Can gravitational waves be waylaid by gravity?

Yesterday, I learnt the answer is ‘yes’. Gravitational waves can be gravitationally lensed. It seems obvious once you think about it, but not something that strikes you (assuming you’re not a physicist) right away.

When physicists solve problems relating to the spacetime continuum, they imagine it as a four-dimensional manifold: three of space and one of time. Objects exist in the bulk of this manifold and visualisations like the one below are what two-dimensional slices of the continuum look like. This unified picture of space and time was a significant advancement in the history of physics.

While Hendrik Lorentz and Hermann Minkowski first noticed this feature in the early 20th century, they did so only to rationalise empirical data. Albert Einstein was the first physicist to fully figure out the why of it, through his theories of relativity.

A common way to visualise the curvature of spacetime around a massive object, in this case Earth. Credit: NASA

Specifically, according to the general theory, massive objects bend the spacetime continuum around themselves. Because light passes through the continuum, its path bends along the continuum when passing near massive bodies. Seen head-on, a massive object – like a black hole – appears to encircle a light-source in its background in a ring of light. This is because the black hole’s mass has caused spacetime to curve around the black hole, creating a cosmic mirage of the light emitted by the object in its background (see video below) as seen by the observer. By focusing light flowing in different directions around it towards one point, the black hole has effectively behaved like a lens.

So much is true of light, which is a form of electromagnetic radiation. And just the way electrically charged particles emit such radiation when they accelerate, massive particles emit gravitational waves when they accelerate. These gravitational waves are said to carry gravitational energy.

Gravitational energy is effectively the potential energy of a body due to its mass. Put another way, a more massive object would pull a smaller body in its vicinity towards itself faster than a less massive object would. The difference between these abilities is quantified as a difference between the objects’ gravitational energies.

Credit: ALMA (NRAO/ESO/NAOJ)/Luis Calçada (ESO)

Such energy is released through the spacetime continuum when the mass of a massive object changes. For example, when two binary black holes combine to form a larger one, the larger one usually has less mass than the masses of the two lighter ones together. The difference arises because some of the mass has been converted into gravitational energy. In another example, when a massive object accelerates, it distorts its gravitational field; these distortions propagate outwards through the continuum as gravitational energy.

Scientists and engineers have constructed instruments on Earth to detect gravitational energy in the form of gravitational waves. When an object releases gravitational energy into the spacetime continuum, the energy ripples through the continuum the way a stone dropped in water instigates ripples on the surface. And just the way the ripples alternatively stretch and compress the water, gravitational waves alternatively stretch and compress the continuum as they move through it (at the speed of light).

Instruments like the twin Laser Interferometer Gravitational-wave Observatories (LIGO) are designed to pick up on these passing distortions while blocking out all others. That is, when LIGO records a distortion passing through the parts of the continuum where its detectors are located, scientists will know it has just detected a gravitational wave. Because the frequency of a wave is directly proportional to its energy, scientists can use the properties of the gravitational wave as measured by LIGO to deduce the properties of its original source.

(As you might have guessed, even a cat running across the room emits gravitational waves. However, the frequency of these waves is so very low that it is almost impossible to build instruments to measure them, nor are we likely to find such an exercise useful.)

I learnt today that it is also possible for instruments like LIGO to be able to detect the gravitational lensing of gravitational waves. When an object like a black hole warps the spacetime continuum around it, it lenses light – and it is easy to see how it would lens gravitational waves as well. The lensing effect is the result not of the black hole’s ‘direct’ interaction with light as much as its distortion of the continuum. Ergo, anything that traverses the continuum, including gravitational waves, is bound to be lensed by the black hole.

The human body evolved eyes to receive information encoded in visible light, so we can directly see lensed visible-light. However, we don’t possess any organs that would allow us to do the same thing with gravitational waves. Instead, we will need to use existing instruments, like LIGO, to detect these particular distortions. How do we do that?

When two black holes are rapidly revolving around each other, getting closer and closer, they shed more and more of their potential energy as gravitational waves. In effect, the frequency of these waves is quickly increasing together with their amplitude, and LIGO registers this as a chirp (see video below). Once the two black holes have merged, both frequency and amplitude drop to zero (because a solitary spinning black hole does not emit gravitational waves).

