Category Archives: Physics

Making Science Inclusive

It’s International Women’s Day, and people seem to find it easier to support girls than women.

I’ve noticed this as a Woman In Science, this eagerness to encourage girls into science with no concern as to what might happen to them as they, inevitably, become Women In Science. Isn’t that what we want for them, to evade the leaks in the pipeline and become role models that future girls can look up to? Don’t we want girls to become Women Who Have It All?

And yet, it’s easier to support girls than women – girls aren’t threatening. Girls aren’t competition. How else to explain the fact that, at this stage in my career, I face more sexism than I ever have before? I went off to college in 2001, at the ripe old age of 16, and you’d think things would have gotten better since then.

But when you’re part of a student cohort, or even a postdoctoral researcher or senior postdoc, you’re classed with other researchers at your level. As a PI on the other hand, running a research group, teaching undergrads, applying for funding, suddenly I am being treated worse than at any previous time in my career. Sure, it’s a demanding job, but I can’t help but notice the female junior academics around me getting saddled with heavier workloads and negative attitudes about their gender that male academics don’t have to deal with. And it doesn’t make my job easier when I can’t go to a conference without being asked ‘who do you work for?’

I have to admit I thought that as a society we’d be over this by now. Naïve student Jessamyn would have assumed there would be no need for gender quotas, in a place as progressive as a university setting, in the year 2019. Lecturer Jessamyn grimly admits that we still need them, and we have a long way to go before we are truly including everyone in higher education, and in science. Science is for everyone, regardless of gender, race, class, sexuality, or background, but it’s a lot of work to make that happen. Initiatives like the Athena SWAN and IOP Juno awards are a step in the right direction, and I’m glad to see my own institution pursuing them, but I think of them as being like the Bechdel Test – a necessary minimum, but not nearly enough to ensure true inclusion. We need to make sure that everyone is part of the story, not just the usual suspects.

Sometimes I find this tough going. I work in physics, a field with a pretty bad diversity problem, and I am used to being the only woman in the room. I wish I didn’t have so much experience being put down or disrespected, and while it may me a minority of physicists who act this way, a few consistent bad experiences can really change the environment. I sometimes wonder about the ethics of encouraging young girls into physics, having had the experiences I have: am I shepherding them into a place where they won’t be valued?

But you know, when I was an undergraduate at the cusp of either leaving physics or doubling down and pursuing a graduate education, I was lucky to end up working for an amazing physicist who I looked up to. She was an inspiration, a wonderful and supportive supervisor, and a role model whether she intended to be or not. In fact she still is – her name is Dr. Natalie Roe, and she is now the Physics Director of Lawrence Berkeley National Lab. I’m sure she faced adversity in her career, and had experiences that differed from her male colleagues, but when I looked at her as a student I saw an incredible scientist and person, who I wanted to emulate. She paved a road for me into physics, and probably for many other young women who she mentored.

Culture change is slow, and frustrating, and in fighting injustice we often feel like we are far behind where we want to be. But I am grateful for everyone who’s pushed for change, and I hope that we can all keep paying this forward, to make science the kind of place where every girl, and every woman, is truly welcome.

Expressing Nanoscience through Dance

Growing up, I was a dancer – I performed in salsa, swing, and ballroom competitions, trained a little in ballet, and was the captain of my high school dance team. When I went off to college to study physics and math, I never imagined that I would get to bring science and dance together someday.

But in 2017, when I was selected for a science/art residency aboard a ship in the Arctic, my roommate for three weeks in close quarters was Deidre Cavazzi, a choreographer specializing in interdisciplinary dance projects. Deidre was a wonderful companion on the journey we shared, and is now a good friend, so I was delighted when she suggested that she might be able to come to Ireland during her next sabbatical to choreograph a dance piece based on my research in nanoscience. I felt really honoured, because I had seen her previous work based on things like the Fibonacci sequence and banned books and to me, the idea of translating these ideas into physical movement and shape and tempo was fascinating.

So it was very exciting when Deidre came to Galway in autumn 2018, supported by a public engagement grant from the Institute of Physics. She came to my lab, talked with me about my research, and read everything she could get her hands on about nanoelectronics and memristors and novel devices that are mimicking the brain. We got a beautiful venue courtesy of the Discipline of Drama, Theatre, and Performance at NUI Galway, and then for two nights during Science Week we invited people to a free event where I gave a short introduction to nanoscience, and then Deidre introduced a dance theatre piece that explored the same concepts, with images from my research and movement choreographed and set to music by Deidre. The full show was recorded, and you can watch it here:

Video of my talk and the dance performance that Deidre created.

