Frances Hellman

Notice: We are in the process of migrating Oral History Interview metadata to this new version of our website.

During this migration, the following fields associated with interviews may be incomplete: Institutions, Additional Persons, and Subjects. Our Browse Subjects feature is also affected by this migration.

We encourage researchers to utilize the full-text search on this page to navigate our oral histories or to use our catalog to locate oral history interviews by keyword.

Please contact [email protected] with any feedback.

ORAL HISTORIES
Frances Hellman

Photo courtesy of Frances Hellman

Interviewed by
David Zierler
Interview date
Location
Video conference
Usage Information and Disclaimer
Disclaimer text

This transcript may not be quoted, reproduced or redistributed in whole or in part by any means except with the written permission of the American Institute of Physics.

This transcript is based on a tape-recorded interview deposited at the Center for History of Physics of the American Institute of Physics. The AIP's interviews have generally been transcribed from tape, edited by the interviewer for clarity, and then further edited by the interviewee. If this interview is important to you, you should consult earlier versions of the transcript or listen to the original tape. For many interviews, the AIP retains substantial files with further information about the interviewee and the interview itself. Please contact us for information about accessing these materials.

Please bear in mind that: 1) This material is a transcript of the spoken word rather than a literary product; 2) An interview must be read with the awareness that different people's memories about an event will often differ, and that memories can change with time for many reasons including subsequent experiences, interactions with others, and one's feelings about an event. Disclaimer: This transcript was scanned from a typescript, introducing occasional spelling errors. The original typescript is available.

Preferred citation

In footnotes or endnotes please cite AIP interviews like this:

Interview of Frances Hellman by David Zierler on April 25, 2021,
Niels Bohr Library & Archives, American Institute of Physics,
College Park, MD USA,
www.aip.org/history-programs/niels-bohr-library/oral-histories/46802

For multiple citations, "AIP" is the preferred abbreviation for the location.

Abstract

Interview with Frances Hellman, professor of physics and of Materials Science and Engineering, Dean of Mathematical and Physical Sciences at UC Berkeley, as well as senior faculty scientist at Berkeley Lab.  Hellman is also president-elect of the APS. Hellman explains why she considers physics her “home” department and why her research agenda spans so many disciplines. She describes the major issues in her incoming leadership of APS and how Berkeley has coped during the pandemic. Hellman recounts her childhood in Manhattan and then Brooklyn and she describes her Quaker education and her early interests in science. She describes her focus on ski racing and her undergraduate experience at Dartmouth, and the formative influence that Bruce Pipes had on her development as a physicist. Hellman discusses her motivations to pursue thesis research at Stanford, where Mac Beasley and Ted Geballe were her co-advisors and where A15 superconductor research was in full gear. She describes her postdoctoral appointment at Bell Labs to work on magnetic thin film materials and magnetic superconductors. Hellman conveys her interest in entrepreneurship and the opportunities that allowed her to join the faculty at UC San Diego, and she describes building up her lab and her interests in thermal links. She reflects broadly on the basic and applied aspects of her research, and she explains her reasons for transferring to Berkeley and her affiliation with the Exploratorium. Hellman describes her administrative responsibilities as department chair in physics and she conveys her recent interests in amorphous materials and specifically ideal glass. At the end of the interview, Hellman discusses her involvement in both the APS and Berkeley’s efforts to make STEM more inclusive and diverse, and she describes her optimism that her work on amorphous materials will lead to key discovery in the field.

 

Transcript

Zierler:

This is David Zierler, oral historian for the American Institute of Physics. It is April 25th, 2021. I'm so happy to be here with Professor Frances Hellman. Frances, it’s good to see you. Thank you so much for joining me on this Sunday.

Hellman:

Oh, it’s totally my pleasure. It’s a slightly rainy day, which in California is a lovely thing, to have a rainy day.

Zierler:

Well, it’s sunny here in New Jersey, so I guess it’s backwards time for us today. Frances, to start, would you please tell me your title and institutional affiliation?

Hellman:

Okay, that sounds like a really straightforward question, but I do have a number of hats. I guess I will just launch into all of them, doing them in order that I think of them. I'm Frances Hellman. I'm a professor of physics at University of California, Berkeley. I am also a professor in the Materials Science and Engineering Department. And, I am the Dean of Mathematical and Physical Sciences at UC Berkeley. I'm also a senior faculty scientist at Lawrence Berkeley National Laboratory in the Material Sciences Division. And I'm currently the president-elect of the American Physical Society. I have various other things I do. I'm a visiting scientist—my technical title is scientist in residence - at the Exploratorium, which I get to as often as I possibly can. Although lately, my overcommitment level has been such that that has been less. But I love going to the Exploratorium. It’s one of the more fun things I do.

Zierler:

So now I know why we're doing this on a Sunday, because you obviously have way too much going on right now.

Hellman:

[laughs] I do have way too much going on right now.