In the event of a lensing, however, LIGO will effectively detect two sets of gravitational waves. One set will arrive at LIGO straight from the source. The second set – originally sent off in a different direction – will become lensed towards LIGO. And because the lensed wave will effectively have travelled a longer distance, it will arrive a short while after the direct wave.

The distance scale here is grossly exaggerated for effect

However, LIGO will not register two chirps; in fact, it will register no chirps at all. Instead, the direct wave and the lensed wave will interfere with each other inside the instrument to produce a characteristically mixed signal. By the laws of wave mechanics, this signal will have increasing frequency, as in the chirp, but uneven amplitude. If it were sonified, the signal’s sound would climb in pitch but have irregular volume.

A statistical analysis published in early 2018 (in a preprint paper) claimed that LIGO should be able to detect gravitationally lensed gravitational waves at the rate of about once per year (and the proposed Einstein Telescope, at about 80 per year!). A peer-reviewed paper published in January 2019 suggested that LIGO’s design specs allow it to detect lensing effects due to a black hole weighing 10-100,000-times as much as the Sun.

Just like ‘direct’ gravitational waves give away some information about their sources, lensed gravitational waves should also give something away about the objects that deflected them. So if we become able to use LIGO, and/or other gravitational wave detectors of the future, to detect gravitationally lensed gravitational waves, we will have the potential to learn even more about the universe’s inhabitants than gravitational-wave astronomy currently allows us to.

Thanks to inputs from Madhusudhan Raman, @ntavish, @alsogoesbyV and @vaa3.

A shorter article about short gamma ray bursts lights up little

  • Identify a simple and well-defined question
  • Describe the question and answer it
  • Get the fuck out

Writing with these three rules in mind makes for a good science article. You stick to the point, you know what details to include and what to leave out and, most importantly, you set straightforward expectations and meet them. The overall effect is for the reader to walk away feeling not disappointed. That’s always a happy ending.

Sadly, not everyone writes like this – rather, more broadly, not all news publishers think of science articles this way. For example, The Hindu regularly publishes science articles so packed with information – about the study as much as its authors – that you’re left confused about what you just read. Was it a profile or was it an explainer? It doesn’t matter because it failed either way.

The latest example of this kind of writing is an article about short gamma ray bursts. The binary neutron star merger known by the gravitational-wave event designation GW170817 was expected by astrophysicists to have unleashed a short gamma ray burst at the moment of collision – but data obtained of the event shows no signs of the expected radio signature. A group of scientists led by Kunal Mooley from Oxford University suggested this could be because GW170817 released a new kind of gamma ray burst.

BusinessLine (a business newspaper with the same publisher and top management as The Hindu) carried an article attempting to discuss all this. Sample the opening para, a mulch of facts and inaccuracies:

Screen Shot 2017-12-23 at 09.18.40

“The one located on the outskirts of Pune”? Sounds like everyone must know about it even if they don’t. “First-ever detection of gravitation waves”? Not really: gravitation waves, a.k.a. gravity waves and unlike gravitational waves, can be observed in Earth’s atmosphere. Also, the first-ever detection of gravitational waves came last year; what came in August was the first-ever detection of a neutron star merger. The three US scientists won the award for building LIGO, not detecting GW170817.

The rest of the article tries to simultaneously explain Mooley and co.’s interpretation of the data and also provide a glimpse of his educational trajectory. Why would I want to know he studied in Pune and Mumbai? Unless this is because the author wanted to drive home the India connection – which is all the more troubling because it plays up an aspect of the researcher’s identity that is irrelevant to their professional accomplishment. I’ve noticed many publications succumbing to this kind of thinking: if researcher is Indian, cover the paper/study/whatever irrespective of the legitimacy, strength and/or novelty of what they’re saying.

The science ought to take precedent, not the researcher’s identity. But when it doesn’t, you typically end up writing something that’s definitely not news and likely trash. You end up wrapping your national pride around a core of stupidity. I recommend the pages of ScoopwhoopThe QuintThe Better IndiaDailyOThe Times of India and The New Indian Express, among others, for examples. It’s also possible that the author was conscious about providing an India connection so readers in India took the article more seriously. I’ve made noise about such behaviour many times before, such as here: science shouldn’t be assessed, or enjoyed, solely according to what it can do for humankind.