One of the best things for me about this project was that Deidre asked me if I wanted to be one of the dancers! I had a wonderful time, and seeing how she brought nanoscience concepts to a whole new context was truly inspiring. You can read more about Deidre’s process on her blog here, where she describes her process and her time in Ireland. I enjoyed my collaboration with her so very much, and we are hoping to repeat it again sometime in the future! But I couldn’t actually sum up the project better than this quote, from one of our audience members:

The scientist and the choreographer had understood each other so well… I loved the blending of science and art, I think both can benefit hugely from each other as each has a unique perspective but are trying to answer similar questions.

Why teach physics?

To me, teaching at the university level has three specific roles:

  • Student empowerment
  • Democratization of knowledge
  • The betterment of humanity

As an educator, my job is first and foremost to empower students by helping them to learn, access information, and create new knowledge. This process is inherently democratic as it is meant to equalize access to knowledge, so that any student can learn any subject regardless of their gender, race, socioeconomic status, or other factors which they cannot control. Empowering individual students in this way leads to a society that contains more and more well-informed, competent and capable individuals, who will use their knowledge and talents to contribute to the betterment of the human condition. My teaching is meant to be a public good.

The author closing TEDxTUM 2017. Photo by Wade Million.

This may seem like an obvious, if lofty goal. However, there are challenges which must be addressed, especially in the teaching of science generally and physics more specifically. Physics requires not only topical knowledge and strong math skills, but also critical thinking and problem solving, which can at first seem at odds with the way that many students have prepared themselves for the Leaving Cert. In the Irish context, teaching first year physics, I try to help students transition from a rote learning mentality to embrace more complex modes of understanding. I must also address a common fallacy about science, that it is a collection of facts. Showing students how these facts are connected, and that science is a creative endeavour at heart, is needed to progress their understanding.

There is also an elitism around physics, sometimes perpetuated by physicists themselves, that must be broken down. Physics relies on math in a way that few other subjects do, and while math is essential to understanding physical concepts, many students experience ‘math anxiety’ that must be addressed. Using language, demonstrations, and math together helps students cement the basic concepts. Studies have shown that the use of humour in lectures also aids students to remember information better, which comes as little surprise to me after my experience running Bright Club.

A talk I gave about using humour to communicate tough topics.

Often, whether or not students choose to pursue a topic depends not so much on the topic itself as on their own identity: whether or not they can see themselves in physics, whether they ‘fit in’. My own experience growing up surrounded by scientists helped show me what a scientific career looked like and how to access it. But students without this background need more from their instructors: to see scientists as real people, to understand the steps to a scientific career and where it can lead them. This is especially important for women and other underrepresented minorities, who are still sorely needed in physics. Even if the students who take my classes choose not to study physics, I aim to leave them with a solid understanding and positive attitude toward physics, to understand that physics is a way to understand the world around us, and an important part of everything we do.

I do my best to actively engage students in lectures, problem-solving, and laboratory, remembering that people often say that you cannot teach anyone physics, but empower them to learn it themselves. My role is to find as many ways as there are students to present the information and connective ideas of physics. I am inspired in this by my own undergraduate physics mentor, Richard Muller at the University of California Berkeley, who was famous as a teacher for his course Physics for Future Presidents. He understood that everyone, not just future physicists, will benefit from a solid understanding of physics, and designed a seminar style course focused more on interesting and relevant science topics rather than historical sequences of discovery. I take the liberal arts view that physics is for everyone, the same way literature is for everyone.

My aim as a teacher is not merely to get as many people as possible to study physics. It is to improve engagement, understanding, and attitudes toward physics, which is a central science of relevance to every single person. I want to empower students to understand physics, improve their access to that knowledge and understanding of how it is connected, and help scientists and non-scientists alike to work for the betterment of humanity. This makes better physicists, but it also makes better people.

The World’s Smallest Movie

One of the most incredible movies ever made is this stop-motion animated film, made by moving carbon monoxide molecules around on the surface of copper at IBM. It’s incredible to think that when we look at this, we are seeing quantum mechanics – bonding at the nanoscale, and echoes of the wave nature of reality.

