“Why has no Indian won a science Nobel this year?”

For all their flaws, the science Nobel Prizes – at the time they’re announced, in the first week of October every year – provide a good opportunity to learn about some obscure part of the scientific endeavour with far-reaching consequences for humankind. This year, for example, we learnt about attosecond physics, quantum dots, and invitro transcribed mRNA. The respective laureates had roots in Austria, France, Hungary, Russia, Tunisia, and the U.S. Among the many readers that consume articles about these individuals’ work with any zest, the science Nobel Prizes’ announcement is also occasion for a recurring question: how come no scientist from India – such a large country, of so many people with diverse skills, and such heavy investments in research – has won a prize? I thought I’d jot down my version of the answer in this post. There are four factors:

1. Missing the forest for the trees – To believe that there’s a legitimate question in “why has no Indian won a science Nobel Prize of late?” is to suggest that we don’t consider what we read in the news everyday to be connected to our scientific enterprise. Pseudoscience and misinformation are almost everywhere you look. We’re underfunding education, most schools are short-staffed, and teachers are underpaid. R&D allocations by the national government have stagnated. Academic freedom is often stifled in the name of “national interest”. Students and teachers from the so-called ‘non-upper-castes’ are harassed even in higher education centres. Procedural inefficiencies and red tape constantly delay funding to young scholars. Pettiness and politicking rule many universities’ roosts. There are ill-conceived limits on the use, import, and export of biological specimens (and uncertainty about the state’s attitude to it). Political leaders frequently mock scientific literacy. In this milieu, it’s as much about having the resources to do good science as being able to prioritise science.

2. Historical backlog – This year’s science Nobel Prizes have been awarded for work that was conducted in the 1980s and 1990s. This is partly because the winning work has to have demonstrated that it’s of widespread benefit, which takes time (the medicine prize was a notable exception this year because the pandemic accelerated the work’s adoption), and partly because each prize most often – but not always – recognises one particular topic. Given that there are several thousand instances of excellent scientific work, it’s possible, on paper, for the Nobel Prizes to spend several decades awarding scientific work conducted in the 20th century alone. Recall that this was a boom time for science, with the advent of quantum mechanics and the theories of relativity, considerable war-time investment and government support, followed by revolutions in electronics, materials science, spaceflight, genetics, and pharmaceuticals, and then came the internet. It was also the time when India was finding its feet, especially until economic liberalisation in the early 1990s.

3. Lack of visibility of research – Visibility is a unifying theme of the Nobel laureates and their work. That is, you need to do good work as well as be seen to be doing that work. If you come up with a great idea but publish it in an obscure journal with no international readership, you will lose out to someone who came up with the same idea but later, and published it in one of the most-read journals in the world. Scientists don’t willingly opt for obscure journals, of course: publishing in better-read journals isn’t easy because you’re competing with other papers for space, the journals’ editors often have a preference for more sensational work (or sensationalisable work, such as a paper co-authored by an older Nobel laureate; see here), and publishing fees can be prohibitively high. The story of Meghnad Saha, who was nominated for a Nobel Prize but didn’t win, offers an archetypal example. How journals have affected the composition of the scientific literature is a vast and therefore separate topic, but in short, they’ve played a big part to skew it in favour of some kinds of results over others – even if they’re all equally valuable as scientific contributions – and to favour authors from some parts of the world over others. Journals’ biases sit on top of those of universities and research groups.

4. Award fixation – The Nobel Prizes aren’t interested in interrogating the histories and social circumstances in which science (that it considers to be prize-worthy) happens; they simply fete what is. It’s we who must grapple with the consequences of our histories of science, particularly science’s relationship with colonialism, and make reparations. Fixating on winning a science Nobel Prize could also lock our research enterprise – and the public perception of that enterprise – into a paradigm that prefers individual winners. The large international collaboration is a good example: When physicists working with the LHC found the Higgs boson in 2012, two physicists who predicted the particle’s existence in 1964 won the corresponding Nobel Prize. Similarly, when scientists at the LIGO detectors in the US first observed gravitational waves in 2016, three physicists who conceived of LIGO in the 1970s won the prize. Yet the LHC and the LIGOs, and other similar instruments continue to make important contributions to science – directly, by probing reality, and indirectly by supporting research that can be adapted for other fields. One 2007 paper also found that Nobel Prizes have been awarded to inventions only 23% of the time. Does that mean we should just focus on discoveries? That’s a silly way of doing science.


The Nobel Prizes began as the testament of a wealthy Swedish man who was worried about his legacy. He started a foundation that put together a committee to select winners of some prizes every year, with some cash from the man’s considerable fortunes. Over the years, the committee made a habit of looking for and selecting some of the greatest accomplishments of science (but not all), so much so that the laureates’ standing in the scientific community created an aspiration to win the prize. Many prizes begin like the Nobel Prizes did but become irrelevant because they don’t pay enough attention to the relationship between the laureate-selecting process and the prize’s public reputation (note that the Nobel Prizes acquired their reputation in a different era). The Infosys Prize has elevated itself in this way whereas the Indian Science Congress’s prize has undermined itself. India or any Indian for that matter can institute an award that chooses its winners more carefully, and gives them lots of money (which I’m opposed to vis-à-vis senior scientists) to draw popular attention.

There are many reasons an Indian hasn’t won a science Nobel Prize in a while but it’s not the only prize worth winning. Let’s aspire to other, even better, ones.

Reimagining science, redux

This article on Founding Fuel has some great suggestions I thought, but it merits sharing with a couple caveats.

First, in narratives about making science “easier to do”, commentators give science-industry linkages more play than science-society ones. This has been true in the past and continues to be. We remember and periodically celebrate the work of Shanti Swarup Bhatnagar and M. Visveshwaraya, but not with nearly equal fanfare that of, say, Yash Pal or the members of the Hoshangabad Science Teaching Programme.

In public dialogues about making the work of scientists more relevant, writers and TV panellists often touch on spending more money to setup larger, better supplied labs and improving ties between the labs and industry, where research is translated into product or service. Spending more on science is necessary, as is the need to support collaborations, regularise funding and grant-giving, improve working conditions for teachers, etc.

More broadly, I acknowledge that the problem is that there isn’t enough good science happening in the country, that the author is recommending various ways in which science-industry linkages and tweaks within the science ecosystem can both change this for the better, and that science-society linkages are unlikely to be of help on this front. However, could this be because we’re asking the wrong question?

That is, what science and industry can do for each other becomes relevant if what we’re seeking is the growth of science, as defined by some parameters (number of citations, number of patents, etc.), as an enterprise in and of itself – as if its fortunes and outcomes weren’t already yoked to other societal endeavours. Growth for growth’s sake. Science-society linkages become relevant on the other hand when the parameters are, say, research and academic liberties, extent of public participation, distribution of opportunities, freedom from government interference, etc. – when quantitative growth is both difficult and more aligned with nation-building.

Ultimately, we don’t need a science that becomes easier to do at the expense of not thinking about whether it needs to be done, or done differently. This is not a veiled comment against ‘blue sky’ research, which must continue, but is directed against ‘black sky’ research – which goes on to pollute our air and water, drills forestland for oil, dams rivers and destabilises ecosystems without thought for the consequences.

Nevertheless, in a system designed increasingly to incentivise working with the private sector, to self-finance one’s work through patents and other licenses, and to translate research into arbitrarily defined “useful” things, such thinking can only become more penalised, more unfavourable. And the science that is rolled into technologies will only be industry friendly, which in the current political climate means Ambani- and/or Adani-friendly, to the detriment of everyone else, especially those on the bottom rungs of society.

