Comsol -leaderboard other pages

Topics

Still seeking solutions

How did winning a Special Breakthrough Prize last year compare with the Nobel Prize?

Steven Weinberg

It came as quite a surprise because as far as I know, none of the people who have been honoured with the Breakthrough Prize had already received the Nobel Prize. Of course nothing compares with the Nobel Prize in prestige, if only because of the long history of great scientists to whom it has been awarded in the past. But the Breakthrough Prize has its own special value to me because of the calibre of the young – well, I think of them as young – theoretical physicists who are really dominating the field and who make up the selection committee.

The prize committee stated that you would be a recognised leader in the field even if you hadn’t made your seminal 1967 contribution to the genesis of the Standard Model. What do you view as Weinberg’s greatest hits?

There’s no way I can answer that and maintain modesty! That work on the electroweak theory leading to the mass of the W and Z, and the existence and properties of the Higgs, was certainly the biggest splash. But it was rather untypical of me. My style is usually not to propose specific models that will lead to specific experimental predictions, but rather to interpret in a broad way what is going on and make very general remarks, like with the development of the point of view associated with effective field theory. Doing this I hope to try and change the way my fellow physicists look at things, without usually proposing anything specific. I have occasionally made predictions, some which actually worked, like calculating the pion–nucleon and pion–pion scattering lengths in the mid-1960s using the broken symmetry that had been proposed by Nambu. There were other things, like raising the whole issue of the cosmological constant before the discovery of the accelerated expansion of the universe. I worried about that – I gave a series of lectures at Harvard in which I finally concluded that the only way I can understand why there isn’t an enormous vacuum energy is because of some kind of anthropic selection. Together with two guys here at Austin, Paul Shapiro and Hugo Martel, we worked out what was the most likely value that would be found in terms of order of magnitude, which was later found to be correct. So I was very pleased that the Breakthrough Prize acknowledged some of those things that didn’t lead to specific predictions but changed a general framework.

I wish I could claim that I had predicted the neutrino mass

You coined the term effective field theory (EFT) and recently inaugurated the online lecture series All Things EFT. What is the importance of EFT today?

My thinking about EFTs has always been in part conditioned by thinking about how we can deal with a quantum theory of gravitation. You can’t represent gravity by a simple renormalisable theory like the Standard Model, so what do you do? In fact, you treat general relativity the same way you treat low-energy pions, which are described by a low-energy non-renormalisable theory. (You could say it’s a low-energy limit of QCD but its ingredients are totally different – instead of quarks and gluons you have pions). I showed how you can generate a power series for any given scattering amplitude in powers of energy rather than some small coupling constant. The whole idea of EFT is that any possible interaction is there: if it’s not forbidden it’s compulsory. But the higher, more complicated terms are suppressed by negative powers of some very large mass because the dimensionality of the coupling constants is such that they have negative powers of mass, like the gravitational constant. That’s why they’re so weak.

If you recognise that the Standard Model is probably a low-energy limit of some more general theory, then you can consider terms that make the theory non-renormalisable and generate corrections to it. In particular, the Standard Model has this beautiful feature that in its simplest renormalisable version there are symmetries that are automatic: at least to all orders of perturbation theory, it can’t violate the conservation of baryon or lepton number. But if the Standard Model just generates the first term in a power series in energy and you allow for more complicated non-renormalisable terms in the Lagrangian, then you find it’s very natural that there would be baryon and lepton non-conservation. In fact, the leading term of this sort is a term that violates lepton number and gives neutrinos the masses we observe. I wish I could claim that I had predicted the neutrino mass, but there already was evidence from the solar neutrino deficit and also it’s not certain that this is the explanation of neutrino masses. We could have Dirac neutrinos in which you have left and right neutrinos and antineutrinos coupling to the Higgs, and in that way get masses without any violation of lepton-number conservation. But I find that thoroughly repulsive because there’s no reason in that case why the neutrino masses should be so small, whereas in the EFT case we have Majorana neutrinos whose small masses are much more natural.

On this point, doesn’t the small value of the cosmological constant and Higgs mass undermine the EFT view by pointing to extreme fine-tuning?

Yes, they are a warning about things we don’t understand. The Higgs mass less so, after all it’s only about a hundred times larger than the proton mass and we know why the proton mass is so small compared to the GUT or Planck scale; it is because the proton gets it mass not from the quark masses, which have to do with the Higgs, but from the QCD forces, and we know that those become strong very slowly as you come down from high energy. We don’t understand this for the Higgs mass, which, after all, is a term in the Lagrangian, not like the proton mass. But it may be similar – that’s the old technicolour idea, that there is another coupling alongside QCD that becomes strong at some energy where it leads to a potential for the Higgs field, which then breaks electroweak symmetry. Now, I don’t have such a theory, and if I did I wouldn’t know how to test it. But there’s at least a hope for that. Whereas regards to the cosmological constant, I can’t think of anything along that line that would explain it. I think it was Nima Arkani-Hamed who said to me, “If the anthropic effect works for the cosmological constant, maybe that’s the answer with the Higgs mass – maybe it’s got to be small for anthropic reasons.” That’s very disturbing if it’s true, as we’re going to be left waving our hands. But I don’t know.