There is a short documentary that explains how the movie was made, manipulating the molecules with scanning tunnelling microscopy, which is also worth a watch:

Women in Science at the End of the World

It is one thing to use science to better understand the world, another to fear the world itself is crumbling all around you. And yet the scientists who were pursuing research during World War II must have felt both these things keenly, as the Great Powers became embroiled in the second major war in a generation.

Against this backdrop, scientific advances were about to become very
important to the course of the war, and the public perception of science was about to be changed indelibly. Researchers in Europe and the United States were digging to the heart of nuclear fission, an understanding of how the nuclei at the heart of atoms could split, changing into other elements in a naturally-occurring process. Fission was also thought to release an unheard of amount of energy, which in wartime led to one obvious thought: was it possible to use fission to build a bomb?

The Project
After a report from the UK was shared with the US Army, which coordinated the results of a series of secret conferences to discuss the possibility of a fission bomb and how it might be designed, the Army Corps of Engineers launched what was called ‘the Manhattan Project’. Major General Leslie Groves was put in charge, and appointed as scientific director Robert Oppenheimer, an expert in neutron collisions at the University of California Berkeley (the only university with a particle accelerator powerful enough to make plutonium, which had been recently discovered in 1941). Oppenheimer’s first task was to find a suitable location to build a lab where a fission bomb could be designed, built, and eventually tested.

As a child, Oppenheimer suffered from tuberculosis and recovered at the Los Alamos Ranch School in the New Mexico mountains. Far from any major settlements, this location seemed ideal to Oppenheimer and he suggested it as the main site for the Manhattan Project. The land at Los Alamos was purchased by the US government in late 1942, with scientific work beginning there in 1943. Initially General Groves had imagined a military installation, with the scientists in uniforms and posted away from their families. But key scientists balked at uniforms and many wished to bring their families with them. The Los Alamos scientists, working in secret, are often considered a boys’ club, plus wives. Yet even in wartime, and facing prejudice their male counterparts did not, women made huge scientific contributions to the success of the Manhattan Project.

Lise Meitner


The physicist whose work set the scene for the development of the fission bomb, though she would not have wished it, was Austrian-born Lise Meitner. Working in pre-war Germany, she and long-time collaborator Otto Hahn developed the theory and the experimental understanding of nuclear fission. But Meitner was uprooted from Germany due to the Nuremberg laws and her Jewish heritage, and she was forced to flee to Stockholm to continue her work. Her German-based colleagues left her name off several key papers, fearing repercussions from the Nazi authorities. And so the Nobel Prize for Physics in 1944 was then awarded to Meitner’s main collaborator, Otto Hahn, for the discovery of nuclear fission. Although the Nobel committee later revealed that the complications of wartime and the difficulty of assessing interdisciplinary work had contributed to Meitner’s omission from the prize, the US Army had recognized her contribution and invited her to join the Manhattan Project. Her experimental and theoretical insight had clearly been critical, and she was no longer ensnared in Nazi Germany as Hahn was. But Meitner refused to join the Project, saying ‘I will
have nothing to do with a bomb!’ Her advancement of the theory of nuclear fission was, nevertheless, critical to the Project’s success.

Lilli Hornig


Among the women who did join the Project was Lilli Hornig, a Czech chemist who specialized in the newly discovered element plutonium. She had a master’s degree from Harvard when the Manhattan Project began and came to Los Alamos married to an explosives scientist, having been told that anyone
with a chemistry background would be welcomed on board.

Upon arrival, however, Horning was offered a typing job, and was only permitted to work on plutonium chemistry after saying she did not know how to type. Hornig was moved to the explosives group once lab management realised the intense radioactivity of plutonium might cause reproductive damage. ‘I tried delicately to point out that they might be more susceptible than I was; that didn’t
go over well,’ she said in an interview with Manhattan Project Voices (a public
archive of oral histories).