Zierler:

In terms of your dual faculty appointments in each of the departments, where are you most likely to have graduate students and to teach undergraduate classes? Just to get a sense of the nature of your affiliations.

Hellman:

That’s a great question, and actually an important question, too. I consider myself first and foremost a professor of physics. Way back when I was applying for jobs, when I was a postdoc and looking for assistant professor positions, I applied for jobs in quite a range of departments, mostly physics but also some materials science, materials science and engineering, I think one chemistry department, and one or two electrical engineering departments. That’s because my research really could lie in any of these departments. I'm a condensed matter experimentalist and my research lies at the intersection of all these fields. I mostly do research in magnetic materials. But that being said, for the last quite a few years, most of my research has involved amorphous, meaning non-crystalline, materials. And in the last about ten years, I've been doing a lot of research on non-magnetic amorphous materials, specifically insulators. That’s its own topic and has led me to having an affiliation with LIGO, for example. I'm now an official member of the LIGO team, and I work on improving their mirror coatings, which are insulators.

So getting back to your question about affiliations, when I was applying for jobs, I reached the decision that I loved teaching physics. And that’s still true. I probably could be in any of these other departments, but I love teaching physics, and I love having physics as a hobby as well as my job. I love learning about all aspects of physics. The breadth of physics, I just find remarkable. So I love being aware of that. I love going to colloquia. And since I love teaching and learning about physics, the natural thing for me is to be in a physics department.

That being said, my graduate students come from a range of departments. It is common that I have graduate students from materials science and engineering—MSE, to use the abbreviation. It’s common for me to have about six graduate students, and of them, right now I have three from physics, two from MSE, and one from electrical engineering. And that’s pretty common. At other times in my career, not that long ago, I had two from chemistry.

Students with a physics background are a more natural fit for me. Their background is a little closer to what I need them to know and be able to understand. But that being said, people in materials science and engineering or electrical engineering often have a physics undergrad degree, and they bring a breadth of knowledge in other areas that is extremely valuable to the group. I guess I first and foremost would call myself a physicist, perhaps a materials physicist.

Zierler:

Do you have strong feelings about the terms “condensed matter” versus “solid state”?

Hellman:

Nope. No strong feelings about that, at all. Both are good! [laughs]

Zierler:

[laughs]

Hellman:

I was already a physicist when that transition happened, which was in the—oh, gosh—1960s, 1970s, something like that.

Zierler:

It depends who you ask, really.

Hellman:

Yes, I believe that. So no, I don’t. I've never worked on superfluid helium, which is one of the classic areas that led to the distinction of not wanting to call it solid state physics. My work on amorphous materials is related to the liquid state, or is related to the liquid state which then condenses, goes through a glass transition, into the solid state. Which is part of the reason we call it condensed matter physics and not solid state physics.

Zierler:

And where is soft matter physics in this for you?

Hellman:

Great question. I get invited to soft matter conferences to speak on my work on amorphous materials and ideal glasses and topics like that. Many of my talks are in sessions on soft matter, and many of the people that I talk with, particularly the theorists that I talk with about the ideas around ideal glasses and related, call themselves soft condensed matter scientists.

I don’t tend to call myself a soft condensed matter person, even though I'm working on non-crystalline materials. And that’s in part because one of the aspects I'm most interested in in amorphous materials is what one might term a hard-core condensed matter topics, meaning correlated electron physics. I love the fact that I observe in amorphous materials what most people regard as properties that you only find in crystalline materials, and I love showing that the theory underpinning the understanding of those properties are therefore an incomplete understanding. Amorphous materials don’t have Bloch waves and we lose our mathematical approach to the condensed state, represented by Bloch waves and periodicity — yet we have virtually all the properties that we are trying to describe with Bloch waves. I have shown that some fairly hard-core properties, like topological insulators and Berry phase curvature, which are topics whose theories rely on the periodic lattice and reciprocal space, are found in amorphous materials.

The thing you don’t have in an amorphous system is reciprocal space; k is no longer a good quantum number. At least as a vector, it’s no longer a good quantum number. Energy E is a perfectly good quantum number, but momentum k as a vector is not. But you find the same phenomena. What I love about that is that you're forced to think about these phenomena slightly differently. You're forced to think about them in real space, for example. You have to think about real space analogs to what you normally talk about in reciprocal space.

We're getting a little far afield, perhaps, but the example I love to use for this is superconductivity, which in the original BCS theory relied on plus-minus k-state pairing. That original formulation of BCS theory is completely a reciprocal space description. [An] electron traveling one way pairs with an electron traveling the other way. And many people don’t know this, but there are amorphous superconductors. There are amorphous superconductors that have significantly higher TC than their crystalline counterpart, molybdenum being the best example that I know of, which is around one Kelvin in its crystalline state, and around seven Kelvin in its amorphous state. This forced a generalization of the superconductivity theory. It was no longer plus-minus k- state pairing; it’s time reversed states. Leading to this beautiful description in which one electron travels through scattering essentially every mean-free path. And a time-reverse of itself travels back through that path, just like light would, through a bunch of scattering events. No matter how complicated the mirrors—as long as there’s no inelastic event. Those time-reserved states are what actually pair.