Second, the article’s author uses Nobel Prize-winning work to describe presumably the extent of what is possible when faculty members at an institute work together or when researchers collaborate with their younger peers. But in the process he frames ‘collaborations that produce Nobel Prizes’ as desirable. This is a problem because doing so overlooks collaborations that didn’t win Nobel Prizes, because laureates are often white men (non-white, non-cis-men may not be able to ‘breach’ such ‘in-groups’ because of structural factors even as solutions to break these barriers are ignored in favour of a flatter ‘prize-winning’ one), and because “Nobel-Prize-winning collaborations” is an oxymoron.

The last is easiest to see: the prizes are awarded only to three people at a time whereas the author himself quotes a study that found that the number of authors per scientific paper increased from 3.2 to 4.4 in 1996-2015.


As a corrective of sorts, to infuse deliberations prompted by the Founding Fuel article with what a focus on industry-oriented development leaves out, let me quote at length from an essay Mukund Thattai published with The Wire three years ago, exploring the existence of “an Indian way of doing science” (emphases mine):

There is a strong case to fund science for the same reason we fund the arts or sport. Science is a cultural activity: it reveals unexpected beauty in the everyday; it captures the imagination of children; it attempts to answer some of humanity’s biggest questions about where we came from. Moreover, scientific ideas can be a potent component of the process by which society arrives at collective decisions about the future. Among the strongest reasons a resource-limited country such as India should fund curiosity-driven science is that the nature of future crises cannot be predicted.

It is impossible to micromanage the long-term research agenda, so the only hope is to cast a wide net. A broad and deep scientific community is a valuable resource that can be called upon to give its inputs on a variety of issues. They cannot be expected to always deliver a solution but can be expected to provide the best possible information available at any time. In this consultative process, it is crucially important to not privilege scientific experts over other participants in the discussion.

… Science thrives within a diversity of questions and methods, a diversity of institutional environments, and a diversity of personal experiences of individual scientists. In the modern era, the practice of science has moved to a more democratic mode, away from the idea of lone geniuses and towards a collective effort of creating hypotheses and sharing results. Any tendency toward uniformity and career professionalisation dilutes and ultimately destroys this diversity. As historian of science Dhruv Raina describes it, a science that is vulnerable to the “pressures of government” is “no longer an open frontier of critical activity”. Instead, science must become “social and reflexive”.

Ideas and themes must bubble up from the broadest possible community. In India, access to such a process is limited by the accident of one’s mother tongue and social class, and this must change. Anyone who wants to should have the opportunity to understand what scientists are doing. Ultimately, this must involve not only scientists but also social scientists, historians, philosophers, artists and communicators – and the public at large.

… Is there such a thing as an “Indian way” of doing science? Science in the abstract is said to transcend national boundaries. In practice it is strongly influenced by local experiences and local history. Unfortunately, even as national missions have faded to the background, they have been replaced by an imitation of Western fashions. It has become common to look to high-profile journals and conferences as arbiters of questions worth asking. This must stop. The key to revitalising Indian science is the careful choice of rich questions. These questions could be driven by new national missions that bring the excitement of a collective effort. Or they could be inspired by observing the complex interactions of the world immediately around us.

There is a great deal of scholarship and scientific inquiry that can arise from the study of India’s traditional knowledge systems. The country’s enormous biodiversity and human genetic diversity are an exciting and bottomless source of scientific puzzles and important secrets. Such questions would allow for a deeper two-way engagement with India’s people. This is not to say Indian scientists cannot work on internationally important problems – quite the opposite. The scientific community in India, working within their own unique contexts, could become the source of important problems that anyone in the world would be excited to work on.

… The internationalisation of science is an important goal in and of itself. While it stimulates cross-fertilisation of ideas and pushes up standards within science, it also creates opportunities for broader global discussions and engagements. The unfortunate hurdles which curtail the ability of Indian academics and students to travel abroad, and the enormous difficulty foreign academics face in obtaining necessary permissions to visit their colleagues in India, serve no purpose. In spite of all this, there is a healthy trend towards stronger international links.

Academic scientists have long played dual roles as teachers and researchers. Within India, science has a remarkably broad appeal. Public science talks are standing-room-only affairs, and famous scientists receive the kind of adulation typically reserved for movie stars. Students across the country are excited about science. Many aspire to become scientists themselves.

Historically, engineering and medical colleges have attracted scientifically-minded students, but this is changing. The Indian Institutes of Science Education and Research have now been running undergraduate programs for over a decade in cities across India. These institutions are to science what the IITs are to engineering, attracting some of the brightest students each year. Science programs within public universities have not fared as well, and must seize every opportunity to reinvent themselves. A science curriculum based not on dry facts but on the history and process of discovery can form the base of a broad education, in conjunction with the humanities and the arts.

Freeman Dyson’s PhD

The physicist, thinker and writer Freeman Dyson passed away on February 28, 2020, at the age of 96. I wrote his obituary for The Wire Science; excerpt:

The 1965 Nobel Prize for the development of [quantum electrodynamics] excluded Dyson. … If this troubled Dyson, it didn’t show; indeed, anyone who knew him wouldn’t have expected differently. Dyson’s life, work, thought and writing is a testament to a philosophy of doing science that has rapidly faded through the 20th century, although this was due to an unlikely combination of privileges. For one, in 1986, he said of PhDs, “I think it’s a thoroughly bad system, so it’s not quite accidental that I didn’t get one, but it was convenient.” But he also admitted it was easier for him to get by without a PhD.

His QED paper, together with a clutch of others in mathematical physics, gave him a free-pass to more than just dabble in a variety of other interests, not all of them related to theoretical physics and quite a few wandering into science fiction. … In 1951, he was offered a position to teach at Cornell even though he didn’t have a doctorate.

Since his passing, many people have latched on to the idea that Dyson didn’t care for awards and that “he didn’t even bother getting a PhD” as if it were a difficult but inspiring personal choice, and celebrate it. It’s certainly an unlikely position to assume and makes for the sort of historical moment that those displeased with the status quo can anchor themselves to and swing from for reform, considering the greater centrality of PhDs to the research ecosystem together with the declining quality of PhD theses produced at ‘less elite’ institutions.

This said, I’m uncomfortable with such utterances when they don’t simultaneously acknowledge the privileges that secured for Dyson his undoubtedly deserved place in history. Even a casual reading of Dyson’s circumstances suggests he didn’t have to complete his doctoral thesis (under Hans Bethe at Cornell University) because he’d been offered a teaching position on the back of his contributions to the theory of quantum electrodynamics, and was hired by the Institute for Advanced Study in Princeton a year later.

It’s important to mention – and thus remember – which privileges were at play so that a) we don’t end up unduly eulogising Dyson, or anyone else, and b) we don’t attribute Dyson’s choice to his individual personality alone instead of also admitting the circumstances Dyson was able to take for granted and which shielded him from adverse consequences. He “didn’t bother getting a PhD” because he wasn’t the worse for it; in one interview, he says he feels himself “very lucky” he “didn’t have to go through it”. On the other hand, even those who don’t care for awards today are better off with one or two because:

  • The nature of research has changed
  • Physics has become much more specialised than it was in 1948-1952
  • Degrees, grants, publications and awards have become proxies for excellence when sifting through increasingly overcrowded applicants’ pools
  • Guided by business decisions, journals definition of ‘good science’ has changed
  • Vannevar Bush’s “free play of free intellects” paradigm of administering research is much less in currency
  • Funding for science has dropped, partly because The War ended, and took a chunk of administrative freedom with it

The expectations of scientists have also changed. IIRC Dyson didn’t take on any PhD students, perhaps as a result of his dislike for the system (among other reasons because he believed it penalises students not interested in working on a single problem for many years at a time). But considering how the burdens on national education systems have shifted, his decision would be much harder to sustain today even if all of the other problems didn’t exist. Moreover, he has referred to his decision as a personal choice – that it wasn’t his “style” – so treating it as a prescription for others may mischaracterise the scope and nature of his disagreement.