Maybe we have 2500 years ahead of us before we get to the next big step

Early last year you posted a preprint “Models of lepton and quark masses” in which you returned to the problem of the fermion mass hierarchy. How was it received?

Even in the abstract I advertise how this isn’t a realistic theory. It’s a problem that I first worked on almost 50 years ago. Just looking at the table of elementary particle masses I thought that the electron and the muon were crying out for an explanation. The electron mass looks like a radiative correction to the muon mass, so I spent the summer of 1972 on the back deck of our house in Cambridge, where I said, “This summer I am going to solve the problem of calculating the electron mass as an order-alpha correction to the muon mass.” I was able to prove that if in a theory it was natural in the technical sense that the electron would be massless in the tree approximation as a result of an accidental symmetry, then at higher order the mass would be finite. I wrote a paper, but then I just gave it up after no progress, until now when I went back to it, no longer young, and again I found models in which you do have an accidental symmetry. Now the idea is not just the muon and the electron, but the third generation feeding down to give masses to the second, which would then feed down to give masses to the first. Others have proposed what might be a more promising idea, that the only mass that isn’t zero in the tree approximation is the top mass, which is so much bigger than the others, and everything else feeds down from that. I just wanted to show the kinds of cancellations in infinites that can occur, and I worked out the calculations. I was hoping that when this paper came out some bright young physicist would come up with more realistic models, and use these calculational techniques – that hasn’t happened so far but it’s still pretty early.

What other inroads are there to the mass/flavour hierarchy problem?

The hope would be that experimentalists discover some correction to the Standard Model. The problem is that we don’t have a theory that goes beyond the Standard Model, so what we’re doing is floundering around looking for corrections in the model. So far, the only one discovered was the neutrino mass and that’s a very valuable piece of data which we so far have not figured out how to interpret. It definitely goes beyond the Standard Model – as I mentioned, I think it is a dimension-five operator in the effective field theory of which the Standard Model is the renormalisable part.

Weinberg delivering a seminar at CERN in 1979

The big question is whether we can cut off some sub-problem that we can actually solve with what we already know. That’s what I was trying to do in my recent paper and did not succeed in getting anywhere realistically. If that is not possible, it may be that we can’t make progress without a much deeper theory where the constituents are much more massive, something like string theory or an asymptotically safe theory. I still think string theory is our best hope for the future, but this future seems to be much further away than we had hoped it would be. Then I keep being reminded of Democritus, who proposed the existence of atoms in around 400 BCE. Even as late as 1900 physicists like Mach doubted the existence of atoms. They didn’t become really nailed down until the first years of the 20th century. So maybe we have 2500 years ahead of us before we get to the next big step.

Recently the LHC produced the first evidence that the Higgs boson couples to a second-generation fermion, the muon. Is there reason to think the Higgs might not couple to all three generations?

Before the Higgs was discovered it seemed quite possible that the explanation of the hierarchy problem was that there was some new technicolour force that gradually became strong as you came from very high energy to lower energy, and that somewhere in the multi-TeV range it became strong enough to produce a breakdown of the electroweak symmetry. This was pushed by Lenny Susskind and myself, independently. The problem with that theory was then: how did the quarks and leptons get their masses? Because while it gave a very natural and attractive picture of how the W and Z get their masses, it left it really mysterious for the quarks and leptons. It’s still possible that something like technicolour is true. Then the Higgs coupling to the quarks and leptons gives them masses just as expected. But in the old days, when we took technicolour seriously as the mechanism for breaking electroweak symmetry, which, since the discovery of the Higgs we don’t take seriously anymore, even then there was the question of how, without a scalar field, can you give masses to the quarks and leptons. So, I would say today, it would be amazing if the quarks and leptons were not getting their masses from the expectation value of the Higgs field. It’s important now to see a very high precision test of all this, however, because small effects coming from new physics might show up as corrections. But these days any suggestion for future physics facilities gets involved in international politics, which I don’t include in my area of expertise.

It’s still possible that something like technicolour is true

Any more papers or books in the pipeline?

I have a book that’s in press at Cambridge University Press called Foundations of Modern Physics. It’s intended to be an advanced undergraduate textbook that takes you from the earliest work on atoms, through thermodynamics, transport theory, Brownian motion, to early quantum theory; then relativity and quantum mechanics, and I even have two chapters that probably go beyond what any undergraduate would want, on nuclear physics and quantum field theory. It unfortunately doesn’t fit into what would normally be the plan for an undergraduate course, so I don’t know if it will be widely
adopted as a textbook. It was the result of a lecture course I was asked to give called “thermodynamics and quantum physics” that has been taught at Austin for years. So, I said “alright”, and it gave me a chance to learn some thermodynamics and transport theory.