And that may sound like a purely mathematical generalization of BCS theory, but it leads to very deep insight. Superconductivity is a remarkable phenomenon. Many refer to it as an emergent phenomenon. It doesn't care about the details of the underlying structure; the coherent electron state emerges, as a pure wave function. Isn’t it more amazing that some material (like amorphous Molybdenum) superconducts where in its normal state the electrons scatter at every atom—meaning that the mean-free path is like one and a half angstroms in the normal state, and when it goes superconducting, it goes into a state where the paired electrons don't scatter at all—well, the specifics of what they are doing, you have to think about a bit, but the pairs travel coherently in a single wave function.

So I would argue that this is an even more dramatic example of an emergent phenomena than superconductivity in crystals. I love that example of the generalization necessitated by thinking about amorphous materials, and I'm constantly, in my work, looking for things like that. Like, can I generalize, where people think they understand the theory, but they're missing a basic point? That was a long-winded answer!

Zierler:

No, that’s great. Where is LBNL in all of this? Is that more of a courtesy appointment, or does your work at the lab really give you access to instrumentation, funding, collaboration, that you wouldn't have just being a university professor?

Hellman:

The latter. LBNL is key to my work. It’s key to the funding of my work, but it’s also key to my work more profoundly. I think of my work as divided into two parts. I used to work in superconductivity, when I was a graduate student. But I turned to magnetism fairly shortly after that. And my LBNL work is all about magnetism. It’s all about spintronics and magnetism, particularly magnetism in the construct of heterostructures, where there are interfaces and strong spin orbit coupling, which leads you to these interesting spintronic effects, where charge and transport get very coupled together.

I have been, for the last, oh, maybe five years, the leader of a group of scientists, a mixture of Berkeley campus and LBNL scientists, seven of us, working in the general area we call non-equilibrium magnetism. LBNL has brought me these collaborators—which has expanded the work I do. Scientists who are expert in scattering techniques, or in microscopy, or in ultrafast pump probe techniques, or in spin orbit torque. And it has brought me funding to do this work. And facilities. It’s tied me in more to the Advanced Light Source at LBNL. Three of the people in our magnetism group work at the Advanced Light Source. LBNL has brought me all those things, including a style of work, to work as a team, as opposed to an individual.

Earlier in my career at Berkeley, there was an opportunity to either split away and apply for an individual grant to DOE, or I could choose to be part of LBNL. You can’t both be DOE-funded as an individual university grant person, and be in LBNL getting money through them. I thought about it for a while, and I realized LBNL is way more than just funding. It’s a key to that collaborative type of research. So I chose to be part of LBNL.

Zierler:

Just as a snapshot in time, in light of your incoming leadership at APS, what are some of the big items for discussion right now? What are some of the big areas that the physics community, as represented by APS, is dealing with front and center circa Spring 2021?

Hellman:

That is a great question, to which I have a partial answer. We discussed a little earlier—my overcommitment level right now is a little extreme. I have a very active research group and I am the dean of Mathematical and Physical Sciences.

Zierler:

The more important point there, Frances, is that you are a scientist who happens to be dean. That’s how you're organizing these things in your mind.

Hellman:

Yes, that’s true. And, I'm the APS president-elect. So these three things—well, I didn't intend to do all three of these things at the same time. I've been involved with the APS in various ways for a long time. I was Physics chair for six years, about ten years ago, and then I've been dean for about six-plus years now. And I've always been adamant that I would keep my research going. I also do a little bit of teaching. As Dean, I'm not asked to do any teaching, but it’s important to me. So I do a little bit of undergraduate teaching, in addition to supervising/working with my research group. I do a lot of teaching through the mentoring of my students, undergrad and grad, and so forth. But I also do a little bit of actual classroom-type teaching, in the form of a freshman seminar on magnetism which I taught nearly every semester that I have been dean.

So [laughs] wearing all these hats is actually one too many, and came about as an accidental evolution where I got asked to do one before I had been asked to do the other. I agreed to run for office for the APS when I had not yet been asked to re-up as dean. And then I couldn't see not running for APS office when I might not be asked to be dean. That wasn’t an obvious thing, that I was going to be asked to be dean again. So I agreed to run for APS office. Then before the election happened, I was asked to become dean again, for another five years. And then the challenge is, do I turn that down because I might win the APS election? I mean, that doesn't make a lot of sense, either. I thought, “Well, I don’t know that I'm going to win the election.” So I agreed to be dean for another five years. And then I did win the election.