Hornig was eventually a witness to the ‘Trinity’ test of the first so-called ‘atomic
bomb’ in Alamogordo, New Mexico. The Trinity detonation used plutonium as its fissile material, standing on the shoulders of Hornig’s work, but after seeing the devastation it was capable of, she signed a letter along with 100 other scientists requesting that the bomb be demonstrated to the Japanese on an uninhabited island. The next two nuclear detonations occurred over the cities of Hiroshima and Nagasaki, killing over 100,000 people. After the war had ended, Hornig went back to graduate school, getting her PhD and becoming a chemistry professor at Brown University. She was a feminist and a passionate advocate for women in science, studying inequality in the sciences alongside her first love of chemistry.

Maria Goeppert Mayer


Maria Goeppert Mayer was a German physicist, whose doctoral thesis was super-
vised by Max Born, the father of quantum mechanics. Goeppert Mayer came to America when her husband took employment as a professor at Johns Hopkins, but nepotism laws at the time prevented the wife of a professor from being employed at the same institution. Initially she worked unpaid, collaborating with others and eventually studying the separation of different atomic isotopes. The couple then moved to Columbia University and there Goeppert Mayer began her work with the Manhattan Project, studying isotope separation of uranium compounds to be purified into fissile fuel.

While comparing different isotopes, she began to notice ‘magic numbers’ of nucleons which led to more stable atomic nuclei. In 1945, she went to Los Alamos to work directly with Edward Teller on the successor to the atomic bomb: the hydrogen bomb, which exploited the energy from the atomic fusion of elements.

After the war, Goeppert Mayer continued to develop her shell model of the atomic nucleus, which she likened to pairs of waltzers at a dance: each nucleon was a waltzer paired with another waltzer, and more waltzing couples could be fitted into the nucleus by having some go clockwise, some anticlockwise, paralleling nuclear spin. She finally achieved her first full-time paid position as a scientist in 1960 at the University of California San Diego, after receiving her PhD in 1930. When she received the Nobel Prize in Physics in 1963, for her shell model of the atomic nucleus, the local newspaper headline read ‘S. D. mother wins Nobel Prize’. Goeppert Mayer was the second female Nobel Laureate in physics, after Marie Curie.

Each of these women had critical scientific and technical contributions to the Manhattan project, but the politics of the time and secrecy surrounding the wartime effort shrouded the impact of their work. The contributions of male scientists to the Project are more widely touted, and although Hornig, Goeppert Mayer, and Meitner are long gone, the scientific establishment today still struggles to appropriately acknowledge the contributions of women.

Furthermore, that a Czech, a German, and an Austrian were so central to ending an international war, displaced from their countries of origin, shows the value of immigration and even of wartime refugees. Despite difficulties accessing academia and key resources, female scientists played a major role in the Manhattan Project, the building of the first atomic bomb, and the end of World War II.

What Is Life? And Other Interdisciplinary Questions

Scientists, like most people, want to understand the world they live in. We examine the physical structure of the world, in the hope that understanding the rules governing it will also lend some clue as to what it all means. Ironically, as the boundaries of scientific knowledge grow, the possibility of any individual scientist grasping this entire meaning gets smaller and smaller. Even within science, the traditional disciplinary boundaries—biology, chemistry, physics—often separate scientists who should really be talking to each other. I have always loved talking to people who know different things than I do, prompted by curiosity and encouraged by family. But it can be rare to see famous scientists doing the same.

Erwin Schrödinger, a physicist famous for his mathematical and philosophical development of quantum mechanics, tried to reach across these boundaries and ask the question that compels so many of us, namely ‘What Is Life?’ His historic lectures were given in 1943, against the backdrop of world war and 75 years behind our current understanding of biology. And in a wise move for anyone trying to understand something new, Schrödinger began by admitting what he didn’t know: to him, the difference between what physics and biology had to say about life was the difference between “a wallpaper and a Raphael tapestry”.

And yet, connecting physical understanding to biology has provided significant insights. When these lectures were given, the DNA molecule itself had already been discovered, but the double helix structure was yet to be found, along with much of our modern understanding of how this blueprint for life works. And yet the stability of the DNA molecule in the cell is directly due to the quantum basis of chemical bonds. Schrödinger expresses amazement that even with the perturbations caused by heat and environment at the molecular level, the naive physicist’s expectation of wild variability is incorrect, and chemical stability holds. While DNA does sometimes mutate in ways that persist through generations, forming the basis of natural selection, its ability to reproduce error free throughout our lives is amazing from a physics perspective. Especially when one considers the consequences of uncontrolled mutation, as the world would see only two years later as radiation-induced mutation caused terrible illness in the survivors of Hiroshima and Nagasaki.