Finally, it’s possible that the newspaper itself wanted to establish all details on record for posterity – but AFAIK, the BusinessLine is not a newspaper of record. This of course is a minor point.

By ditching the extraneous details, the author and the editor could’ve had the space to focus on the science more, using better language and without the painful economy of words it’s currently striving to. They could even have devoted some words to discussing whether other astronomers have disputed Mooley’s interpretation (they have), an exercise that would’ve made the article more reliable than it is. And to those who’re saying the article was probably kept short because there might not have been space in the newspaper, I’ve a bigger complaint: why wasn’t a short version published in print and a longer version online?

In all, I don’t think BusinessLine is taking its science journalism seriously. The time is past when they could’ve gotten ahead simply by being one of the few publications in the country to write articles about short gamma ray bursts. But given the complacency with which the article seems to have been composed and edited, maybe that time shouldn’t have existed in the first place. It surely doesn’t now.

Featured image: An artist’s illustration of a bright gamma-ray burst. Caption and credit: NASA/Wikimedia Commons.

Onto drafting the gravitational history of the universe

It’s finally happening. As the world turns, as our little lives wear on, gravitational wave detectors quietly eavesdrop on secrets whispered by colliding blackholes and neutron stars in distant reaches of the cosmos, no big deal. It’s going to be just another day.

On November 15, the LIGO scientific collaboration confirmed the detection of the fifth set of gravitational waves, made originally on June 8, 2017, but announced only now. These waves were released by two blackholes of 12 and seven solar masses that collided about a billion lightyears away – a.k.a. about a billion years ago. The combined blackhole weighed 18 solar masses, so one solar mass’s worth of energy had been released in the form of gravitational waves.

The announcement was delayed because the LIGO teams had to work on processing two other, more spectacular detections. One of them involved the VIRGO detector in Italy for the first time; the second was the detection of gravitational waves from colliding neutron stars.

Even though the June 8 is run o’ the mill by now, it is unique because it stands for the blackholes of lowest mass eavesdropped on thus far by the twin LIGO detectors.

LIGO’s significance as a scientific experiment lies in the fact that it can detect collisions of blackholes with other blackholes. Because these objects don’t let any kind of radiation escape their prodigious gravitational pulls, their collisions don’t release any electromagnetic energy. As a result, conventional telescopes that work by detecting such radiation are blind to them. LIGO, however, detects gravitational waves emitted by the blackholes as they collide. Whereas electromagnetic radiation moves over the surface of the spacetime continuum and are thus susceptible to being trapped in blackholes, gravitational waves are ripples of the continuum itself and can escape from blackholes.

Processes involving blackholes of a lower mass have been detected by conventional telescopes because these processes typically involve a light blackhole (5-20 solar masses) and a second object that is not a blackhole but instead usually a star. Mass emitted by the star is siphoned into the blackhole, and this movement releases X-rays that can be spotted by space telescopes like NASA Chandra.

So LIGO’s June 8 detection is unique because it signals a collision involving two light blackholes, until now the demesne of conventional astronomy alone. This also means that multi-messenger astronomy can join in on the fun should LIGO detect a collision of a star and a blackhole in the future. Multi-messenger astronomy is astronomy that uses up to four ‘messengers’, or channels of information, to study a single event. These channels are electromagnetic, gravitational, neutrino and cosmic rays.

The masses of stellar remnants are measured in many different ways. This graphic shows the masses for black holes detected through electromagnetic observations (purple); the black holes measured by gravitational-wave observations (blue); neutron stars measured with electromagnetic observations (yellow); and the masses of the neutron stars that merged in an event called GW170817, which were detected in gravitational waves (orange). GW170608 is the lowest mass of the LIGO/Virgo black holes shown in blue. The vertical lines represent the error bars on the measured masses. Credit: LIGO-Virgo/Frank Elavsky/Northwestern
The masses of stellar remnants are measured in many different ways. This graphic shows the masses for black holes detected through electromagnetic observations (purple); the black holes measured by gravitational-wave observations (blue); neutron stars measured with electromagnetic observations (yellow); and the masses of the neutron stars that merged in an event called GW170817, which were detected in gravitational waves (orange). GW170608 is the lowest mass of the LIGO/Virgo black holes shown in blue. The vertical lines represent the error bars on the measured masses. Credit: LIGO-Virgo/Frank Elavsky/Northwestern