Martinus J G Veltman 1931–2021

Martinus Veltman

Eminent physicist and Nobel laureate Martinus Veltman passed away in his home town of Bilthoven, the Netherlands, on 4 January. Martinus (Tini) Veltman graduated in physics at Utrecht University, opting first for experimental physics but later switching to theory. After his conscript military service in 1959, Léon Van Hove offered him a position as a PhD student. Veltman started in Utrecht, but later followed Van Hove to the CERN theory division.

CERN opened up a new world, and Tini often mentioned how he benefited from contacts with John Bell, Gilberto Bernardini and Sam Berman. The latter got him interested in weak interactions and in particular neutrino physics. During his time there, Tini spent a short period at SLAC where he started to work on his computer algebra program “Schoonschip”. He correctly foresaw that practical calculations of Feynman diagrams would become more and more complicated, particularly for theories with vector bosons. Nowadays extended calculations beyond one loop are unthinkable without computer algebra.

In 1964 Murray Gell-Mann proposed an algebra of conserved current operators for hadronic matter, which included the weak and electromagnetic currents. He argued that commutators of two currents taken at the same instant in time should “close”, meaning that these commutators can be written as linear combinations of the same set of currents. From this relation one could derive so-called sum rules that can be compared to experiments. Facing the technical problems with this approach, Tini came up with an alternative proposal. In a 1966 paper he simply conjectured that the hadronic currents for the electromagnetic and weak interactions had to be covariantly conserved, where he assumed that the weak interactions were mediated by virtual vector bosons, just as electromagnetic processes are mediated by virtual photons. The current conservation laws therefore had to contain extra terms depending on the photon field and the fields associated with the weak intermediate vector bosons. Quite surprisingly, he could then demonstrate that these new conservation equations suffice to prove the same sum rules. A more important aspect of his approach was only gradually realised, namely that the conservation laws for these currents are characteristic of a non-abelian gauge theory as had been written down more than 10 years earlier by Yang and Mills. Hence Veltman started to work on the possible renormalisability of Yang–Mills theory.

From his early days at CERN it was clear that Tini had a strong interest in confronting theoretical predictions with experimental results

Meanwhile, Veltman had left CERN towards the end of 1966 to accept a professorship at Utrecht. At the end of 1969 a prospective PhD student insisted on working on Yang–Mills theories. Veltman, who was already well aware of many of the pitfalls, only gradually agreed, and so Gerard ’t Hooft joined the programme. This turned out to be a very fruitful endeavour and the work was proudly presented in the summer of 1971. Veltman and ’t Hooft continued to collaborate on Yang–Mills theory. Their 1972 papers are among the finest that have been written on the subject. In 1999 they shared the Nobel Prize in Physics “for elucidating the quantum structure of electroweak interactions”.

With the renormalisability of the electroweak theory established, precision comparisons with experiment were within reach, and Veltman started to work on these problems with postdocs and PhD students. One important tool was the so-called rho parameter (a ratio of the masses of the W and Z bosons and the weak mixing angle). Its experimental value was close to one, which showed that only models in which the Higgs field starts as a doublet are allowed. From the small deviations from one, it was possible to estimate the mass of the top quark, which was not yet discovered. Later, when CERN was planning to build the LEP collider, the emphasis changed to the calculation of one-loop corrections for various processes in e+e collisions. As a member of the CERN Scientific Policy Committee (SPC), Veltman strongly argued that LEP should operate at the highest possible energy, well above the W+W threshold, to study the electroweak theory with precision. The Standard Model has since passed all of these tests.

From his early days at CERN it was clear that Tini had a strong interest in confronting theoretical predictions with experimental results, and in the organisation needed to do so. To this end, he was one of a small group of colleagues in the Netherlands to push for a national institute for subatomic physics – Nikhef, which was founded in 1975. In 1981 Tini moved to the University of Michigan in Ann Arbor, returning to the Netherlands after his retirement in 1996.

Veltman made a lasting impact on the field of particle physics, and inspired many students. Until recently he followed what was happening in the field, regularly attending the September meetings of the SPC. Our community will miss his sharp reasoning and clear-eyed look at particle physics that are crucial for its development.

Connecting physics with society

Student analysing ATLAS collisions

Science and basic research are drivers of technologies and innovations, which in turn are key to solving global challenges such as climate change and energy. The United Nations has summarised these challenges in 17 “sustainable development goals”, but it is striking how little connection with science they include. Furthermore, as found by a UNESCO study in 2017, the interest of the younger generation in studying science, technology, engineering and mathematics is falling, despite jobs in these areas growing at a rate three times faster than in any other sector. Clearly, there is a gulf between scientists and non-scientists when it comes to the perception of the importance of fundamental research in their lives – to the detriment of us all.