In trying to define life, Schrödinger comes to the idea of order and disorder, and the physicist’s idea of entropy. Although entropy, which is a measure of disorder, is bound to increase over time subject to the Second Law of Thermodynamics, Schrodinger posited that living beings were effectively decreasing their local entropy by exporting it, increasing order within the cell even if the broader environment became less ordered. Cheating the Second Law of Thermodynamics is a necessity for living cells, living beings, and even our planet to maintain local order. Schrodinger then concludes that living beings must be ‘negative entropy machines’, converting energy to local order, a perspective only a physicist could have come up with.

Schrödinger’s willingness to admit what he did not know, and try to combine modern biology and modern physics even during wartime to unify humanity in knowledge, put me in mind of another, less famous, transdisciplinary scientist.

A man pipetting.

My father, Eric Fairfield, was a biochemistry professor who left academia to work on the Human Genome Project. We talked about science a lot, especially once I chose to pursue physics. I know he was proud of me for becoming a scientist, even though as a biochemist he could not resist ribbing me for my limited understanding of biology. When I read Schrodinger’s statement:

THE CLASSICAL PHYSICIST’S EXPECTATION, FAR FROM BEING TRIVIAL, IS WRONG

I could nearly hear it in my dad’s voice. The “naive physicist’s approach” to understand the cell by looking to statistical physics and randomnessi misses the stability of chemical bonds in the DNA molecule and other cellular components. In a messy, changeable environment, the blueprints that make us have persisted through thousands of generations. As my dad used to say, biology had this figured out a long time before we even knew what questions to ask.

But this isn’t to say that physicists have no business asking questions in biology, or vice versa. Biology is built on the laws of physics and chemistry, even if the exact details of how are still being puzzled out today. And questions that my dad put to me, as part of his own research, often had me questioning both physics and biology. How does a cell know what organ to build a piece of? What biochemical signals lead to the evolution of our own sensory organs, like ears or eyes? How does higher level order arise from molecule level decisions?

I enjoyed discussing these questions with him, and asking my own about the chemistry of the nanomaterials I studied, and their current and possible future biological applications. But the last big biological questions my dad asked ended up being about cancer, a scientific issue that has absorbed the careers of many researchers. Colon cancer took my father’s life last year, and when I have a question about biology, I can no longer call him to see if he’s thought about it before. At his memorial service, many friends commented how much they enjoyed talking science with my dad, whether they had a scientific background or not. He enjoyed discussing and debating these topics with anyone, even if and sometimes especially if they had a vastly different perspective to his own. But I think science would get a little bit further if we had more scientists like my father, or like Erwin Schrödinger, who were willing to cross disciplinary boundaries, admit what they don’t know, and see where they can go from here.

My scientific colleagues may find this to be a very personal response to a scientific matter. But Schrödinger himself dedicated What Is Life? to the memory of his parents. We are all searching for answers together, inspired by those who have come before, and certain to be surprised by what comes next.

When Your Science Hero is Problematic

We all have heroes, people we look up to and whose achievements spur us on to do our own personal best. And, especially in this era where women are saying #metoo and finally being heard, we have probably all had the experience of finding out that one of our heroes has done some less than heroic things. This has come up a lot for me recently with the deaths of some very famous scientists and science fiction writers, men I greatly admired when I was a kid, who I’m now discovering were frequently awful to women (i.e., people like me).

I think this happens more than usual in science, a traditionally male-dominated field where a culture of elitism and privilege has been embedded for a long time. And it’s tempting to view things in black and white: either my hero is amazing for their achievements or they are garbage for their behaviour. We know in our personal lives that people are multi-faceted, yet we’re slow to allow public figures that same understanding. If a famous male scientist discovers lots of things, and is a great collaborator with other men but acts differently toward women, consciously or unconsciously, how are we meant to think about that?

As a physicist who loves to write, I’ve had to consider this before, because one of my early science heroes was Richard Feynman. Feynman was a brilliant theoretical physicist, a Nobel laureate, and worked on the Manhattan project building the atomic bomb in my hometown of Los Alamos. He also wrote a series of very enjoyable popular science books, which were also quite personal and effortlessly engaging. A quote from an interview that immediately stuck with me:

Omni: As we came back to the office, you stopped to discuss a lecture on color vision you’ll be giving. That’s pretty far from fundamental physics, isn’t it? Wouldn’t a physiologist say you were ‘poaching’?