And then I had a moment where I thought, “Well, maybe I could just do everything. I've managed to do a lot of things in my past..” And then COVID happened. And that multiplied almost everything. It did reduce my APS job a bit, because I'm not traveling as much as I would have been. But my dean job went up tenfold. It has just been overwhelming. That was a longwinded way of saying I am stepping down as dean at the end of June, so that I'll have a lot more time to invest into my APS responsibilities.

What are the priorities for APS? Well, the first thing is, I don’t think anybody knows what the post-COVID world looks like. I think everybody is trying to figure out what are the positive things that we can take from all our Zoom calls and our ability to talk to people all over the world. It’s obvious the things we've lost. We've lost those hallway meetings, the casual conversations, the informal sorts of settings. I think everybody is craving—literally craving—getting back together again, in-person meetings. But, there are probably some things to be learned out of how we ran our meetings. The same thing is true of how we run our classes. What do we learn, what positives can we take from this, as we move back into a post-COVID world? The topic of antiracism and now merging as well with the topic of anti-Asian racism specifically—

Zierler:

As if there weren’t enough challenges to deal with!

Hellman:

Yeah, well, we should have seen that coming. We've all been a bit—I think we've just all been a bit oblivious. I really think it has been there all along; we just were oblivious. Now we've become very conscious of the anti-Black racism. We've been very conscious of the anti-women misogyny, or just sexism; it’s not even necessarily misogyny. I think the Black Lives Matter movement and the very public deaths has just brought it all to the forefront. I think we're all starting to recognize, it’s just not okay. It’s really not okay that we still have such low representation of women and even lower of people in color in physics. It’s just not okay.

And to pass it off as, “Oh, well they don’t really want to do this” is completely missing the point. Yeah, many don’t want—“they” don’t want to do physics the way it’s currently constructed. But we have to take better ownership of what that sentence actually says. They don’t want to do this because we make it very unpleasant for them to do this. Our Black colleagues face just crazy things, of people thinking, for example, “Who is this trying to get into the building?” and asking them for ID, where they never ask a white person for ID. I see that as a huge issue facing our community. And a bit of a generational divide, where the older generation of scientists seem to really believe they are not racist. And they're not getting that that’s just not true—they are racist; they just don’t recognize that they are. They're not racist in the sense of they're not out lynching people, or deliberately excluding people. But there are implicit things that are happening all over the place.

Zierler:

It’s also perhaps a lack of acknowledging that there’s systemic racism.

Hellman:

Exactly. That term systemic—I've had arguments with people, where they say, “There’s no such thing. That doesn't mean anything.” In my opinion, “Well, yeah, there really is such a thing, and it does mean something.” So that’s a huge issue. I think there’s also a big issue facing our country—perhaps the world, but let’s just stick to our country—that some anti-science sentiment has emerged . Scientists have a belief that we are right, and that we are smarter than everybody else in the room, that then doesn't translate well into making people work with us. So we have to change how we present ourselves, and how we interact with people.

Us just talking at people is not the right answer. We may know the answer, but it’s like when you teach a class. I always say to students—you can ask me how to solve a problem, but the problem with that plan is my showing you how to solve the problem is not going to actually help you solve the problem. It’s a different process that has to happen. I have to help you get unstuck. I have to help you find solutions such that you actually know how to solve the problem, not just know how I solve the problem.

We have to do a better job as scientists at helping people discover their own love of science, their own curiosity—I refer to this as “the scientist in everybody.” I believe there is a scientist in everybody. We often talk about the poet in everybody, or the artist in everybody. I also believe there’s a scientist in everybody. It may not look exactly like scientists that we currently call scientists. They may not have the math background as an example. But they are scientists. Kids are curious. We have to get people back to appreciating that they too are scientists, and that what scientists do is of value.

Zierler:

And there’s a duality of value in that as well, because not only does it enhance the relationship between scientists and the broader public, but it also enhances the movement of inclusivity.

Hellman:

It does.

Zierler:

The idea that there isn’t an other-ism at play here.

Hellman:

Correct, absolutely.

Zierler:

That everybody has their inner scientist, and that we need more people who might not think of themselves as scientists, to think about science as a career for themselves.

Hellman:

Yep, exactly. And there’s a range of scientist careers. Some of us are researchers at R1 universities. Others of us teach high school science classes. Others are science advisors to like politicians, or are lobbyists on behalf of science, or patent lawyers. All of industry is filled with scientists that are needed for technological advances. There’s a huge range of careers. I'm a very strong believer that it’s not a hierarchy. It’s not like the people who are academics at the top universities are somehow the top of some hierarchy. All these people have important strengths and are valuable to our society.

I am, in some ways, very smart, and I have a good memory for things, and I have a good knack for working out how to do things. But there’s other ways in which I am not as smart as other people —my husband Warren, for example, is much smarter than me about various things. And it’s quite funny when I suddenly realize this. For example, I don’t recognize people, for example, very well. And he’s like, “We just saw them yesterday.” Or, “We just heard them on the radio yesterday. How do you not remember that?” That’s not my skill set, apparently. People have different talents, and we have to learn to value them.