Some in the community are resistant to communicate physics spin-offs because this is not our primary purpose

Try asking your neighbours, kids, family members or mayor of your city whether they know about the medical and other applications that come from particle physics, or the stream of highly qualified people trained at CERN who bring their skills to business and industry. While the majority of young people are attracted to physics by its mindboggling findings and intriguing open questions, our subject appeals even more when individuals find out about its usefulness outside academia. This was one of the key outcomes of a recent survey, Creating Ambassadors for Science in Society, organised by the International Particle Physics Outreach Group (IPPOG).

Do most “Cernois” even know about the numerous start-ups based on CERN technologies or the hundreds of technology disclosures from CERN, 31 of which came in 2019 alone? Or about the numerous success stories contained within the CERN impact brochure and the many resources of CERN’s knowledge-transfer group? Even though “impact” is gaining attention, anecdotally when I presented  these facts to my research colleagues they were not fully aware. Yet who else will be our ambassadors, if not us?

Some in the community are resistant to communicate physics spin-offs because this is not our primary purpose. Yet, millions of people who have lost their income as a result of COVID-19 are rather more concerned about where their next rent and food payments are coming from, than they are about the couplings of the Higgs boson. Reaching out to non-physicists is more important than ever, especially to those with an indifferent or even negative attitude to science. Differentiating audiences between students, general public and politicians is not relevant when addressing non-scientifically educated people. Strategic information should be proactively communicated to all stakeholders in society in a relatable way, via eye-opening, surprising and emotionally charged stories about the practical applications of curiosity-driven discoveries.

Barbora Bruant Gulejova

IPPOG has been working to provide such stories since 2017 – and there is no shortage of examples. Take the touchscreen technology first explored at CERN 40 years ago, or humanitarian satellite mapping carried out for almost 20 years by UNOSAT, which is hosted at CERN. Millions of patients are diagnosed daily thanks to tools like PET and MRI, while more recent medical developments include innovative radioisotopes from MEDICIS for precision medicine, the first 3D colour X-ray images, and novel cancer treatments  based on superconducting accelerator technology. In the environmental arena, recent CERN spin-offs include a global network of air-quality sensors and fibre-optic sensors for improved water and pesticide management, while CERN open-source software is used for digital preservation in libraries and its computing resources have been heavily deployed in fighting the pandemic.

Building trust

Credibility and trust in science can only be built by scientists themselves, while working hand in hand with professional communicators, but not relying only on them. Extracurricular activities, such as those offered by IPPOG, CERN, other institutions and individual initiatives, are crucial in changing the misperceptions of the public and bringing about fact-based decision-making to the young generation. Scientists should develop a proactive strategic approach and even consider becoming active in policy making, following the shining examples of those who helped realise the SESAME light source in the Middle East and the South East European International Institute for Sustainable Technologies.

Particle physics already inspires some of the brightest minds to enter science. But audiences never look at our subject with the same eyes once they’ve learned about its applications and science-for-peace initiatives.

Jack Steinberger 1921–2020

Jack Steinberger

Jack Steinberger, a giant of the field who witnessed and shaped the evolution of particle physics from its beginnings to the confirmation of the Standard Model, passed away on 12 December aged 99. Born in the Bavarian town of Bad Kissingen in 1921, his father was a cantor and religious teacher to the small Jewish community, and his mother gave English and French lessons to supplement the family income. In 1934, after new Nazi laws had excluded Jewish children from higher education, Jack’s parents applied for him and his brother to take part in a charitable scheme that saw 300 German refugee children transferred to the US. Jack found his home as a foster child, and was reunited with his parents and younger brother in 1938.

Jack studied chemistry at the University of Chicago until 1942, when he joined the army and was sent to the MIT radiation laboratory to work on radar bomb sights. He was assigned to the antenna group where his attention was brought to physics. After the war he returned to Chicago to embark on a career in theoretical physics. Under the guidance of Enrico Fermi, however, he switched to the experimental side of the field, conducting mountaintop investigations into cosmic rays. He was awarded a PhD in 1948. Fermi, who was probably Jack’s most influential physics teacher, described him as “direct, confident, without complication, he concentrated on physics, and that was enough”.

In 1949 Steinberger went to the Radiation Lab at the University of California at Berkeley, where he performed an experiment at the electron synchrotron that demonstrated the production of neutral pions and their decay to photon pairs. He stayed only one year in Berkeley, partly because he declined to sign the anti-communist loyalty oath, and moved on to Columbia University.

In the 1960s the construction of a high-energy, high-flux proton accelerator at Brookhaven opened the door to the study of weak interactions using neutrino-beam experiments. This marked the beginning of Jack’s interest in neutrino physics. Along with Mel Schwarz and Leon Lederman, he designed and built the experiment that established the difference between neutrinos associated with muons and those associated with electrons, for which they received the 1988 Nobel Prize in Physics.