Feynman: Physiology? It has to be physiology? Look, give me a little time and I’ll give a lecture on anything in physiology. I’d be delighted to study it and find out all about it, because I can guarantee you it would be very interesting. I don’t know anything, but I do know that everything is interesting if you go into it deeply enough.

As someone who is omnivorous about knowledge, I found that quote resonated with me deeply. Science is fascinating because it shows us how the world works, how things which might appear separate are deeply connected, and the overlapping intricacies behind the everyday we take for granted. I now do my research on nanoscience, a strongly interdisciplinary field that draws from chemistry, electrical engineering, materials science, and plenty more beyond the physics that I got my degrees in. I admired Feynman for not letting other people dictate the questions he could ask, for being a physicist in what felt like a subversive and wide-ranging way. He was also famous for his sense of humour, his love of non-scientific things like playing bongos, and for generally not being as formal and rigid about anything as physicists tend to be.

The author having a Feynman bongo moment at the No-Ball Prizes. Photo by Ian Bowkett.

Of course, if you read Feynman’s books you’ll also find less inspiring stories, if you are a female scientist. He writes about doing his calculations in a Hooters, negging women in bars, and pretending to be an undergraduate to pick up grad students’ wives. This is less subversive, and more what we might generously call ‘of a time’. Feynman did plenty to promote the status of women in physics, encouraging his own sister to study it and eventually get a PhD. But reading through these differing accounts of his behaviour, female physicists are left wondering whether this great man of science would have seen them as colleagues and equals, or as prey.

I still find a lot in Feynman to look up to, as a physicist who did amazing work but cared about communication and didn’t give in to pressure to conform. However I can still acknowledge the women he mistreated, or perhaps even drove out of the field which is a terrible loss to science. He had a complexity to him, and my initial hero-worship of Feynman when I was younger has been replaced by equally complex feelings, of respect for his scientific and communication work alongside frustration at his mistreatment of women. But there’s no such thing as a perfect hero anyway, and if I needed one in physics, I might be waiting a long time. We have many historical women in physics to look up to, like Lise Meitner or Emmy Noether, and yet often these women were denied resources and opportunities that their male colleagues had, which can make them feel like amazing but also tragic figures. I would hope that women working in science today can be heroic without the tragedy.

Perhaps looking for heroes in science is a fundamentally flawed endeavor. Science is at its heart collaborative, and the sheer scope of human knowledge means that it is impossible for one person, toiling alone, to conquer it all. We must talk to each other, work together, and build on existing work, as famously stated by Isaac Newton: “If I have seen further it is by standing on the shoulders of giants.” The great man theory is as flawed when it comes to science as it is when it comes to history. We all seek out role models, but we must recognize that they worked with others, seen and unseen, and that science is a societal effort and not the work of a lone genius.

While Feynman is long gone, there are other scientists still living, still contributing, and still behaving badly. It’s important that we not let them off the hook. Feynman lived decades ago, and certainly the standards of behaviour were different then, but today’s harassers and discriminators have no such excuse. If science is truly a collaborative effort, then it loses strength every time a person is pushed out of science by harassment. We can have complicated feelings about prominent scientists of the past, but there are a lot of people working in science today who are doing it right, and can serve as inspirations.

For example, tomorrow is the first ever LGBT STEM day, being celebrated with events around the world. Our Irish LGBT STEM network, House of STEM, has done so much to organise and promote this event, and founder Shaun O’Boyle explains why it’s desperately needed here:

The past is full of problematic yet successful scientists. Yet I’m hopeful that the future will have a broader array of amazing scientists, working together, who are also amazing people.

Threading a Nano-needle

One of the most exciting things about working in nanoscience is the incredible level of precision with which we can probe the world around us. This includes not just the physical world but also the biological world, which is a lot messier than most physicists are comfortable with!

There is currently intense interest in sequencing DNA via translocation through a nanopore, like threading a needle with the molecule that contains instructions for building life. Protein nanopores are the basis of new technology for DNA sequencing that made the news recently, costing only $1000 for full DNA sequencing with a palm-sized device from Oxford Nanopore. The protein nanopore is pictured in the image below, with DNA unspooling to pass through the pore.