Zierler:

Frances, I'll ask one last question that’s in the moment, then we'll go back and develop your personal narrative. In light of your deanship coming to an end in late June—and as you say, of course, COVID happened and that overtook everyone—obviously there was no playbook for what we all dealt with over this past, what are we, now, fourteen months? Something like that. What do you wish you knew now at the beginning of this pandemic, and how would you grade yourself in terms of your part in guiding Berkeley as best as possible through this multifaceted emergency?

Hellman:

All right, I'm going to answer this in two parts. I will talk about the Berkeley part first. One of the biggest challenges in COVID was not knowing how long it would last and what we were going to have to do. So we spent so much time creating plans that never got implemented. Just constantly, constantly creating plans for—“Oh, maybe we'll be able to start in-person classes at the start of Fall semester.” Nope, that’s not going to happen. “But maybe we can have in-person classes starting at this other date.” Constantly planning for things that didn't happen, so that everybody was just working much too hard. And I don’t know if we could have let that go, but in hindsight I wish we had.

To tell you a related story from when I moved to Berkeley—I was a professor at UC San Diego from like 1987 until 2004, and I moved to Berkeley in 2004. And when I moved, my laboratory wasn’t ready, but I had a research group. I said, “While my lab gets done, I have to have a laboratory.” So they gave me this little room, down on the basement floor. And I picked two or three experiments that I wanted my group to be able to do, while my lab got built. And we worked so hard. My whole group, we worked so hard, in that little tiny space, which had not enough ventilation, so the pumps kept turning themselves off because they were overheating. And in the end, we did do one experiment. We measured the magnetoresistance of amorphous Gadolinium-Germanium. It was a key point for a paper that we finished. And I was so excited about it. We did it in this horrible, little tiny space. And I put the data on my office door—“We did this!” We spent almost a year, and we got one experiment done. In hindsight, perhaps I should have just sent everybody home. I should have told them to just take a vacation. Just take a break! Do some reading, so something. But I was so determined that we were not going to shut down for the year.

I felt like this time of COVID has the same element. We were so determined to not shut down. But we were all running around a lot. And I don’t know that we could ever have done this differently, because I don’t think it was ever going to be possible to just say, “Oh, well, we're just going to go home and—we'll just pay you all somehow, but we're just going to wait to see when we can reopen.” There’s a part of me that wishes we could have just waited to see when we could reopen. I don’t know if that’s really a lesson learned. I don’t think we could have done it differently, but it was painful, how much effort got spent, with little effect.

Personally, I remember the day we got sent home. I was in my office at Berkeley. I think I walked home. It’s about a 40-minute walk. My husband was off doing his job. And I was mildly excited about the whole thing. I had all these plans. I was already busy thinking through what projects I was going to give to each of the students in my research group to keep them productive. I had a plan for every student. We're setting up a new scanning calorimeter, and so a couple of them were going to do simulations of the thermodynamics of the nano calorimeters that we needed. Others were going to do density functional theory calculations of electronic density of states in amorphous topological insulators, a topic I'm interested in, etc.

I had all these plans for what each of the students was going to work on, and I was slightly excited because I thought of all the things I was going to do. Like for example [laughs] I remember I was going to clean up the house. I was going to finally get rid of all those extra clothes in my closets. I was going to learn the mandolin. I am a member of a family band, and I don’t play an instrument; I am vocals only. But I have been under a certain amount of pressure for quite a while to learn an instrument. And so I have a mandolin that I borrowed from somebody. And for the past about three years now, I have been paying $20 a month to an online mandolin teaching class, but I have not participated in a single class. Not even one. I have literally not even logged in once. But I every month pay my $20 dutifully.

So as I was walking home, I'm like, “I can finally do this.” I was on my way home, and I had all these plans of what I was going to do. And here we are, 14 months later; I still have not done a single mandolin lesson. The house is definitely not cleaner. So what lesson would I personally take from that? I guess maybe give myself a little more of a break? I don’t think any of us knew how long this would go on. I remember being all excited that I had a freezer that had two pizzas in it. I thought, “Oh, that’s great. I'm all set for food for a little while, because I've got two pizzas in the freezer.” And of course, that’s not even a month. The things you think at the beginning are going to be helpful to you—we've learned all kinds of things, how to do all sorts of things that we didn't know how to do when we started.

Zierler:

I'll just share with you that at the beginning, I was going through the same thing, trying to figure out what I was going to do now that I'm not traveling and interviewing scientists in person. I reached out to top-ten department chairs in physics, asking, “Who’s emeritus that would be great for me to interview on Zoom? I've never done that before. Let me do it now. That’ll keep me busy for a couple weeks.” One of them was—I got in touch with Wick Haxton, and as a result, John Clarke, Mary Gaillard, and Marvin Cohen. And I thought, “Wow, this is amazing.” And then, here we are; I'm just doing everybody now. [laughs] Frances, let’s take it all the way back to the beginning. First, let’s start with your parents. I know this is a very interesting story. Please tell me about them and where they're from.