However, questions about whether Dyson might have acted differently if he’d had to really fight the PhD system, which he certainly had problems with, are moot. I’m not discussing his stomach for a struggle nor am I trying to find fault with Dyson’s stance; the former is a pointless consideration and the latter would be misguided.

Instead, it seems to me to be a question of what we do know: Dyson didn’t get a PhD because he didn’t have to. His privileges were a part of his decision and cemented its consequences, and a proper telling of the account should accommodate them even if only to suggest a “Dysonian pride” in doing science requires a strong personality as well as a conspiracy of conditions lying beyond the individual’s control, and to ensure reform is directed against the right challenges.

Featured image: Freeman Dyson, October 2005. Credit: ioerror/Wikimedia Commons, CC BY-SA 2.0.

Why are the Nobel Prizes still relevant?

Note: A condensed version of this post has been published in The Wire.

Around this time last week, the world had nine new Nobel Prize winners in the sciences (physics, chemistry and medicine), all but one of whom were white and none were women. Before the announcements began, Göran Hansson, the Swede-in-chief of these prizes, had said the selection committee has been taking steps to make the group of laureates more racially and gender-wise inclusive, but it would seem they’re incremental measures, as one editorial in the journal Nature pointed out.

Hansson and co. seems to find the argument that the Nobel Prizes award achievements at a time where there weren’t many women in science tenable when in fact it distracts from the selection committee’s bizarre oversight of such worthy names as Lise Meitner, Vera Rubin, Chien-Shiung Wu, etc. But Hansson needs to understand that the only meaningful change is change that happens right away because, even for this significant flaw that should by all means have diminished the prizes to a contest of, for and by men, the Nobel Prizes have only marginally declined in reputation.

Why do they matter when they clearly shouldn’t?

For example, according to the most common comments received in response to articles by The Wire shared on Twitter and Facebook, and always from men, the prizes reward excellence, and excellence should brook no reservation, whether by caste or gender. As is likely obvious to many readers, this view of scholastic achievement resembles a blade of grass: long, sprouting from the ground (the product of strong roots but out of sight, out of mind), rising straight up and culminating in a sharp tip.

However, achievement is more like a jungle: the scientific enterprise – encompassing research institutions, laboratories, the scientific publishing industry, administration and research funding, social security, availability of social capital, PR, discoverability and visibility, etc. – incorporates many vectors of bias, discrimination and even harassment towards its more marginalised constituents. Your success is not your success alone; and if you’re an upper-caste, upper-class, English-speaking man, you should ask yourself, as many such men have been prompted to in various walks of life, who you might have displaced.

This isn’t a witch-hunt as much as an opportunity to acknowledge how privilege works and what we can do to make scientific work more equal, equitable and just in future. But the idea that research is a jungle and research excellence is a product of the complex interactions happening among its thickets hasn’t found meaningful purchase, and many people still labour with a comically straightforward impression that science is immune to social forces. Hansson might be one of them if his interview to Nature is anything to go by, where he says:

… we have to identify the most important discoveries and award the individuals who have made them. If we go away from that, then we’ve devalued the Nobel prize, and I think that would harm everyone in the end.

In other words, the Nobel Prizes are just going to look at the world from the top, and probably from a great distance too, so the jungle has been condensed to a cluster of pin-pricks.

Another reason why the Nobel Prizes haven’t been easy to sideline is that the sciences’ ‘blade of grass’ impression is strongly historically grounded, with help from notions like scientific knowledge spreads from the Occident to the Orient.

Who’s the first person that comes to mind when I say “Nobel Prize for physics”? I bet it’s Albert Einstein. He was so great that his stature as a physicist has over the decades transcended his human identity and stamped the Nobel Prize he won in 1921 with an indelible mark of credibility. Now, to win a Nobel Prize in physics is to stand alongside Einstein himself.

This union between a prize and its laureate isn’t unique to the Nobel Prize or to Einstein. As I’ve said before, prizes are elevated by their winners. When Margaret Atwood wins the Booker Prize, it’s better for the prize than it is for her; when Isaac Asimov won a Hugo Award in 1963, near the start of his career, it was good for him, but it was good for the prize when he won it for the sixth time in 1992 (the year he died). The Nobel Prizes also accrued a substantial amount of prestige this way at a time when it wasn’t much of a problem, apart from the occasional flareup over ignoring deserving female candidates.

That their laureates have almost always been from Europe and North America further cemented the prizes’ impression that they’re the ultimate signifier of ‘having made it’, paralleling the popular undercurrent among postcolonial peoples that science is a product of the West and that they’re simply its receivers.

That said, the prize-as-proxy issue has contributed considerably as well to preserving systemic bias at the national and international levels. Winning a prize (especially a legitimate one) accords the winner’s work with a modicum of credibility and the winner, of prestige. Depending on how the winners of a prize to be awarded suitably in the future are to be selected, such credibility and prestige could be potentiated to skew the prize in favour of people who have already won other prizes.

For example, a scientist-friend ranted to me about how, at a conference he had recently attended, another scientist on stage had introduced himself to his audience by mentioning the impact factors of the journals he’d had his papers published in. The impact factor deserves to die because, among other reasons, it attempts to condense multi-dimensional research efforts and the vagaries of scientific publishing into a single number that stands for some kind of prestige. But its users should be honest about its actual purpose: it was designed so evaluators could take one look at it and decide what to do about a candidate to whom it corresponded. This isn’t fair – but expeditiousness isn’t cheap.

And when evaluators at different rungs of the career advancement privilege the impact factor, scientists with more papers published earlier in their careers in journals with higher impact factors become exponentially likelier to be recognised for their efforts (probably even irrespective of their quality given the unique failings of high-IF journals, discussed here and here) over time than others.

Brian Skinner, a physicist at Ohio State University, recently presented a mathematical model of this ‘prestige bias’ and whose amplification depended in a unique way, according him, on a factor he called the ‘examination precision’. He found that the more ambiguously defined the barrier to advancement is, the more pronounced the prestige bias could get. Put another way, people who have the opportunity to maintain systemic discrimination simultaneously have an incentive to make the points of entry into their club as vague as possible. Sound familiar?

One might argue that the Nobel Prizes are awarded to people at the end of their careers – the average age of a physics laureate is in the late 50s; John Goodenough won the chemistry prize this year at 97 – so the prizes couldn’t possibly increase the likelihood of a future recognition. But the sword cuts both ways: the Nobel Prizes are likelier than not to be the products a prestige bias amplification themselves, and are therefore not the morally neutral symbols of excellence Hansson and his peers seem to think they are.

Fourth, the Nobel Prizes are an occasion to speak of science. This implies that those who would deride the prizes but at the same time hold them up are equally to blame, but I would agree only in part. This exhortation to try harder is voiced more often than not by those working in the West, with publications with better resources and typically higher purchasing power. On principle I can’t deride the decisions reporters and editors make in the process of building an audience for science journalism, with the hope that it will be profitable someday, all in a resource-constrained environment, even if some of those choices might seem irrational.

(The story of Brian Keating, an astrophysicist, could be illuminating at this juncture.)

More than anything else, what science journalism needs to succeed is a commonplace acknowledgement that science news is important – whether it’s for the better or the worse is secondary – and the Nobel Prizes do a fantastic job of getting the people’s attention towards scientific ideas and endeavours. If anything, journalists should seize the opportunity in October every year to also speak about how the prizes are flawed and present their readers with a fuller picture.

Finally, and of course, we have capitalism itself – implicated in the quantum of prize money accompanying each Nobel Prize (9 million Swedish kronor, Rs 6.56 crore or $0.9 million).