But these nanopores can also be created in silicon, graphene or other thin materials. These solid state nanopores emulate the very small biological pores which can be found in the lipid bilayer around cells and their nuclei. Macroelectrodes in solution on either side of a solid state nanopore drive an ionic current in the carrier solution through the nanopore. As the DNA passes through the nanopore, it physically blocks the ionic current through the nanopore, allowing detection of translocation events. Additional electrodes can be added across the nanopore to enhance sensitivity to DNA. While research in this area is ongoing, it is thought that noise in the electrical signal through the nanopore can eventually be lowered—by applying coatings, slowing translocation speeds and improving fabrication techniques—to enable base pair sensitivity for DNA sequencing. Using solid state nanopores for sequencing could lead to more reliability and lower costs than protein sequencing, and is a major research area of the Drndić group at UPenn, the group I worked in for my PhD!

To me, nanopore sequencing is an amazing example of how direct electrical interaction with nanostructures can yield important information about not just the world around us, but our own place in it.

Cassini’s Grand Finale

After nearly 20 years in space and a final series of dives through Saturn’s ring, the spacecraft Cassini is on its last descent. It will crash into the planet later today, ending an incredible scientific mission to an amazing place.

Cassini and its instruments helped investigate Saturn’s atmosphere, its rocky rings, its strange polar hexagon. It also expanded our knowledge of Saturn’s moons, from the geysers and hidden oceans of Enceladus to the rocks and lakes on the surface of Titan. Cassini’s Huygens probe, which landed on Titan, was the furthest space landing of anything humanity has built.

Originally launched in 1997, Cassini’s mission was supposed to end in 2008. But it received two major mission extensions, nearly doubling its lifetime. It has sent 635 gigabytes of data back, which mightn’t sound like a lot except that all of it was on 1997 era technology, through a billion kilometers of space.

The spacecraft is being crashed into Saturn because it’s running out of fuel for orbital maneuvers, and scientists don’t want to risk accidentally crashing it into one of Saturn’s moons which might contain life, and contaminating them. I feel personally invested in Cassini’s mission and final resting place, because my signature is on board! When I was growing up, one of my Girl Scout troop leaders was an astrophysicist working on the Cassini Plasma Spectrometer before the spacecraft was launched. She arranged for us to all come in and get to see some of the instruments, and in the end our signatures were all added to the 600,000 signatures on the disc of human culture that was included with the spacecraft. This disc, modeled on the Golden Record aboard Voyager, is a time capsule of human culture for other spacefaring civilizations to find. But while the Voyager discs are traveling beyond our solar system, Cassini will be meeting its end here, to become part of the planet it studied.

Cassini’s voyage has been such an inspiration, a feat of technical and scientific exploration which I, along with millions of others, have loved watching from here on Earth.

You can read much more about Cassini’s scientific discoveries here, and watch Cassini’s final descent today starting at 7AM EDT here, on the NASA livestream. Godspeed, Cassini.

Sound and Waves

When I hear my mother’s voice, it sounds different from my father’s voice, and different from a bird or a drum. Why are the sounds we hear so varied, and how do they travel to our ears?

soundwaves

Sound is created when something moves rapidly, and creates a wave in the air around it. Our vocal cords do this, as does the skin on a drum, pushing the wave out into the world. This wave is made up of bands of air: more pressure, less pressure, high and low, back and forth as long as the sound lasts. Sound can only travel through something whose pressure can be changed, like air and water. So if you’re floating in space: perfect quiet.

But have you ever noticed how sound changes as it echoes around a gym? That’s because sound waves change when they bounce off things. A musical note will sound differently in a glass room than in one lined with velvet cushions. This affects musical instruments too! And the size of an instrument influences the sound it makes, from the deep growl of the tuba to the light chirp of a flute. Generally, bigger instruments make deeper sounds, with fewer waves per second.

And sound is not just high or low. Of course, it’s also soft or loud. But more interesting are differences that lead to a new tone or feel. For example, a violin and a flute might play the same note at the same volume, but they still won’t sound the same. Waves have amazing abilities to send subtle differences within a sound. And luckily for us, our ears use delicate hairs to detect these waves as they move through the air. Nerves connect the hairs to our brain, connecting us to the full orchestra of sound.