Hellman:

Oh, gosh. Well, they are interesting. There is no obvious science in my past, in my heritage. My dad was born and raised, more or less, in San Francisco. His parents were interesting. His mother in particular—Ruth, my paternal grandmother—was a WAC, Women’s Auxiliary Air Corp. They flew planes in World War II. They mostly flew supply planes, but then I've also heard stories of them flying target practice. They would haul the banner behind that the other planes would shoot at. And she had amazing stories—like they’d be shuttling planes from a factory across the country. In those days, planes did not fly all the way across the country. So they would land somewhere, and the military commanding officer would say, “What the heck is going on here? Women don’t fly planes.” And he wouldn't let them take off! [laughs] And they’d have to call Washington to get permission for them to refuel the planes and take off again. She was an amazing woman. She died when I was relatively young, scuba diving, when she was about 60. She was remarkable.

My dad was a native San Franciscan, and my mom was British. She grew up in England, a wartime child, World War II child. She was a dancer, so she stopped high school when she was 16, after she took the A-levels. Which you take when you're 16 in Britain, to enable you to go to the university. Her father would not let her stop school to become a dancer until she passed her A-levels, which she did, then she stopped school and danced professionally. She was a soloist with the London Festival Ballet. She and my dad met on a boat, transiting the Atlantic. My dad went to UC Berkeley, and he was doing a junior year trip abroad with his buddies, and so they were on a boat traveling to England, and from there to Germany and so forth. And my mom was returning from a set of ballet performances in the U.S. And so they met on this boat.

And over the course of the next two or three years, they saw each other twice [laughs] for like a week each time. She went to San Francisco one time on tour, and she stayed with his parents. And he went one more time to London and more or less courted her. This was over the course of three years. They had about two weeks together, and then he sent her a telegram saying, “Will you be Mrs. Hellman?” And she said, “Yes.” And so they got married, then she flew across to the United States. And in those days, there were phones, but it’s not like you routinely went home at that point.

My father was very non-technical, a very successful businessman. A very entrepreneurial kind of businessman. He believed in investing in people. He would find somebody he believed in, and then invest in them. He did well by himself and by them. Nobody on that side of the family has any technical background—many are bankers, and there’s some lawyers in there, things like that.

So I've always been of the belief that it was my mother who had the innate science skills, but never expressed, because she left high school at 16. She always believed herself to be terrible at math, but I just don’t believe that.

I'm the oldest of four, there’s four kids in my family. I was born in Britain, so presumably could never be president. Although I've never been clear on that, because I was born of U.S. citizens while my father was in the military. But I have no interest anyway in being the president of the United States. We should be clear on that.

Zierler:

Well, you say that now. Let’s see after APS. Maybe you'll find out.

Hellman:

Yeah, yeah. [laughs] My dad was in the military, so it seems like a little funny to think that if he’s in the US military stationed in Germany—

Zierler:

What does your birth certificate say?

Hellman:

It’s a British birth certificate. It says I was born in a British hospital—because my mother went home to England to have me. But we went back to Germany, to the army base Dad was stationed with by the time I was 6 weeks old. I have a giant smallpox vaccine on my arm—huge—like the size of a quarter—because my smallpox vaccine was done in a U.S. Army base, and they don’t routinely do children. And then each of my sibs was born as my parents moved around. My dad went to business school at Harvard, and so my sister was born in Boston. And then the next two siblings were born in New York, where he then had a large fraction of his career.

Yeah, so I was the oldest. I was a very rebellious teenager. It was extremely important to me to establish independence from my parents. I was very immersed in ski racing. When I was in high school, the most important thing to me was my ski racing, not my school. I applied to colleges entirely based on their ranking in the ski world. I applied to Dartmouth College, Middlebury College, University of Vermont, and Johnson State. I had this idea of how hard they were to get into. This sort of ranking in my head, whether accurate or not. But they were all chosen because they were ski racing schools.

Zierler:

Was science on your radar at all in choosing schools?

Hellman:

Yes, it was. It absolutely was, but as sort of a secondary feature. In high school, we were living in New York, and we all got very involved in ski racing. And it is pretty much impossible to competitively ski-race while living in New York City. There’s only so far you can get commuting every weekend up to Vermont.

Zierler:

And what neighborhood? Where did you live in New York?

Hellman:

When we were young, the very first place was on East 85th Street, Upper East Side, in a tiny apartment. I was old enough to remember this tiny apartment. Then we moved to Brooklyn, and we moved into a house in Brooklyn.

Zierler:

Where in Brooklyn?