The detection also signals that LIGO is sensitive to such low-mass events. The three other sets of gravitational waves LIGO has observed involved black holes of masses ranging from 20-25 solar masses to 60-65 solar masses. The previous record-holder for lowest mass collision was a detection made in December 2015, of two colliding blackholes weighing 14.2 and 7.5 solar masses.

One of the bigger reasons astronomy is fascinating is its ability to reveal so much about a source of radiation trillions of kilometres away using very little information. The same is true of the June 8 detection. According to the LIGO scientific collaboration’s assessment,

When massive stars reach the end of their lives, they lose large amounts of their mass due to stellar winds – flows of gas driven by the pressure of the star’s own radiation. The more ‘heavy’ elements like carbon and nitrogen that a star contains, the more mass it will lose before collapsing to form a black hole. So, the stars which produced GW170608’s [the official designation of the detection] black holes could have contained relatively large amounts of these elements, compared to the stellar progenitors of more massive black holes such as those observed in the GW150914 merger. … The overall amplitude of the signal allows the distance to the black holes to be estimated as 340 megaparsec, or 1.1 billion light years.

The circumstances of the discovery are also interesting. Quoting at length from a LIGO press release:

A month before this detection, LIGO paused its second observation run to open the vacuum systems at both sites and perform maintenance. While researchers at LIGO Livingston, in Louisiana, completed their maintenance and were ready to observe again after about two weeks, LIGO Hanford, in Washington, encountered additional problems that delayed its return to observing.

On the afternoon of June 7 (PDT), LIGO Hanford was finally able to stay online reliably and staff were making final preparations to once again “listen” for incoming gravitational waves. As part of these preparations, the team at Hanford was making routine adjustments to reduce the level of noise in the gravitational-wave data caused by angular motion of the main mirrors. To disentangle how much this angular motion affected the data, scientists shook the mirrors very slightly at specific frequencies. A few minutes into this procedure, GW170608 passed through Hanford’s interferometer, reaching Louisiana about 7 milliseconds later.

LIGO Livingston quickly reported the possible detection, but since Hanford’s detector was being worked on, its automated detection system was not engaged. While the procedure being performed affected LIGO Hanford’s ability to automatically analyse incoming data, it did not prevent LIGO Hanford from detecting gravitational waves. The procedure only affected a narrow frequency range, so LIGO researchers, having learned of the detection in Louisiana, were still able to look for and find the waves in the data after excluding those frequencies.

But what I’m most excited about is the quiet announcement. All of the gravitational wave detection announcements before this were accompanied by an embargo, lots of hype building up, press releases from various groups associated with the data analysis, and of course reporters scrambling under the radar to get their stories ready. There was none of that this time. This time, the LIGO scientific collaboration published their press release with links to the raw data and the preprint paper (submitted to the Astrophysical Journal Letters) on November 15. I found out about it when I stumbled upon a tweet from Sean Carroll.

And this is how it’s going to be, too. In the near future, the detectors – LIGO, VIRGO, etc. – are going to be gathering data in the background of our lives, like just another telescope doing its job. The detections are going to stop being a big deal: we know LIGO works the way it should. Fortunately for it, some of its more spectacular detections (colliding intermediary-mass blackholes and colliding neutron stars) were also made early in its life. What we can all look forward to now is reports of first-order derivatives from LIGO data.

In other words, we can stop focusing on Einstein’s theories of relativity (long overdue) and move on to what multiple gravitational wave detections can tell us about things we still don’t know. We can mine patterns out of the data, chart their variation across space, time and their sources, and begin the arduous task of drafting the gravitational history of the universe.

Featured image credit: Lovesevenforty/pixabay.