Then again, this figure pales in comparison to the amounts that academic institutions know they can rake in by instrumentalising the prestige in the form of donations from billionaires, grants and fellowships from the government, fees from students presented with the tantalising proximity to a Nobel laureate, and in the form of press coverage. L’affaire Epstein even demonstrated how it’s possible to launder a soiled reputation by investing in scientific research because institutions won’t ask too many questions about who’s funding them.

The Nobel Prizes are money magnets, and this is also why winning a Nobel Prize is like winning an Academy Award: you don’t get on stage without some lobbying. Each blade of grass has to mobilise its own PR machine, supported in all likelihood by the same institute that submitted their candidature to the laureates selection committee. The Nature editorial called this out thus:

As a small test case, Nature approached three of the world’s largest international scientific networks that include academies of science in developing countries. They are the International Science Council, the World Academy of Sciences and the InterAcademy Partnership. Each was asked if they had been approached by the Nobel awarding bodies to recommend nominees for science Nobels. All three said no.

I believe those arguments that serve to uphold the Nobel Prizes’ relevance must take recourse through at least one of these reasons, if not all of them. It’s also abundantly clear that the Nobel Prizes are important not because they present a fair or useful picture of scientific excellence but in spite of it.

Are the papers behind this year’s Nobel Prizes in the public domain?

Note: One of my editors thought this post would work for The Wire as well, so it’s been republished there.

“… for the greatest benefit of mankind” – these words are scrawled across a banner that adorns the Nobel Prize’s homepage. They are the words of Alfred Nobel, who instituted the prizes and bequeathed his fortunes to run the foundation that awards them. The words were chosen by the prize’s awarders to denote the significance of their awardees’ accomplishments.

However, the scientific papers that first described these accomplishments in the technical literature are often not available in the public domain. They languish behind paywalls erected by the journals that publish them, that seek to cash in on their importance to the advancement of science. Many of these papers are also funded by public money, but that hasn’t deterred journals and their publishers from keeping the papers out of public reach. How then can they be for the greatest benefit of mankind?

§

I’ve listed some of the more important papers published by this year’s laureates; they describe work that earned them their respective prizes. Please remember that my choice of papers is selective; where I have found other papers that are fully accessible – or otherwise – I have provided a note. This said, I picked the papers from the scientific background document first and then checked if they were accessible, not the other way round. (If you, whoever you are, are interested in replicating my analysis but more thoroughly, be my guest; I will help you in any way I can.)

A laureate may have published many papers collectively for which he was awarded (this year’s science laureates are all male). I’ve picked the papers most proximate to their citation from the references listed in the ‘advanced scientific background’ section available for each prize on the Nobel Prize website. Among publishers, the worst offender appears – to no one’s surprise – to be Elsevier.

A paper title in green indicates it’s in the public domain; red indicates it isn’t – both on the pages of the journal itself. Some titles in red maybe available in full elsewhere, such as in university archives. The names of laureates in the papers’ citations are underlined.

Physiology/medicine

“for their discoveries of molecular mechanisms controlling the circadian rhythm”

The paywall for papers by Young and Rosbash published in Nature were lifted by the journal on the day their joint Nobel Prize was announced. Until then, they’d been inaccessible to the general public. Interestingly, both papers acknowledge funding grants from the US National Institutes of Health, a tax-funded body of the US government.

Michael Young

Restoration of circadian behavioural rhythms by gene transfer in Drosophila – Nature 312, 752 – 754 (20 December 1984); doi:10.1038/312752a0 link

Isolation of timeless by PER protein interaction: defective interaction between timeless protein and long-period mutant PERL – Gekakis, N., Saez, L., Delahaye-Brown, A.M., Myers, M.P., Sehgal, A., Young, M.W., and Weitz, C.J. (1995). Science 270, 811–815. link

Michael Rosbash

Feedback of the Drosophila period gene product on circadian cycling of its messenger RNA levels – Nature 343, 536 – 540 (08 February 1990); doi:10.1038/343536a0 link

The period gene encodes a predominantly nuclear protein in adult Drosophila – Liu, X., Zwiebel, L.J., Hinton, D., Benzer, S., Hall, J.C., and Rosbash, M. (1992). J Neurosci 12, 2735–2744. link

Jeffrey Hall

Molecular analysis of the period locus in Drosophila melanogaster and identification of a transcript involved in biological rhythms – Reddy, P., Zehring, W.A., Wheeler, D.A., Pirrotta, V., Hadfield, C., Hall, J.C., and Rosbash, M. (1984). Cell 38, 701–710. link

P-element transformation with period locus DNA restores rhythmicity to mutant, arrhythmic Drosophila melanogaster – Zehring, W.A., Wheeler, D.A., Reddy, P., Konopka, R.J., Kyriacou, C.P., Rosbash, M., and Hall, J.C. (1984). Cell 39, 369–376. link

Antibodies to the period gene product of Drosophila reveal diverse tissue distribution and rhythmic changes in the visual system – Siwicki, K.K., Eastman, C., Petersen, G., Rosbash, M., and Hall, J.C. (1988). Neuron 1, 141–150. link

Physics

“for decisive contributions to the LIGO detector and the observation of gravitational waves”

While results from the LIGO detector were published in peer-reviewed journals, the development of the detector itself was supported by personnel and grants from MIT and Caltech. As a result, the Nobel laureates’ more important contributions were published as a reports since archived by the LIGO collaboration and made available in the public domain.

Rainer Weiss

Quarterly progress reportR. Weiss, MIT Research Lab of Electronics 105, 54 (1972) link

The Blue BookR. Weiss, P.R. Saulson, P. Linsay and S. Whitcomb link

Chemistry

“for developing cryo-electron microscopy for the high-resolution structure determination of biomolecules in solution”

The journal Cell, in which the chemistry laureates appear to have published many papers, publicised a collection after the Nobel Prize was announced. Most papers in the collection are marked ‘Open Archive’ and are readable in full. However, the papers cited by the Nobel Committee in its scientific background document don’t appear there. I also don’t know whether the papers in the collection available in full were always available in full.

Jacques Dubochet

Cryo-electron microscopy of vitrified specimens – Dubochet, J., Adrian, M., Chang, J.-J., Homo, J.-C., Lepault, J., McDowall, A. W., and Schultz, P. (1988). Q. Rev. Biophys. 21, 129-228 link

Vitrification of pure water for electron microscopyDubochet, J., and McDowall, A. W. (1981). J. Microsc. 124, 3-4 link

Cryo-electron microscopy of viruses – Adrian, M., Dubochet, J., Lepault, J., and McDowall, A. W. (1984). Nature 308, 32-36 link

Joachim Frank

Averaging of low exposure electron micrographs of non-periodic objectsFrank, J. (1975). Ultramicroscopy 1, 159-162 link

Three-dimensional reconstruction from a single-exposure, random conical tilt series applied to the 50S ribosomal subunit of Escherichia coli – Radermacher, M., Wagenknecht, T., Verschoor, A., and Frank, J. (1987). J. Microsc. 146, 113-136 link

SPIDER-A modular software system for electron image processingFrank, J., Shimkin, B., and Dowse, H. (1981). Ultramicroscopy 6, 343-357 link

Richard Henderson

Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopyHenderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, E., and Downing, K. H. (1990). J. Mol. Biol. 213, 899-929 link

The potential and limitations of neutrons, electrons and X-rays for atomic resolution microscopy of unstained biological moleculesHenderson, R. (1995). Q. Rev. Biophys. 28, 171-193 link (available in full here)

§

By locking the red-tagged papers behind a paywall – often impossible to breach because of the fees involved – they’re kept out of hands of less-well-funded institutions and libraries, particularly researchers in countries whose currencies have lower purchasing power. More about this here and here. But the more detestable thing with the papers listed above is that the latest of them (among the reds) was published in 1995, fully 22 years ago, and the earliest, 42 years go – both on cryo-electron microscopy. Both represent almost unforgivable durations across which to have paywalls, with the journals Nature and Cell further attempting to ride the Nobel wave for attention. It’s not clear if the papers they’ve liberated from behind the paywall will always be available for free hence either.