Hellman:

Brooklyn Heights, on State Street in Brooklyn Heights. I'm not sure why I still remember that. I went to Brooklyn Friends School, a Quaker school. I've sometimes wondered why I never asked my parents how it ended up that we went to a Quaker school, but we did.

Zierler:

What’s your parents’ religious backgrounds?

Hellman:

My dad’s Jewish, was Jewish. My mom was—well, more nuanced—my mom in some sense might have been Church of England, Presbyterian or something like that. But her father was a member of the British Communist Party, so he was avowedly anti-religion. I remember him quite well. He was this lovely, sweet man, who every time we’d visit—about every two years, we’d travel to England, and my sister and I, being the oldest— got to stay at Grandma and Grandpa’s house, and everybody else stayed in a hotel. And every time, we would stay there and paint the shed in the backyard of my grandparents’ house. They had built this house right before the war. It’s a street with all similar houses. It was amazing to me that they built a house, just before the war, and then lived there, the whole rest of their lives.

My sister and I stayed in my mother’s old room. There was a garden, with a shed, and we would paint the shed every visit. And I remember it being this big production to paint the shed. And of course I've seen the shed since then, and it’s tiny! It’s this little tiny shed! It’s just big enough to hold your tools. You can walk into it; there is an inside space. But I had this image of it being this huge thing. So we would paint the shed every year, and have kippers and tea with my grandma. The English drink more tea than anybody can imagine. Life with my grandparents was a constant stream—some coffee mixed in there, but mostly tea. Like 20 cups a day kind of amounts.

Zierler:

[laughs]

Hellman:

We would go to a park, and we would have tea before we left the house, and then they would bring a thermos, so that when we got to the park, we could have another cup of tea. Where was I going with this? Oh, science!

Zierler:

Well, I wanted to ask, on the religious front, with your parents being an interfaith union, was this a like menorah and Christmas tree kind of December situation, or not even that?

Hellman:

It was messy. Their wedding, in the 50’s in Britain, was complicated to arrange due to the interfaith yet pretty non-religious nature of it. My dad’s mother and father were Jewish, and semi-practicing, I guess. They went to High Holy Days, but it wasn’t a deep part of their lives. When I was growing up, we spent every Christmas with those grandparents, with a Christmas tree, and Christmas presents. Judaism was not front and center in their lives. Maybe I'll tell you a funny story about that. We did not grow up being Jewish, my sibs and I. My dad was Jewish by cultural heritage but not strongly Jewish in other ways. At one point, growing up in New York City, my parents decided that they needed to get us out of the house on Sunday mornings. So they sent us to Sunday school! [laughs] Which seemed completely fine to everybody. Nobody thought much about it. Until one time, my grandfather, my Jewish grandfather, was visiting— I'm about seven or eight at this point—he asked me what I would like for my Christmas present. Again, the irony—my Jewish grandfather is asking me what I’d like for my Christmas present. And my answer was I wanted a statue of the Virgin Mary with a little baby Jesus in her arms.

Zierler:

[laughs]

Hellman:

And that crossed a threshold for everybody.

Zierler:

I was waiting where the red line was, and there it is.

Hellman:

That was the red line. That crossed the threshold for everyone. Everyone was like, “That is just a bit too much!”

Zierler:

[laughs]

Hellman:

So we got taken out of that Sunday school, and then got sent to a Unitarian Sunday school.

Zierler:

That sounds about right.

Hellman:

I remember being incredibly excited about it, because they gave me a Bible with my name engraved on the front in gold letters, which was very cool. And we went to this Quaker school. So religion was a funny mixture. And, my maternal grandfather was a Communist. So he did not believe in any of this. In the 1960s, he saw the flaws of Russia, but not very clearly. I remember he had a Russian camera that was constantly failing. It was always a source of jokes in our family, like he’d go to take pictures, and the shutter wouldn't close. And he and my father would have these endless debates, my father being rather a staunch capitalist, and my grandfather a communist. Very polite, but debates. A long tradition in the family. He was never allowed to come to the U.S. Being a member of a foreign Communist Party, you were not allowed to come to the U.S. And so he never came to visit. So we grew up in a very areligious household. And I and my sibs are technically not Jewish, because our mother is not Jewish.

Zierler:

Well, that’s by the Orthodox standard. There are other ways of looking at this.

Hellman:

Correct.

Zierler:

You need not define yourself by one strand of Judaism, so you choose.

Hellman:

Exactly. And in Hitler’s Germany, you didn't have much of a choice. So there’s no question in Hitler’s Germany where I would have been. My sister, the sister just younger than me, actually chose to convert, in a formal way. She went through an adult version of a Bat Mitzvah, with my dad who also went through the ceremony.

Zierler:

How adorable.

Hellman:

Judaism became increasingly important to him in later years. It was very, very sweet, the two of them together, doing this ceremony. And it did become important to him. Although in a kind of secular way.