On that 'Last Word on Nothing' post

A post published on the Last Word On Nothing blog yesterday has been creating quite the stir on Twitter. Excerpt:

While I can appreciate that this is an important scientific discovery, I still have a hard time mustering excitement over gravitational waves. I would not have read these articles had I not embarked on this experiment. And I wanted to stop reading some of these articles as I was conducting the experiment. Space is not my thing. I don’t think it ever will be, at least not without a concerted effort on my part to get a basic handle on physics and astronomy. …

Physics writers, this is how you nab the physics haters — human emotion. You can explain gravitational waves using the cleanest, clearest, most eloquent words that exist — and you should! — but I want the story of the scientists in all their messy, human glory.

Cassandra Willyard, the post’s author, was writing about the neutron-star collision announcement from LIGO. Many of those who are dissing the point the post is making are saying that Willyard is vilifying the ‘school’ of science writing that focuses on the science itself over its relationship with the human condition. I think she’s only expressing her personal opinion (as the last line in the excerpt suggests) – so the levels of indignation that has erupted in some pockets of the social media over these opinions suggests Willyard may have touched off some nerves.

I myself belong to the school that prefers to excite science readers over the science itself over its human/humanist/humanitarian aspects. In the words of Tracy, who wrote them as a comment on Willyard’s post,

So many amazing things happen in this universe without a human noticing it, reflecting on it, understanding it, being central to it. So many wondrous mysteries abound despite the ego. The human story is just one of billions.

And I will concede from personal experience that it’s quite difficult as a result to sell such stories to one’s editors as well as readers. I’ve written about this many times before, e.g. here; edited excerpt:

I couldn’t give less of a fuck for longer pieces, especially because they’re all the same: they’re concerned with science that is deemed to be worthy of anyone’s attention because it is affecting us directly. And I posit that they’ve kept us from recognising an important problem with science journalism in the country: it is becoming less and less concerned with the science itself; what has been identified as successful science journalism is simply a discussion – no matter how elaborate and/or nuanced – of how science impacts us. Instead, I’d love to read a piece reported over 5,000 words about molecules, experiments, ideas. It should be okay to want to write only about particle physics because that’s all I’m interested in reading. Okay to want to write only about this even if I don’t have any strength to hope that QCD will save lives, that Feynman diagrams will help repeal AFSPA, that the LHC will accelerate India’s economic growth, that the philosophies of fundamental particles will lead to the legalisation of same-sex marriage. I haven’t been presented with any evidence whatsoever to purchase my faith in the possibility that the obscurities of particle physics will help humans in any way other than to enlighten them, that there is neither reward nor sanction in anxiously bookending every articulation of wonder with the hope that we will find a way to profit from all of our beliefs, discoveries and perceptions.

For many people in this ‘school’, this fight is almost personal because it’s arduous and requires tremendous conviction, will and resilience on one’s part to see coverage of such kind through. In this scenario, to have a science writer come forward and say “I won’t write about this science because I don’t understand this science” can be quite dispiriting. It’s a science writer’s job to disentangle some invention, discovery or whatever and then communicate it to those who are interested in knowing more about it. So when Willyard writes in her post that “The day I write about a neutron star collision is the day hell will freeze over” – it’s a public abdication of an important responsibility, and arguably one of the most complicated responsibilities in journalism in the Information Age thanks to its fiercely non-populist nature.

(Such a thing happened recently with Natalie Wolchover as well. Her words – written against topological physics – were more disappointing to come across because Wolchover writes very good physics pieces for Quanta. And while she apologised for the “flippancy” of her tweet shortly after, saying that she’d been in a hurry at 5.45 am, that’s precisely the sort of sentiment that shouldn’t receive wider coverage without the necessary qualifications. So my thanks to Chad Orzel for the thread he published in response.)

However, it must be acknowledged that the suggestion Willyard makes (in the second paragraph of the excerpt) is quite on point. To have to repeatedly pander to the human condition in one way or another when in fact you think the science in and of itself is incredibly cool can become frustrating over time – but this doesn’t mean that a fundamental disconnect between writers like me and the statistically average science reader out there doesn’t exist. If I’m to get her attention, then I’ve found from experience that one must begin with the humans of science and then flow on to the science itself. As Alice Bell recommends here, you start upstream and go downstream. And once you’ve lured them in, you can begin to discuss the science more freely.

