis have their way when even the fruits of honest, well-directed efforts are ridden with guilt, and non-applied research becomes implicitly disfavoured and discouraged.

Salam got his first shot at winning a Nobel Prize when he thought to question an idea that many physicists until then took for granted. He would eventually be vindicated but only after he had been rebuffed by Wolfgang Pauli, forcing him to drop his line of inquiry. It was then taken up and to its logical conclusion by two Chinese physicists, Tsung-Dao Lee and Chen-Ning Yang, who won the Nobel Prize for physics in 1957 for their efforts.

Whenever you have a good idea, don’t send it for approval to a big man. He may have more power to keep it back. If it’s a good idea, let it be published.

Abdus Salam

Salam would eventually win a Nobel Prize in 1979, together with Steven Weinberg and Sheldon Glashow – the same year in which Gen. Zia-ul-Haq had Bhutto hung to death after a controversial trial and set Pakistan on the road to Islamisation, hardening its stance against the Ahmadiya sect. But since the general was soon set to court the US against its conflict with the Russians in Afghanistan, he attempted to cast himself as a liberal figure by decorating Salam with the government’s Nishan-e-Imtiaz award.

Such political opportunism contrived until the end to keep Salam out of Pakistan even if, according to one of his sons, it “never stopped communicating with him”. This seems like an odd place to be in for a scientist of Salam’s stature, who – if not for the turmoil – could have been Pakistan’s Abdul Kalam, helping direct national efforts towards technological progress while also striving to be close to the needs of the people. Instead, as Pervez Hoodbhoy remarks in the documentary:

Salam is nowhere to be found in children’s books. There is no building named after him. There is no institution except for a small one in Lahore. Only a few have heard of his name.

Pervez Hoodbhoy

In fact, the most prominent institute named for him is the one he set up in Trieste, Italy, in 1964 (when he was 38): the Abdus Salam International Centre for Theoretical Physics. Salam had wished to create such an institution after the first time he had been forced to leave Pakistan because he wanted to support scientists from developing countries.

Salam sacrificed a lot of possible scientific productivity by taking on that responsibility. It’s a sacrifice I would not make.

Steven Weinberg

He also wanted the scientists to have access to such a centre because “USA, USSR, UK, France, Germany – all the rich countries of the world” couldn’t understand why such access was important, so refused to provide it.

When I was teaching in Pakistan, it became quite clear to me that either I must leave my country, or leave physics. And since then I resolved that if I could help it, I would try to make it possible for others in my situation that they are able to work in their own countries while still [having] access to the newest ideas. … What Trieste is trying to provide is the possibility that the man can still remain in his own country, work there the bulk of the year, come to Trieste for three months, attend one of the workshops or research sessions, meet the people in his subject. He had to go back charged with a mission to try to change the image of science and technology in his own country.

In India, almost everyone has heard of Rabindranath Tagore, C.V. Raman, Amartya Sen and Kailash Satyarthi. One reason our memories are so robust is that Jawaharlal Nehru – and “his insistence on scientific temper” – was independent India’s first prime minister. Another is that India has mostly had a stable government for the last seven decades. We also keep remembering those Nobel laureates because of what we think of the Nobel Prizes themselves. This perception is ill-founded at least as it currently stands: of the prizes as the ultimate purpose of human endeavour and as an institution in and of itself – when in fact it is just one recognition, a signifier of importance sustained by a bunch of Swedish men that has been as susceptible to bias and oversight as any other historically significant award has been.

However, as Salam (the documentary) so effectively reminds us, the Nobel Prize is also why we remember Abdus Salam, and not the many, many other Ahmadi Muslim scientists that Pakistan has disowned over the years, has never communicated with again and to whom it has never awarded the Nishan-e-Imtiaz. If Salam hadn’t won the Nobel Prize, would we think to recall the work of any of these scientists? Or – to adopt a more cynical view – would we have focused so much of our attention on Salam instead of distributing it evenly between all disenfranchised Ahmadi Muslim scholars?

One way or another, I’m glad Salam won a Nobel Prize. And one way or another, the Nobel Committee should be glad it picked Salam, too, for he elevated the prize to a higher place.

Note: The headline originally indicated the documentary was released in 2019. It was actually released in 2018. I fixed the mistake on October 6, 2019, at 8.45 am.

Gerald Guralnik (1936-2014)

Of the six scientists who came up with the idea of a Higgs boson in the mid-1960s, independently or in collaboration with others, I’ve met all of one. Tom Kibble was at the Institute of Mathematical Science, Chennai, in January 2013 for a conference. He was 80 years old then, and looked quite frail. Every time somebody tapped his shoulder before taking a photograph, he would break into a self-effacing smile. It was clear he was surprised by the attention he was receiving. Kibble thought he didn’t deserve it.