He was a curious and imaginative physicist with an extraordinary rigour

Jack joined CERN in 1968, working on experiments at the Proton Synchrotron exploring CP violation in neutral kaons. In the 1970s, with the advent of new neutrino beams at the Super Proton Synchrotron, Jack became a founding member of the CERN–Dortmund–Heidelberg–Saclay (CDHS) collaboration. Running from 1976 to 1984, CDHS produced a string of important results using neutrino beams to probe the structure of the nucleon and the Standard Model in general. In particular, the collaboration confirmed the predicted variation of the structure function of the valence quarks with Q2 (nicknamed “scaling violations”), a milestone in the establishment of QCD.

When the Large Electron–Positron (LEP) collider was first proposed, a core group from CDHS joined physicists from other institutions to develop a detector for CERN’s new flagship collider. This initiative grew into the ALEPH experiment, and Jack, a curious and imaginative physicist with an extraordinary rigour, was the natural choice to become its first spokesperson in 1980, a position he held until 1990. From the outset, he stipulated that standard solutions should be adopted across the whole detector as far as possible. This led to the end-caps reflecting the design of the central detector, for example. Jack was also insistent that all technologies considered for the detector first had to be completely understood. As the LEP era got underway, this level of discipline was reflected in ALEPH’s results.

Next to physics, music formed an important part of Jack’s life. He organised gatherings of amateur, and occasionally professional, musicians at his house. These were usually marathons of Bach, starting in the late afternoon and continuing until the late evening. In his autobiography, Jack summarised: “I play the flute, unfortunately not very well, and have enjoyed tennis, mountaineering and sailing, passionately.”

Jack retired from CERN in 1986 and went on to become a professor at the Scuola Normale Superiore di Pisa. President Ronald Reagan awarded him the National Medal of Science in 1988. In 2001, on the occasion of his 80th birthday, the city of Bad Kissingen named its gymnasium in his honour. Jack continued his association with CERN throughout his 90s. He leaves his mark not just on particle physics but on all of us who had the opportunity to collaborate with him.

Accelerating talent at CERN

Natalia Magdalena Koziol

CERN enjoys a world-class reputation as a scientific laboratory, with the start-up of the Large Hadron Collider and the discovery of the Higgs boson propelling the organisation into the public spotlight. Less tangible and understood by the public, however, is that to achieve this level of success in cutting-edge research, you need the infrastructure and tools to perform it. CERN is an incredible hub for engineering and technology – hosting a vast complex of accelerators, detectors, experiments and computing infrastructure. Thus, CERN needs to attract candidates from across a wide spectrum of engineering and technical disciplines to fulfil its objectives.

CERN employs around 2600 staff members who design, build, operate, maintain and support an infrastructure used by a much larger worldwide community of physicists. Of these, only 3% are research physicists. The core hiring needs are for engineers, technicians and support staff in a wide variety of domains: mechanical, electrical, engineering, vacuum, cryogenics, civil engineering, radiation protection, radio­frequency, computing, software, hardware, data acquisition, materials science, health and safety… the list goes on. Furthermore, there are also competences needed in human resources, legal matters, communications, knowledge transfer, finance, firefighters, medical professionals and other support functions.

On the radar

CERN’s hiring challenge takes on even greater meaning when one considers the drive to attract students, graduates and professionals from across CERN’s 32 Member and Associate Member States. In what is already a competitive market, attracting people from a large multitude of disciplines to an organisation whose reputation revolves around particle physics can be a challenge. So how is this challenge tackled? CERN now has a well-established “employer brand”, developed in 2010 to promote its opportunities in an increasingly digitalised environment. The brand centres around factors that make working at CERN the rich experience that it is, namely challenge, purpose, imagination, collaboration, integrity and quality of life – underpinned by the slogan “Take part”. This serves as an identity to devise attractive campaigns through web content, video, online, social media and job-portal advertisements to promote CERN as an employer of choice to the audiences we seek to reach: from students to professionals, apprenticeships to PhDs, across all diversity dimensions. The intention is to put CERN “on the radar” of people who wouldn’t normally identify CERN as a possibility in their chosen career path.

CERN doesn’t just bring together people from a large scope of fields but unites people from all over the world

As no single channel exists that will allow targeting of, for example, a mechanical technician in all CERN Member States, creative and innovative approaches have to be utilised. The varying landscapes, cultural preferences and languages come into play, and this is compounded by the different job-seeking behaviours of students, graduates and experienced professionals through a constantly evolving ecosystem of channels and solutions. A widespread presence is key. The cornerstones are: an attractive careers website; professional networks such as LinkedIn to promote CERN’s employment opportunities and proactively search for candidates; social media to increase visibility of hiring campaigns; and being present on various job portals, for example in the oil, gas and energy arenas. Outreach events, presence at university career fairs and online webinars further serve to present CERN and its diverse opportunities to the targeted audiences.