Read all this in the context of the Nobel Prizes not being awarded to more than three people at a time and maybe you’ll see how much of scientific knowledge is truly out of bounds of most of humankind.

Featured image credit: Pexels/pixabay.

Are the papers behind this year's Nobel Prizes in the public domain?

Note: One of my editors thought this post would work for The Wire as well, so it’s been republished there.

“… for the greatest benefit of mankind” – these words are scrawled across a banner that adorns the Nobel Prize’s homepage. They are the words of Alfred Nobel, who instituted the prizes and bequeathed his fortunes to run the foundation that awards them. The words were chosen by the prize’s awarders to denote the significance of their awardees’ accomplishments.

However, the scientific papers that first described these accomplishments in the technical literature are often not available in the public domain. They languish behind paywalls erected by the journals that publish them, that seek to cash in on their importance to the advancement of science. Many of these papers are also funded by public money, but that hasn’t deterred journals and their publishers from keeping the papers out of public reach. How then can they be for the greatest benefit of mankind?

§

I’ve listed some of the more important papers published by this year’s laureates; they describe work that earned them their respective prizes. Please remember that my choice of papers is selective; where I have found other papers that are fully accessible – or otherwise – I have provided a note. This said, I picked the papers from the scientific background document first and then checked if they were accessible, not the other way round. (If you, whoever you are, are interested in replicating my analysis but more thoroughly, be my guest; I will help you in any way I can.)

A laureate may have published many papers collectively for which he was awarded (this year’s science laureates are all male). I’ve picked the papers most proximate to their citation from the references listed in the ‘advanced scientific background’ section available for each prize on the Nobel Prize website. Among publishers, the worst offender appears – to no one’s surprise – to be Elsevier.

A paper title in green indicates it’s in the public domain; red indicates it isn’t – both on the pages of the journal itself. Some titles in red maybe available in full elsewhere, such as in university archives. The names of laureates in the papers’ citations are underlined.

Physiology/medicine

“for their discoveries of molecular mechanisms controlling the circadian rhythm”

The paywall for papers by Young and Rosbash published in Nature were lifted by the journal on the day their joint Nobel Prize was announced. Until then, they’d been inaccessible to the general public. Interestingly, both papers acknowledge funding grants from the US National Institutes of Health, a tax-funded body of the US government.

Michael Young

Restoration of circadian behavioural rhythms by gene transfer in Drosophila – Nature 312, 752 – 754 (20 December 1984); doi:10.1038/312752a0 link

Isolation of timeless by PER protein interaction: defective interaction between timeless protein and long-period mutant PERL – Gekakis, N., Saez, L., Delahaye-Brown, A.M., Myers, M.P., Sehgal, A., Young, M.W., and Weitz, C.J. (1995). Science 270, 811–815. link

Michael Rosbash

Feedback of the Drosophila period gene product on circadian cycling of its messenger RNA levels – Nature 343, 536 – 540 (08 February 1990); doi:10.1038/343536a0 link

The period gene encodes a predominantly nuclear protein in adult Drosophila – Liu, X., Zwiebel, L.J., Hinton, D., Benzer, S., Hall, J.C., and Rosbash, M. (1992). J Neurosci 12, 2735–2744. link

Jeffrey Hall

Molecular analysis of the period locus in Drosophila melanogaster and identification of a transcript involved in biological rhythms – Reddy, P., Zehring, W.A., Wheeler, D.A., Pirrotta, V., Hadfield, C., Hall, J.C., and Rosbash, M. (1984). Cell 38, 701–710. link

P-element transformation with period locus DNA restores rhythmicity to mutant, arrhythmic Drosophila melanogaster – Zehring, W.A., Wheeler, D.A., Reddy, P., Konopka, R.J., Kyriacou, C.P., Rosbash, M., and Hall, J.C. (1984). Cell 39, 369–376. link

Antibodies to the period gene product of Drosophila reveal diverse tissue distribution and rhythmic changes in the visual system – Siwicki, K.K., Eastman, C., Petersen, G., Rosbash, M., and Hall, J.C. (1988). Neuron 1, 141–150. link

Physics

“for decisive contributions to the LIGO detector and the observation of gravitational waves”

While results from the LIGO detector were published in peer-reviewed journals, the development of the detector itself was supported by personnel and grants from MIT and Caltech. As a result, the Nobel laureates’ more important contributions were published as a reports since archived by the LIGO collaboration and made available in the public domain.

Rainer Weiss

Quarterly progress reportR. Weiss, MIT Research Lab of Electronics 105, 54 (1972) link

The Blue BookR. Weiss, P.R. Saulson, P. Linsay and S. Whitcomb link

Chemistry

“for developing cryo-electron microscopy for the high-resolution structure determination of biomolecules in solution”

The journal Cell, in which the chemistry laureates appear to have published many papers, publicised a collection after the Nobel Prize was announced. Most papers in the collection are marked ‘Open Archive’ and are readable in full. However, the papers cited by the Nobel Committee in its scientific background document don’t appear there. I also don’t know whether the papers in the collection available in full were always available in full.

Jacques Dubochet

Cryo-electron microscopy of vitrified specimens – Dubochet, J., Adrian, M., Chang, J.-J., Homo, J.-C., Lepault, J., McDowall, A. W., and Schultz, P. (1988). Q. Rev. Biophys. 21, 129-228 link

Vitrification of pure water for electron microscopyDubochet, J., and McDowall, A. W. (1981). J. Microsc. 124, 3-4 link

Cryo-electron microscopy of viruses – Adrian, M., Dubochet, J., Lepault, J., and McDowall, A. W. (1984). Nature 308, 32-36 link

Joachim Frank

Averaging of low exposure electron micrographs of non-periodic objectsFrank, J. (1975). Ultramicroscopy 1, 159-162 link

Three-dimensional reconstruction from a single-exposure, random conical tilt series applied to the 50S ribosomal subunit of Escherichia coli – Radermacher, M., Wagenknecht, T., Verschoor, A., and Frank, J. (1987). J. Microsc. 146, 113-136 link

SPIDER-A modular software system for electron image processingFrank, J., Shimkin, B., and Dowse, H. (1981). Ultramicroscopy 6, 343-357 link

Richard Henderson

Model for the structure of bacteriorhodopsin based on high-resolution electron cryo-microscopyHenderson, R., Baldwin, J. M., Ceska, T. A., Zemlin, F., Beckmann, E., and Downing, K. H. (1990). J. Mol. Biol. 213, 899-929 link

The potential and limitations of neutrons, electrons and X-rays for atomic resolution microscopy of unstained biological moleculesHenderson, R. (1995). Q. Rev. Biophys. 28, 171-193 link (available in full here)

§

By locking the red-tagged papers behind a paywall – often impossible to breach because of the fees involved – they’re kept out of hands of less-well-funded institutions and libraries, particularly researchers in countries whose currencies have lower purchasing power. More about this here and here. But the more detestable thing with the papers listed above is that the latest of them (among the reds) was published in 1995, fully 22 years ago, and the earliest, 42 years go – both on cryo-electron microscopy. Both represent almost unforgivable durations across which to have paywalls, with the journals Nature and Cell further attempting to ride the Nobel wave for attention. It’s not clear if the papers they’ve liberated from behind the paywall will always be available for free hence either.