He, Carl Hagen and Gerald Guralnik comprised one of the three teams that conceived the mechanism to explain how some fundamental particles acquired mass in the early universe, over time making possible chemical reactions, stars, life, and many things besides. The other two teams comprised Francois Englert and Robert Brout, and Peter Higgs; Higgs’ name has today become attached to the name of the mechanism. For their work, Higgs and Englert were awarded the 2013 Nobel Prize in physics. Brout couldn’t receive the prize because he had died in 2011. Kibble, Hagen and Guralnik were left out because of limits on how many people the prize could be awarded to at a time.

Fair share of obstacles

On April 26, 2014, Gerald Guralnik died of a heart attack in Rhode Island after delivering a lecture at Brown University. He was 77. In those seven decades, he had become one of the world’s leading experts on theoretical particle physics, which, through the 1960s, was entering its boom time as the world would later discover. In this period, he co-scripted one of the most enduring quests in modern physics research.

Before I started writing this, I visited the Wikipedia page for the Physical Review Letters papers published by the three groups that first called the world’s attention to their findings. In the second line, Peter Higgs is mentioned as having worked with Satyen Bose – undoubtedly the consequence of a grave misapprehension that pervaded India when the 2013 Nobel Prizes were announced. Many believed Satyen Bose had been neglected for his work, but he just hadn’t worked on the Higgs boson, only on the underlying theory that controls the lives and times of all bosons. If such are the facile issues that concern some misguided Indians today, Guralnik tackled more than a fair share in his time.

sb1

For a few years after Kibble, Hagen and Guralnik published their paper, their work wasn’t taken seriously. Guralnik wrote in Huffington Post in August 2012 that, in the summer of 1965, Werner Heisenberg – the originator of the notorious uncertainty principle – thought Guralnik’s ideas were junk. The New York Times wrote that Robert Marshak, a famous theoretical physicist, told Guralnik that if he wished to survive in physics, he “must stop thinking about this sort of problem and move on,” advice that Guralnik “wisely obeyed”. According to Kibble, however, Marshak later admitted that he had been misguided.

Deference over primacy

Nevertheless, some other scientists had starting working on Guralnik & co.’s theories. By the 1970s, Sheldon Glashow, Abdus Salam and Steven Weinberg had succeeded in ironing out many of its inconsistencies and won the Nobel Prize for physics in 1979 for their work… even though it would be 50 more years to prove via experiment that the Higgs mechanism was for real. This is because there was no disputing that the implications of the work of Kibble, Hagen, Guralnik, Higgs, Brout and Englert were revolutionary, at least among those who were willing to accept it.

To this end, the 1979 prizewinners and the ‘Higgs Six’ were aware of and deferential toward the contributions of others to the development of this new theory. In fact, Higgs, who has often wound up being the centre of attention when talk of his eponymous mechanism comes up, has said that he’d rather call it the ABEGHHK’tH mechanism (A denoted Phillip Warren Anderson; ‘tH, Gerardus ‘t Hooft).

But others were less considerate, which didn’t go down well with Guralnik. As Kibble wrote in his obituary in Nature, “Guralnik came to feel that our early paper was often unfairly neglected. He gave talks and wrote papers pointing out our distinctive contribution, of which he was justifiably proud, and in which he was unquestionably the prime mover.” This doesn’t mean he went on to become a sour, old bat, of course, but only that Guralnik seemed to appreciate the gravitas of his work much more than others at the time. When  Higgs and Englert shared the 2013 Nobel Prize in physics, Guralnik told Brown Daily Herald that he was “a little hurt”, but happier for the recognition that his peers – and by extension his work – had received.

(It is, in fact, hard to say if he is as celebrated as Higgs is today, physicists notwithstanding. Such are the consequences of asymmetric recognition, a sort of ceiling effect that silences avant garde advancements until the world is ready to hear them. This is also a complaint I’ve heard from far too many Indian scientists and whose efforts to remedy it I don’t begrudge them even if it only seems like an infantile squabble over primacy.)

In fact, after his work in establishing the theoretical foundations of the Higgs mechanism, which itself is a cornerstone of a unified theory that describes both the electromagnetic and weak nuclear forces of nature, Guralnik proceeded to make a lot of other contributions. He worked on computational approaches to quantum field theory, quantum chromodynamics (i.e., the theory of the strong nuclear force), the application of chaos theory to particle physics, and string theory. His was a versatile genius, in part combative and in part pliant. Rest in peace.