Storytelling is an essential ingredient in promoting our opportunities, as are the experiences of those already working at CERN. In the words of Håvard, an electromechanical technician from Norway: “I get to challenge myself in areas and with technology you don’t see any other place in the world.” Gunnar, a firefighter from Germany describes, “I am working as a firefighter in one of the most international fire brigades at CERN in what is a very complex, challenging and interesting environment.” Katarina, a computing engineer from Sweden, says, “The diversity of skills needed at CERN is so much larger than what most people know!” While Julia, a former mechanical engineering technical student from the UK put it simply: “I never knew that CERN recruited students for internships.” Natasha, a former software engineering fellow from Pakistan, summed it up with, “Here I am, living my dreams, being a part of an organisation that’s helping me grow every single day.” Each individual experience is a rich insight for potential candidates to identify with and recognise the possibility of joining CERN in their own right.

CERN doesn’t just bring together people from a large scope of fields but unites people from all over the world. Working as summer, technical or doctoral student, as a graduate or professional, builds skills and knowledge that are highly transferable in today’s demanding and competitive job market, along with lasting connections. As the cherry on the cake, a job at CERN paves the way to become CERN’s future alumni and join the ever-growing High-Energy Network. Take part!

Quantum sensing for particle physics

AION’s 10 m stage

A particle physics-led experiment called AION (Atomic Interferometric Observatory and Network) is one of several multidisciplinary projects selected for funding by the UK’s new Quantum Technologies for Fundamental Physics programme. The successful projects, announced in January following a £31 million call for proposals from UK Research and Innovation (UKRI), will exploit recent advances in quantum technologies to tackle outstanding questions in fundamental physics, astrophysics and cosmology.

We have an opportunity to change the way we search for answers to some of the biggest mysteries of the universe

Mark Thomson

UKRI and university funding of about £10 million (UKRI part £7.2 million) will enable the AION team to prepare the construction of a 10 m-tall atomic interferometer at the University of Oxford to explore ultra-light dark matter and provide a pathway towards detecting gravitational waves in the unexplored mid-frequency band ranging from several mHz to a few Hz. The setup will use lasers to drive transitions between the ground and excited states of a cloud of cold strontium atoms in free fall, effectively acting as beam splitters and mirrors for the atomic de Broglie waves (see figure). Ultralight dark matter and exotic light bosons would be expected to have differential effects on the atomic transition frequencies, while a passing gravitational wave would generate a strain in the space through which the atoms fall freely. Either would create a difference between the phases of atomic beams following different paths – the greater their separations, the greater the sensitivity of the experiment.

“AION is a uniquely interdisciplinary mission that will harness cold-atom quantum technologies to address key issues in fundamental physics, astrophysics and cosmology that can be realised in the next few decades,” says AION principal investigator Oliver Buchmueller of Imperial College London, who is also a member of the CMS collaboration. “The AION project will also significantly contribute to MAGIS, a 100 m-scale partner experiment being prepared at Fermilab, and we are exploring the possibility of utilising a shaft in the UK or at the LHC for a similar second 100 m detector.”

Six other quantum-technology projects involving UK institutes are under way thanks to the UKRI scheme. One, led by experimental particle physicist Ruben Saakyan of University College London, will use ultra-precise B-field mapping and microwave spectrometry to determine the absolute neutrino mass in tritium beta-decay beyond the 0.2 eV sensitivity projected for the KATRIN experiment. Others include the use of new classes of detectors and coherent quantum amplifiers to search for hidden structure in the vacuum state; the development of ultra-low-noise quantum electronics to underpin searches for axions and other light hidden particles; quantum simulators to mimic the extreme conditions of the early universe and black holes; and the development of quantum-enhanced superfluid technologies for cosmology.

The UKRI call is part of a global effort to develop quantum technologies that could bring about a “second quantum revolution”. Several major international public and private initiatives are under way. Last autumn, CERN launched its own quantum technologies initiative.

“With the application of emerging quantum technologies, I believe we have an opportunity to change the way we search for answers to some of the biggest mysteries of the universe,” said Mark Thomson, executive chair of the UK’s Science and Technology Facilities Council. “These include exploring what dark matter is made of, finding the absolute mass of neutrinos and establishing how quantum mechanics fits with gravity.”

The hitchhiker’s guide to weak decays

Unexplored territory

Most travellers know that it is essential to have a good travel guide when setting out into unexplored territory. A book where one can learn what previous travellers discovered about these surroundings, with both global information on the language, history and traditions of the land to be explored, and practical details on how to overcome day-to-day difficulties. Andrzej Buras’s recent book, Gauge Theory of Weak Decays, is the ideal guide for both new physicists and seasoned travellers, and experimentalists and theoreticians alike, who wish to start a new expedition into the fascinating world of weak meson decays, in pursuit of new physics.