Read all this in the context of the Nobel Prizes not being awarded to more than three people at a time and maybe you’ll see how much of scientific knowledge is truly out of bounds of most of humankind.

Featured image credit: Pexels/pixabay.

The Indian Science Congress has gutted its own award by giving it to Appa Rao Podile

Featured image credit: ratha/Flickr, CC BY 2.0.

I hadn’t heard of the Millennium Plaque of Honour before yesterday, January 3. From what I was able to read up before filing my report in The Wire (about embattled Hyderabad University vice-chancellor Appa Rao Podile receiving the plaque at the ongoing Indian Science Congress):

  1. It has been awarded by the Indian Science Congress Association since 2003, when it was instituted as the ‘Science & Society Award’
  2. Its name was changed to the New Millennium Plaque of Honour in 2005
  3. It carries a citation, a literal plaque and a cash component of Rs 20,000 “to cover incidental expenses”
  4. It is awarded to two eminent scientists at the Science Congress every year

If the annual event was considered prestigious or even very laudable until 2014, I’m not entirely sure (although it certainly wasn’t a very gala affair). But in 2015 and after, it’s certainly taken a beating. In 2015, particularly, the congress was invaded by right-wing nuts convinced that Vedic age scholars had flown planes to Mars and transplanted animal heads onto human bodies. Proceedings were relatively free of controversy in 2016 before taking another turn for the worse in 2017: by giving a Millennium Plaque to Appa Rao (as well as to Avula Damodaram, but that’s a lesser problem we’ll come to later).

A day or so ago, in a conversation on Twitter, both R. Prasad (The Hindu‘s science editor) and Gautam Desiraju (a celebrated chemist at the Indian Institute of Science, Bengaluru) agreed that many Indian events had off late been banking on legitimacy ‘loaned’ from foreign institutions. For example, a large part of the Indian Science Congress’s public outreach every year involves blaring that X Nobel laureates will be in attendance. Nobel laureates are eminent people men, sure, but often they don’t do much other than give a talk and just be in attendance. And their presence doesn’t do much for the quality of the conference, overall in a decline, either (see footnote). In December 2012, P. Balaram, the director of IIT-Madras, wrote in an editorial in the journal Current Science,

… few practising scientists of note consider the Congress as an important event. Pomp and ceremony take precedence over substance. Over the years the Congress has been reduced to an occasion where the inaugural session appears to be the raison de etre for the meeting. The traditional opening address by the Prime Minister predictably reiterates governmental commitment to support science and invariably promises to remove the many bureaucratic hurdles that sometimes loom larger than life in the minds of many scientists. The presence of the executive head of government invests the inaugural event with an importance that is often not commensurate with the quality of the scientific sessions that follow. The occasion is also used to showcase a couple of Nobel laureates, who fly in to speak to audiences with little appetite for excessively technical talks. The organisers, bolstered by considerable government backing, are always good hosts; the distinguished foreign presence ensuring that the Congress always acquires a degree of respectability rarely supported by the scientific program.

In such times, the value of reinforcing local rewards, recognitions, symbols, ideals, etc. is as important as respecting and re-legitimising them as well. This means that an award like the Millennium Plaque of Honour (despite its pompous name), instituted as it has been by the Indian Science Congress, should be given on every occasion to scientists truly deserving of the award and, more importantly, never to anyone who will lower by association the prestige accorded to the award.

Appa Rao is capable of doing the latter. Particularly after Rohith Vemula’s suicide last year (and more generally for a half-year period before that), Appa Rao, as vice-chancellor, was responsible for allowing partisan interventions from the Bharatiya Janata Party (BJP) to interfere in university student politics as well as for violently quelling student protests that followed on the University of Hyderabad campus. Shortly after the news of Vemula’s death broke, the Times of India also reported that Appa Rao had acquired his vice-chancellorship through political connections, especially with BJP minister Venkaiah Naidu and Telugu Desam Party chief Chandrababu Naidu.

A relevant passage from our coverage of the incidents:

Police, CRPF and RAF forces came to the campus, and students assembled on the lawns outside the VC’s lodge were brutally removed and lathi charged. Some students were badly injured and had to be taken to hospitals, sources have said. Students have also said that they were abused and insulted, and female students were threatened with rape. Students from minority communities were allegedly called “terrorists”.

It’s impossible to overlook the fact that his only presence in recent memory was as a craven but powerful stooge, and in fact almost never for his work as a scientist. He hasn’t done anything memorable of late nor as he displayed the integrity due a vice-chancellor of a public institution. In fact, shortly after the student protests, I had also published evidence of plagiarism in three of his research papers. If he has won a Millennium Plaque, then it only means the ‘honour’ doesn’t stand for research excellence anymore as much as for neglecting one’s duties and for perverting the all-important autonomy of an important position.

Worse yet, it seems an award of the Indian Science Congress has become subverted into becoming an instrument of negotiation for political agents: “You let me interfere in your duties, I will give you a fancy-sounding award”. The other recipient of the same award this year, Avula Damodaram, doesn’t inspire confidence, either – although I concede I have no evidence following my suspicions (yet). Damodaram is the vice-chancellor of Sri Venkateswara University, Tirupati, the same institute that’s hosting the science congress this year. Binay Panda, a bioinformatician and friend, wasn’t surprised:

§

Footnote: Mukund Thattai, a biologist at the National Centre for Biological Sciences, Bengaluru, conducted a poll on Twitter asking why people took to science. The option ‘once saw a Nobel laureate’ clocked in last:

Seventy-four is not a great sample size but 1% is a far more abysmal number.

Awards week

I went into this year’s Nobel Prize Announcements Week a little confused about why I was excited. For me the prizes have always highlighted the recipients’ work, and that’s likelier than not a field of study I’ve probably never heard of (with the exceptions being physics – though I don’t presume I’m familiar with all of it – and, occasionally, literature), but then I’m also forced to think about whether the institution of the prizes isn’t becoming outmoded. It probably is; in fact, with physics I can say more forcefully that many of its rules already are out of another era.

But before I could write the obligatory criticism, an amazing article by Roberta Sinatra et al appeared in Nature Physics, titled A century of physics. Using Web of Science data, it discusses not just how and why the breadth of physics literature has increased over the years but also the motivations of the various sub-fields that have emerged under physics – especially concerning the growing need for multidisciplinarity, a topic that the Nobel Prizes for physics aren’t equipped to acknowledge. Check the piece out if you’ve the time, it’s deliciously detailed.

Anyway, as the announcements started to roll in, it was simply fortunate that the first two (for medicine/physiology and physics) afforded critical perspectives on India – allowing me to substitute the “Are the Nobels important” question with the “Is this how we screwed up” question. You could argue that this is in fact a subtle acknowledgement of the Nobel Prizes’ importance – it is but only insofar as I can say “Here’s what not winning a Nobel tells us about how we’re screwing up in xyz situations”. To wit: With the medicine prize, I used the example of Youyou Tu’s finding artemisinin with the guidance of an ancient Chinese text to look at how India’s popularising its ancient knowledge the wrong way. An excerpt:

And here emerges an instructive lesson about what Tu did differently – to not just extract artemisinin but also to preserve the dignity as well as intellectual context of Ge Hong’s work in which she found her answer. After she extracted an effective form of artemisinin in 1972, Tu arranged for its structure to be studied at the Chinese Academy of Sciences in 1975, performed clinical trials in accordance with the best practices of the field by 1977, published her research (though not in English until the 1980s due to the prevailing political environment), and finally participated in the study of large-scale production mechanisms.