(PS: Some areas of Twitter have gone nuts, claiming Willyard shouldn’t be called a science journalist. I’m making no such judgment call. To be clear, I’m only criticising a peer’s words. I still consider Willyard to be a science journalist – though my fingers cry as I type this because it’s so embarrassing to have to spell it out – and possibly a good one at that going by her willingness to introspect.)

Featured image credit: Pexels/pixabay.

Neutron stars

When the hype for the announcement of the previous GW detection was ramping up, I had a feeling LIGO was about to announce the detection of a neutron-star collision. It wasn’t to be – but in my excitement, I’d written a small part of the article. I’m sharing it below. I’d also recommend reading this post: The Secrets of How Planets Form.

Stars die. Sometimes, when that happens, their outer layers explode into space in a supernova. Their inner layers collapse inwards under the gravity of their own weight in a violent rush. If the starstuff can be packed dense enough, the collapse produces a blackhole – a volume of space where the laws of quantum mechanics and relativity break down and the particles of matter are plunged into a monumental identity crisis. However, if the dying star wasn’t heavy enough when it blew up, then the inward rush will create a very, very, very dense object – but not a blackhole: a neutron star.

Neutron stars are the densest objects in the universe that astronomers can observe. The only things we know are denser than them are blackholes.

You’d think observed means ‘saw’, but what is ‘seeing’ but the light – a form of electromagnetic energy – from an event reaching our eyes? We can’t directly ‘see’ blackholes collide because the collision doesn’t release any electromagnetic energy. So astronomers have built a special kind of eyes – called gravitational wave detectors – that can observe ripples of gravitational energy that the collision lets loose.

The Laser Interferometer Gravitational-wave Detector (LIGO) we already know about. Its twin eyes, located in Washington and Louisiana, US, have detected three blackhole-blackhole collisions thus far. Two of the scientists who helped build it are hot favourites to win the Nobel Prize for physics next week. The other set of eyes involved in the last find is Virgo, a detector in Italy.

You’ve been told that blackholes are freaks of nature. Heavy objects bend spacetime around themselves. Blackholes are freaks because they step it up: they fold it. They’re so heavy that when spacetime bends around them, it goes all the way around and becomes a three-dimensional loop. Thus, a blackhole traps one patch of the cosmos around a vanishingly small heart of darkness. Even light, if it comes close enough, becomes trapped in this loop and can never escape. This is why astronomers can’t observe blackholes directly, and use gravitational-wave detectors instead.

But neutron stars they can observe. They’re exactly what their names suggest: a ball of neutrons. And neutrons experience a force of nature called the strong nuclear force, and it can be 100,000 billion billion billion times stronger than gravity. This makes neutron stars extremely dense and altogether incredibly heavy as well. On their surface, a classic can of Coke will weigh 355,000 billion tonnes, a thousand-times heavier than all the humans on Earth combined.

Sometimes, a neutron star is ravaged by a powerful magnetic field. This field focuses charged particles on the neutron star’s surface into a tight beam of radiation shooting off into space. If the orb is also spinning, then this beam of radiation sweeps through space like the light from a lighthouse sweeps over the sea near it. Such neutron stars are called pulsars.

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 nomenclature of uncertainty

The headline of a Nature article published on December 9 reads ‘LIGO black hole echoes hint at general relativity breakdown’. The article is about the prediction of three scientists that, should LIGO find ‘echoes’ of gravitational waves coming from blackhole-mergers, then it could be a sign of quantum-gravity forces at play.

It’s an exciting development because it presents a simple and currently accessible way of probing the universe for signs of phenomena that show a way to unite quantum physics and general relativity – phenomena that have been traditionally understood to be outside the reach of human experiments until LIGO.

The details of the pre-print paper the three scientists uploaded on arXiv were covered by a number of outlets, including The Wire. And The Wire‘s and Forbes‘s headlines were both questions: ‘Has LIGO already discovered evidence for quantum gravity?’ and ‘Has LIGO actually proved Einstein wrong – and found signs of quantum gravity?’, respectively. Other headlines include:

  • Gravitational wave echoes might have just caused Einstein’s general theory of relativity to break down – IB Times
  • A new discovery is challenging Einstein’s theory of relativity – Futurism
  • Echoes in gravitational waves hint at a breakdown of Einstein’s general relativity – Science Alert
  • Einstein’s theory of relativity is 100 years old, but may not last – Inverse

The headlines are relevant because: Though the body of a piece has the space to craft what nuance it needs to present the peg, the headline must cut to it as quickly and crisply as possible – while also catching the eye of a potential reader on the social media, an arena where all readers are being inundated with headlines vying for attention.