The physics of weak decays is one of the most active and interesting frontiers in particle physics, from both the theoretical and the experimental points of view. Major steps in the construction of the Standard Model (SM) have been made possible only thanks to key observations in weak decays. The most famous example is probably the suppression of flavour-changing neutral currents in kaon decays, which led Glashow, Iliopoulos and Maiani to postulate, in 1970, the existence of the charm quark, well before its direct discovery. But there are many other examples, such as the heaviness of the top quark, inferred from the large matter-to-antimatter oscillation frequency of neutral B mesons, again well before its discovery. In the post-Higgs-discovery era, weak decays are a privileged observatory in which to search for signals of physics beyond the SM. The recent “B-physics anomalies”, reported by LHCb and other experiments, could indeed be the first hint of new physics. The strategic role that weak decays play in the search for new physics is further reinforced by the absence on the horizon, at least in the near future, of a collider with a centre-of-mass energy exceeding that of the LHC, while the LHC and other facilities still have a large margin of improvement in precision measurements.

As Buras describes with clarity, signals of new physics in the weak decays of K, D, and B mesons, and other rare low-energy processes, could manifest themselves as deviations from the precise predictions of the corresponding decay rates that we are able to derive within the SM. In the absence of a reference beyond-the-SM theory, it is not clear where, and at which level of precision, these deviations could show up. But general quantum field theory arguments suggest that weak decays are particularly sensitive probes of new physics, as they can often be predicted with high accuracy within the SM.

The two necessary ingredients for a journey in the realm of weak decays are therefore precise SM predictions on the one hand, and a broad-spectrum investigation of beyond-the-SM sensitivity on the other. These are precisely the two ingredients of Buras’s book. In the first part, he guides the reader though all the steps necessary to arrive to the most up-to-date predictions for rare decays. This part of the book offers different paths to different readers: students are guided, in a very pedagogical way, from tree-level calculations to high-precision multi-loop calculations. Experienced readers can directly find up-to-date phenomenological expressions that summarise the present knowledge on virtually any process of current experimental interest. This part of the book can also be viewed as a well-thought-out summary of the history of precise SM calculations for weak decays, written by one of its most relevant protagonists.

Beyond the Standard Model

The second part of the book is devoted to physics beyond the SM. Here the style is quite different: less pedagogical and more encyclopaedic. Employing a pragmatic approach, which is well motivated to discuss low-energy processes, extensions of the SM are classified according to properties of hypothetical new heavy particles, from Z′ bosons to leptoquarks, and from charged Higgs bosons to “vector-like” fermions. This allows Buras to analyse the impact of such models on rare processes in a systematic way, with great attention paid to correlations between observables.

To my knowledge, this book is the first comprehensive monograph of this type, covering far more than just the general aspects of SM physics, as may be found in many other texts on quantum field theory. The uniqueness of this book lies in its precious details on a wide variety of interesting rare processes. It is a key reference that was previously missing, and promises to be extremely useful in the coming decades.

Alexei Onuchin 1934–2021

Alexei Onuchin

Alexei Onuchin, one of the pioneers of experiments at colliding beams, passed away on 9 January in Novosibirsk, Russia.

Onuchin was born in 1934 in a small village in the Gorky (Nizhny Novgorod) region. After graduating from high school with honours, he decided to try his hand at science and in 1953 he entered the physics department of Moscow State University. In 1959 he graduated with honours and was invited by Gersh Budker to work at the newly organised Institute of Nuclear Physics at Novosibirsk (INP).

At INP, Onuchin enjoyed many important roles. He took part in experiments at the world’s first electron–electron collider (VEP-1), actively worked on the preparation of a detector for the electron–positron collider VEPP-2, supervised the construction of the MD-1 detector for the VEPP-4 collider, was one of the leaders of the KEDR detector experiment at the VEPP-4M collider and was a great enthusiast of the detector project for the proposed Super Charm-Tau Factory. He was also an organiser and for many years the leader of the Budker INP group working at the BaBar experiment at SLAC.

During his career, Alexei made a great contribution to the development of experimental techniques in particle physics. It was this that determined the high level of experiments carried out at Budker INP and other laboratories. These include the development and production of multiwire proportional chambers for the MD-1, various counters based on Cherenkov radiation and the creation of a calorimeter based on liquid krypton, among many others.

Cherenkov counters held a special place in Alexei’s heart from the very beginning of his career as a student in the laboratory of Nobel-laureate Pavel Cherenkov. Starting from pioneering water-threshold counters in the experiment at VEPP–2, he later developed the MD–1 Cherenkov counters filled with ethylene pressurised to 25 bar, and finally suggested the aerogel counters with wavelength shifters (ASHIPH) now operating in the KEDR detector. For this work, in 2008 Alexei Onuchin was awarded the Cherenkov Prize of the Russian Academy of Sciences.

Alexei was a great teacher. Among his former students are professors, group leaders and members of the Russian Academy of Sciences. He was also a caring father and loving husband, who raised a large family with four children, five grandchildren and three great-grandchildren. He will always be remembered by his family, friends and colleagues.