What was demonstrated at the ISC in January, on the other hand, belies a lazier attempt at translating old knowledge into newer contexts. The current government’s support for phylotherapy allows researchers to forward non-peer-reviewed results in obscure, self-published journals that do nothing to advance its contents’ credibility when a better alternative would have been to organise and digitise the literature, make it more accessible, and support credible institutions in exploring the knowledge – blend the ancient with the modern, so to speak.

The physics prize was easier to connect to India: it went for the discovery of neutrino oscillations, to study which India is supposed to be building a neutrino observatory but isn’t thanks to political impediments (though not entirely environmental impediments). Again, an excerpt:

Building on similarly advanced principles of detection, India and China are also constructing neutrino detectors.

At least, India is supposed to be. China on the other hand has been labouring away for about a year now in building the Jiangmen Underground Neutrino Observatory (JUNO). India’s efforts with the India-based Neutrino Observatory (INO) in Theni, Tamil Nadu have, on the other hand, ground to a halt. The working principles behind both INO and JUNO are targeted at answering the mass-ordering questions. And if answered, it would almost definitely warrant a Nobel Prize in the future.

INO’s construction has been delayed because of a combination of festering reasons with no end in sight. The observatory’s detector is a 50,000-ton instrument called the iron calorimeter that is to be buried underneath a kilometre of rock so as to filter all particles but neutrinos out. To acquire such a natural shield, the principal institutions involved in its construction – the Department of Atomic Energy (DAE) and the Institute of Mathematical Sciences, Chennai (Matscience) – have planned to hollow out a hill and situate the INO in the resulting ‘cave’. But despite clearances acquired from various pollution control boards as well as from the people living in the area, the collaboration has faced repeated resistance from environmental activists as well as politicians who, members of the collaboration allege, are only involved for securing political mileage.

I like to imagine that such analytical comparisons are a curious, twisted reflection of a larger trend playing out in my glorious country. While the way we’re doing some of our science and pseudoscience is actively repelling international recognition, many winners of the prestigious Sahitya Akademi award, conferred for literary excellence, are returning their trophies decrying Prime Minister Narendra Modi’s silence over the Dadri lynching incident as well as the religiously motivated persecution and murder of rationalists that Nayantara Sahgal, who kicked off the returnings, called a “reign of terror”.

Circling back: The chemistry prize, however, I couldn’t make much sense of. My friend Akshat Rathi was quicker: for example, he told me how the prize, for “mechanistic studies of DNA repairs”, had overlooked this year’s Lasker Award winners (traditionally, these awardees are likelier to be Nobel Laureates). And finally, the literature prize – announced today – was a brilliant stroke of luck simply because it was awarded to Svetlana Alexievich, two of whose books I’ve actually read (one of which I highly recommend: Voices from Chernobyl). I wrote about her here.

Incidentally, The Wire also had a couple pieces concerning the Nobel Prize before the announcements rolled in: one to talk about the CRISPR/Cas9 tool for gene editing by Nandita Jayaraj and another, by me, that discussed plausible reasons why three particular Indians were passed up for the prize (M.K. Gandhi, Meghnad Saha and Satyen Bose).

The Nobel intent

A depiction of Alfred Nobel in the Nobel Museum in Stockholm. Credit: sol_invictus/Flickr, CC BY 2.0
A depiction of Alfred Nobel in the Nobel Museum in Stockholm. Credit: sol_invictus/Flickr, CC BY 2.0

About three weeks from now, the Nobel Foundation will announce the winners of the 2015 Nobel Prizes. Every year, commentators, opinionators and enthusiasts try to guess who will win the awards – some of them have become famous because they’ve been able to guess the winners with uncanny accuracy. However, as it happens, the prizewinners’ profiles have sometimes exposed patterns which tell us how they might have been selected over others. For example, winners of the physics prize have also typically been awarded the Wolf Prize. For another, like a recent study showed, winners of the medicine and physiology prizes seem to have had similar qualitative preferences for their inter-institutional collaborations.

More light is likely to be shed on its opaque selection process by the Nobel Foundation’s decision to open up its archives and reveal the name of not just all nominees but also the nominators who got those names on the rosters each year.  The complete list for all prizes – except economics – awarded between 1901 and 1964 is now available for the first time. The lists for awards given after 1965 are not visible because they’re sealed for 50 years. With the information, the question of “Who nominated whom?” is worth asking not just for trivia’s sake but also because it throws up clues about the politics behind decisions, the kinds of names that were ignored for the prizes, why they were ignored, and how the underpinning rationale has changed through various social periods.

There are three famous examples with which to illustrate these issues.

Mohandas Gandhi

The first is of M.K. Gandhi. The Nobel Committee admitted in 2001 that overlooking Gandhi had been one of its most infamous mistakes. In 1937, in a total of 63 nominations by prominent people, Gandhi received his first: from Ole Colbjørnsen, a Norwegian politician. Colbjørnsen would nominate Gandhi in 1938 and 1939 as well. After that, the name of Gandhi among the nominees reappears in 1947, put there by G.B. Pant, B.G. Kher and Mavalankar, and in 1948, this time with the endorsement of Frede Castberg (a Norwegian jurist), six professors of the University of Bordeaux, five from Columbia University, the American Friends Service Committee, Christian Oftedal (a Norwegian politician) and the American economist Emily Greene Balch. Gandhi was assassinated in January 1948, and since the Foundation doesn’t allow posthumous awards, his ‘case’ ended that year.

The winners in the years he was nominated in were

  • 1937 – Robert Cecil
  • 1938 – Nansen International Office for Refugees
  • 1939 – No winner
  • 1947 – AFSC and Friends Service Council
  • 1948 – No winner

The committee declined to award the prize in 1948 because “there was no suitable living candidate”. This was with reference to Gandhi, who may have received the prize had he not been killed that year. There have also been some discussions on whether the committee could have made an exception for Gandhi and awarded it posthumously, especially since the nominations had arrived a few days before his death and because his death was quite unexpected, too (incidentally, posthumous awards of the Physics Prize were allowed until 1974 if the awardee was alive at the time of nomination). On the other hand, even if these arguments had been taken seriously, they wouldn’t have fetched the Peace Prize for Gandhi – why he wasn’t chosen alludes to a different issue.

The nomination process is essentially one of filtering, and though it differs for each prize, they are all variations of the following: some 3,000 individuals around the world are asked to send in their preliminary nominations, out of which the Nobel Committee filters out and passes on an order of magnitude fewer names to relevant institutions. Finally, the institutions, represented by members on the committee, vote on the day of the prize, with the result being announced immediately after the counting. The person/persons/institutions with the most votes wins the prizes. There is a distinct committee for each of the prizes.

The number of nominators increases every year – to also include the previous year’s winners, for one – so the names of the first winners were essentially sourced from a handful of individuals.

In 1999, Øyvind Tønnesson, then nobelprize.org’s Peace Director, wrote that in Gandhi’s time, the members of the committee weren’t in favour of him for two reasons. First, many of them couldn’t help but blame Gandhi for some of the incidents of violence in India during his supposedly peaceful resistance, going as far as to claim he should’ve known that his actions would precipitate violence – for example, and especially, the Chauri Chaura incident in 1922. Second, as Tønnesson wrote, the members preferred awardees “who could serve as moral and religious symbols in a world threatened by social and ideological conflicts”, and on that note were opposed to the political implications of Gandhi’s movement – especially his role in effectuating the Partition as well as an inability to quell the widespread violence that followed.