For example, with the quantum gravity pre-print paper, the headline has two specific responsibilities:

  1. To be cognisant of the fact that scientists have found gravitational-wave echoes in LIGO data at the 2.9-sigma level of statistical significance. Note that 2.9 sigma is evidently short of the threshold at which some data counts as scientific evidence (and well short of that at which it counts as scientific fact – at least in high-energy physics). Nonetheless, it still presents a 1-in-270 chance of, as I’ve become fond of saying, an exciting thesis.
  2. To make reading the article (which follows from the headline) seem like it might be time well spent. This isn’t exactly the same as catching a reader’s attention; instead, it comprises catching one’s attention and subsequently holding and justifying it continuously. In other words, the headline shouldn’t mislead, misguide or misinform, as well as remain constantly faithful to the excitement it harbours.

Now, the thing about covering scientific developments from around the world and then comparing one’s coverage to those from Europe or the USA is that, for publications in those countries, what an Indian writer might see as an international development is in fact a domestic development. So Nature, Scientific American, Forbes, Futurism, etc. are effectively touting local accomplishments that are immediately relevant to their readers. The Wire, on the other hand, has to bank on the ‘universal’ aspect and by extension on themes of global awareness, history and the potential internationality of Big Science.

This is why a reference to Einstein in the headline helps: everyone knows him. More importantly, everyone was recently made aware of how right his theories have been since they were formulated a century ago. So the idea of proving Einstein wrong – as The Wire‘s headline read – is eye-catching. Second, phrasing the headline as a question is a matter of convenience: because the quasi-discovery has a statistical significance of only 2.9 sigma, a question signals doubt.

But if you argued that a question is also a cop-out, I’d agree. A question in a headline can be interpreted in two ways: either as a question that has not been answered yet but ought to be or as a question that is answered in the body. More often than not and especially in the click-bait era, question-headlines are understood to be of the latter kind. This is why I changed The Wire copy’s headline from ‘What if LIGO actually proved Einstein wrong…’ to ‘Has LIGO actually proved Einstein wrong…’.

More importantly, the question is an escapism at least to me because it doesn’t accurately reflect the development itself. If one accounts for the fact that the pre-print paper explicitly states that gravitational-wave echoes have been found in LIGO data only at 2.9 sigma, there is no question: LIGO has not proved Einstein wrong, and this is established at the outset.

Rather, the peg in this case is – for example – that physicists have proposed a way to look for evidence of quantum gravity using an experiment that is already running. This then could make for an article about the different kinds of physics that rule at different energy levels in the universe, and what levels of access humanity has to each.

So this story, and many others like it in the past year that all dealt with observations falling short of the evidence threshold but which have been worth writing about simply because of the desperation behind them, have – or could have – prompted science writers to think about the language they use. For example, the operative words/clause in the respective headlines listed above are:

  • Nature – hint
  • IB Times – might have just caused
  • Futurism – challenging
  • Science Alert – hint
  • Inverse – may not

Granted that an informed skepticism is healthy for science and that all science writers must remain as familiar with this notion as with the language of doubt, uncertainty, probability (and wave physics, it seems). But it still is likely the case that writers grappling with high-energy physics have to be more familiar than others, dealing as the latest research does with – yes – hope and desperation.

Ultimately, I may not be the perfect judge of what words work best when it comes to the fidelity of syntax to sentiment; that’s why I used a question for a headline in the first place! But I’m very interested in knowing how writers choose and have been choosing their words, if there’s any friction at all (in the larger scheme) between the choice of words and the prevailing sentiments, and the best ways to deal with such situations.

PS: If you’re interested, here’s a piece in which I struggled for a bit to get the words right (and finally had to resort to using single-quotes).

Featured image credit: bongonian/Flickr, CC BY 2.0