A day with particles

Outreach must continue, even in a pandemic: if visitors can’t come to the lab, we need to find ways to bring the lab to them. Few outreach initiatives do this as charmingly as “A day with particles” — a short independent film by three ATLAS physicists at Charles University in Prague. Mixing hand-drawn animations, deft sound design and a brisk script targeted at viewers with no knowledge of physics, the 30 minute film follows a day in the life of postdoc Vojtech Pleskot. In its latest pitstop in a worldwide tour of indie film festivals, it won “BEST of FEST (Top Geek)” last week at GeekFestToronto.

We want to break stereotypes about scientists

Vojtech Pleskot

“We just want to show that scientists are absolutely normal people, and that no one needs to fear them,” says Pleskot, who wrote and directed the film alongside producer Martin Rybar and animator Daniel Scheirich. Modest and self-effacing when I interviewed them, the three physicists produced the film with no funding and no prior expertise, beating off competition from well funded projects to win the Canadian award. Even within the vibrant but specialist niche of high-energy-physics geekery, competition included “The world of thinking”, featuring interviews with Ed Witten, Freeman Dyson and others, and a professionally produced film dramatising a love letter from Richard Feynman to his late wife Arline, who passed away while he was working on the Manhattan Project. But Pleskot, Rybar and Scheirich won the judges over with their idiosyncratic distillation of life at the rock-face of discovery. The trio place their film in the context of growing scepticism of science and scientists. “We want to break stereotypes about scientists,” adds Pleskot.

Not every stereotype is broken, and there is room to quibble about some of the details, but grassroots projects such as A Day With Particles boast a quirky authenticity which is difficult to capture through institutional planning, and is well placed to connect emotionally with non-physicists. The film is beautifully paced. Wide-eyed enthusiasm for physics cuts to an adorable glimpse of Pleskot’s two “cute little particles” having breakfast. A rapid hop from Democritus to Rutherford to the LHC cuts to tracking shots of Pleskot making his way through the streets of Prague to the university. The realities of phone conferences, failed grid jobs and being late for lab demonstrations are interwoven with a grad student dancing to discuss her analysis, conversations on free-diving with turtles and the stories of beloved professors recalling life in the communist era. Life as a physicist is good. And life as a physicist is really like this, they insist. “I hope that science communicators will share it far and wide,” says Connie Potter (CERN and ATLAS), who commissioned the film for the 2020 edition of ICHEP, and who was also recognised by the Toronto festival for her indefatigable “indie spirit” in promoting it.

A Day With Particles will next be considered at the World of Film International Festival in Glasgow in June, where it has been selected as a semi-finalist.

Scientific journeys of a “Sputnik kid”

Scientific journeys

H Frederick Dylla is a “Sputnik kid”, whose curiosity and ingenuity led him on a successful 50-year career in physics, from plasma to accelerators and leading the American Institute of Physics. His debut book, Scientific Journeys: A Physicist Explores the Culture, History and Personalities of Science, is a collection of essays that puts a multidisciplinary historical perspective on the actors and events that shaped the world of science and scholarly publishing. Through geopolitical and economic context and a rich record of key events, he highlights innovations that have found their use in social and business applications. Those cited as having contributed to global technological progress range from the web and smartphones to medical imaging and renewable energy.

Dylla begins with the story of medieval German abbess, mystic, composer and medicinal botanist Hildegard of Bingen

The book is divided in five chapters: “signposts” (in the form of key people and events in scientific history); mentors and milestones in his life; science policy; communicating science; and finally a brief insight into the relationship between science and art. He begins with the story of medieval German abbess, mystic, composer and medicinal botanist Hildegard of Bingen: “a bright signpost of scholarship”. Dylla goes on to explore the idea that a single individual at the right time and place can change the course of history. Bounding through the centuries, he highlights the importance of science policy and science communication, the funding of big and small science alike, and the contemporary challenges linked to research, teaching science and scholarly publishing. Examples among these, says Dylla, are the protection of scientific integrity, new practices of distance learning and the weaknesses of the open-access model. The book ends bang up to date with a thought on the coronavirus pandemic and science’s key role in overcoming it.

Intended for teachers, science historians and students from high school to graduate school, Dylla’s book puts a face on scientific inventions. The weightiest chapter, mentors and milestones, focuses on personalities who have played an important role in his scientific voyage. Among the many named, however, Mildred Dresselhaus – the “queen of carbon” – is the only female scientist featured in the book besides Hildegard. Though by beginning the book with a brilliant but at best scientifically adjacent abbess who preceded Galileo by four centuries Dylla tacitly acknowledges the importance of representing diversity, the book unintentionally makes it discomfortingly clear how scarce role models for women can be in the white-male dominated world of science. The lack of a discussion on diversity is a missed opportunity in an otherwise excellent book.

bright-rec iop pub iop-science physcis connect