Oddly enough, the Nobel Peace Prize is essentially a political prize, and its credibility often can’t be dissociated from the clout of members of the voting committee. In fact, alongside the Literature Prize, the duo has often been the subject of controversy simply by illustrating the linguistic and cultural differences between the Scandinavian electors and their multitudes of candidates. In 1965, U. Thant, then the Secretary-General of the United Nations, was not given the award because the chair of the Nobel Committee then, Gunnar Jahn, was opposed to him despite a majority having favoured Thant for defusing the Cuban missile crisis. One plausible reason that has been advanced, based on Jahn’s track record when he was the chair, was that Thant was only doing his duty and that none of his initiatives to secure peace in the world stepped beyond that ambit – contrary to the actions of the recipients of the 1947 Peace Prize, in Jahn’s opinion. Another incident betrayed how Jahn’s influence was inordinate, too, despite all assurances toward the selection process being democratic: he threatened to resign if Linus Pauling wasn’t awarded the Peace Prize in 1963 while the majority had voted against the chemist.

Another contention has centred on the measures of worthiness. Why can’t the Nobel Prize be awarded to more than three people at a time? Why is the time-difference between the award-winning work being done and the award being given so huge? And on what grounds will each prospective laureate be judged precisely? In the case of the 2013 Nobel Prize for physics, Peter Higgs and Francois Englert were named the recipients for work done 49 years ago, in 1964, even as four others who’d done the same work in that year were ignored. Jorge Luis Borges has been repeatedly overlooked for the Literature Prize with rumours abounding that the committee was not supportive of his conservative political views and because he’d received a prize from Chilean dictator Augusto Pinochet. On the other hand, some of the greatest writers in history have been politically motivated to produce their best works, so in not specifying the bases on which candidates can be rejected, the Nobel Committee makes the Literature Prize an exercise in winning the approval of a group of Scandinavians who may or may not have a sound knowledge of non-European politics.

Meghnad Saha

Meghnad Saha was an astrophysicist known for an eponymous equation that allowed astronomers to determine how much various elements had been ionised in a star based on its temperature. Saha first published his results in 1920, which were built upon by Irving Langmuir in 1923. Ever since, the equation has also been known as the Saha-Langmuir equation. Presumably for this work, Saha was nominated for the Physics Prize by Dehendra Bose and Sisir Mitra in 1930, by Arthur Compton in 1937, by Mitra again in 1939, by Compton again in 1940, and by Mitra again in 1951* and 1955. On February 16, 1956, Saha passed away.

While his equation has become applicable in different high-energy physics contexts, at the time of its conception it was advertised as being for astrophysics. And in that context, however, a shortcoming was spotted among Saha’s assumptions by Ralph Fowler and Edward Arthur Milne in 1923, who then improved the equation to fix the consequences of that shortcoming. Even so, there appeared to have been some misconceptions in the wider astrophysics community, especially in Europe, about who was the originator – not of the equation but of the more important underlying theory, which Saha called the theory of selective radiation pressure. In 1917, he was financially strained and was faced with a disappointing prospect: that the paper he’d send to the Astrophysical Journal detailing the theory couldn’t be printed unless he bore some of the printing costs, which was out of the question. So he had the paper published in the Journal of the Department of Science at Calcutta University instead, “which had no circulation worth mentioning”.

To quote from the Vigyan Prasar archives, which in turn quotes from Saha himself,

“… I might claim to be the originator of the Theory of Selective Radiation Pressure, though on account of discouraging circumstances, I did not pursue the idea to develop it. E.A. Milne apparently read a note of mine in Nature 107, 489 (1921) because in his first paper on the subject ‘Astrophysical Determination of Average of an Excited Calcium Atom’, in Month. Not. R. Ast. Soc., Vol.84, he mentioned my contribution in a footnote, though nobody appears to have noticed. His exact words are: ‘These paragraphs develop ideas originally put forward by Saha’.”

Later in the same article, now quoting one of Saha’s students, Daulat Kothari:

It is pertinent to remark that the ionisation theory was formulated by Saha working by himself in Calcutta, and the paper quoted above was communicated by him from Calcutta to the Philosophical Magazine – incorrect statements to the contrary have sometimes been made. Further papers soon followed. It is not too much to say that the theory of thermal ionisation introduced a new epoch in astrophysics by providing for the first time, on the basis of simple thermodynamic consideration and elementary concepts of the quantum theory, a straight forward interpretation of the different classes of stellar spectra in terms of the physical condition prevailing in the stellar atmospheres.

Had Saha’s work appeared in the Astrophysical Journal in 1917, would his fortunes have been different?

And given that the publishing volume has been growing very fast of late, do the prizes remain representative of the research being conducted? This question may be suppressed by arguing that the prizes are awarded to remarkable research, of the kind that is so momentous that it can’t but see the light of day. At the same time, as in Saha’s case, how much research passes under the radar of the Foundation even if it’s most in need of the kind of visibility the award can bring? And perhaps this is the more important question: of the dozens of nominations the Foundation has received every year for the Nobel Prizes, how many lost out because they published their work in the so-called low impact-factor (i.e. low-visibility) journals?

Satyendra Nath Bose

A third example is of Satyendra Nath Bose. Despite seminal work done in the 1920s, including on a topic that was quickly recognised as being radical and employed by multiple Nobel-Prize-winning scientists later, Bose was never awarded the Physics Prize. Perhaps his greatest honour for performing that work, apart from contributing to the science itself, was the British physicist Paul A.M. Dirac naming a significant class of fundamental particles after him (bosons). When Higgs and Englert were awarded the Physics Prize in 2013 for having conceived the theory behind the Higgs boson in 1964, a cry went up around India calling for Bose to recognised for his work and be awarded a share of the prize that year. The demand was thoroughly misguided because the Bose-Einstein statistics describe all bosons whereas the Higgs Six had focused on one peculiar boson. If anything, Bose could have been awarded the prize separately: he was nominated by Kedareswar Banerji in 1956, by Daulat Kothari in 1959 and by S. Bagchi in 1962.

In contrast, the only other Indian to have won the Physics Prize (before 1964), C.V. Raman, was nominated by no less than 10 people, including Ernest Rutherford, Louis-Victor de Broglie, Johannes Stark and Niels Bohr – all then or future laureates – in the same year. A case of “who nominated whom”, then? Not quite. Another reported flaw of the Physics Prize has been that it has favoured discoveries over inventions, with the 2014 edition being the most recent of a handful of exceptions to that rule. And among those discoveries, the prize’s selectors have consistently preferred experimental proof. That would explain the unseemly gap between Higgs’s and Englert’s papers in 1964 and their awards in 2013 – and it would also explain why Bose never won the prize himself. Bose’s work in statistics helped understand an already observed anomaly but it provided no other new predictions against which his theory could be tested. In 1924, Einstein would make that prediction: of a unique state of matter since called the Bose-Einstein condensate (BEC). The BEC was first experimentally observed in 1995, fetching three physicists the 2001 Physics Prize. That the statistics would also explain the superfluidity of liquid helium-4 was first suggested by Fritz London in 1938 and proved by Lev Landau in 1941 (so winning the 1962 Physics Prize).

However, this is not a defence of Bose not winning the prize as much as a cautionary note: the helpful thing to remember would be that though the Nobel Prizes may rank among the most prestigious distinctions, they have a character of their own, and that human enterprise cannot be divided as Nobel-class and non-Nobel-class, as if it were an aircraft carrier. For in the more than 800 laureates the Nobel Foundation has counted since 1901, the omissions stand out as much as the rest: apart from the few already mentioned, Chinua Achebe, Jocelyn Bell Burnell, Rosalind Franklin, Václav Havel, Lise Meitner, J.R.R. Tolkien and John Updike come to mind. In Bell Burnell’s case, in fact, another man receiving the Physics Prize for a discovery she made only highlights another failure of the Nobel Foundation and has since become an example often invoked to highlight the plight of women in science.

*Also in 1951, Saha nominated Arnold Sommerfeld, a German physicist infamous for being overlooked for a Nobel Prize despite having received more than 80 nominations over many years.

The Wire
September 15, 2015