Zierler:

Frances, growing up on the Upper East Side is going to warp your perceptions of class. But what were your perceptions? Did you have a sense that your father was very successful and that translated to an upper-class childhood, or not necessarily?

Hellman:

Wow, you're asking a remarkable question, because I spent pretty much my entire adult life denying that I came from privilege. My first job was when I was 11. I earned $100 from a job I held when I was 11. And I remember I was so excited by that, that it was my money, that I had earned my money, no one else’s, and at 11. And that was true all the way through. I was always very adamant that I had to earn my own money, and I didn't want my parents’ money, and it was very much a point of pride for me that I was going to make it on my own.

It’s an interesting thing to look back on now. That whole thing with my sequence of ski-racing colleges—I got into Dartmouth, so I went to Dartmouth. But I was quite adamant when I went to Dartmouth that I was not going to take money from my parents. Now, the privilege part of this, which I've come to recognize since—it’s kind of a privileged statement—I did not, in fact, take money from them. They never sent me a check. But they did pay my tuition, and they did pay my room the times that I had rooms in the university. But I always had jobs, and whenever I was off-campus, then I would pay my own way. And I always paid all my own clothes, all my own books, all my own optional things.

I had jobs the whole way through college. I worked at Dunkin Donuts for a quite a while, which was one of my more interesting non-technical kinds of jobs. So I didn't recognize my privilege—I look back on it now and I say, “Wow, I had privilege that I refused to admit.” I was determined that I was going to make it on my own, and I did not need my parents’ help. Growing up on the Upper East Side and going to a private girls’ school —there’s a lot of angst going on now in that school about what that meant, and how terrible we all were around the very few people of color who were in that school.

There’s a way in which, when I was growing up, having money was viewed as a bad thing. So I very much incorporated that into the way I thought of myself and my family. Having money, meant you weren’t making it on your own. And I always thought it was quite ironic that the people who would say this —in modern language, the one-percenters— first of all usually don’t recognize they're one-percenters. Secondly, half the people who were protesting against the one percenters actually are in the one percent; they just don’t want to admit it. That was very clear to me even when I was a child, that the people who were being dismissive —“You, oh, your father has money”—they were very privileged, also. So it was complex, growing up. And it did lead me to feel very strongly that I wanted to make it on my own. So a complex question you've asked there.

Zierler:

Now, when you got to Dartmouth, is it a gradual process whereby science becomes more important to you than skiing?

Hellman:

Science actually became important to me while I was still in high school. I had an interesting and kind of weird high school experience, mainly because of the ski racing. Ski racing was incredibly important to me. I was attending one of these private girls’ schools in NYC, and there was a moment in tenth grade where I qualified for the Junior Nationals, which is quite competitive. And it happened that the Junior Nationals—or maybe it was the Junior Easterns, which is the predecessor to the Junior Nationals—was during exam week, or perhaps during exam preparation week.

I was at that point doing this hybrid thing. I was attending school in New York City, but I would be gone to Vermont from October through April. My parents and one other family had started this ski racing academy, which was a tutoring system, basically. There were around 15 kids, and we lived in these two houses; a girls’ house and a boys’ house. And there were two tutors, one science and math, and one arts, humanities, and social sciences. We would spend every morning studying, and then every afternoon training for ski racing. And we would do that from October through April.

Then there was this moment where I qualified, and so we said to my school in NYC, “She needs to stay for the Junior Easterns.” And they said, “Well, she can’t.” And my parents said, “What do you mean, she can’t?” They said, “She has to come back. If she doesn't come back for prep and exam weeks, she can’t come back at all.” And we all looked at each other and said “I can’t go back. I've just qualified for the Junior Easterns. That doesn't make any sense at all!” And to my parents’ credit, they said, “She’s not coming back.” So we had this little moment of playing chicken with this whole thing. And they said, “Then don’t come back at all.” And so we said, “Okay.”

And so there was a brief moment where I was enrolled in like a public high school in Vermont. And then one of the other private girls’ schools in New York—where my sister went—reached out and said, “We’d be happy to have her come to us, and we will accommodate her schedule.” The whole thing was sort of silly. The first school kept saying, “If we let you not come back, we’d have to let everybody not come back.” And I'm thinking, “That doesn't even make sense. Not everybody has qualified for the Junior Nationals”. I was doing well in my classes. I thought, “I could understand this if there was a problem, but there isn’t a problem. There’s just your rules, is the problem.”

I first transferred to the school in Vermont, where I nominally finished tenth grade—ironically, because I never actually went there, so it was all very theoretical. Then this other school in New York said, “We will happily take you. We will work with you. We get it. As long as you're doing well in your classes, we will work with you. If you start doing badly, that’s different. But as long as you're doing well, we will work with you.” And so I transferred to this other school.

And then a strange thing happened. That school was in the middle of changing when they did physics. So in the end, I had, in 11th grade, a high school physics class, where there was only five of us in the class. These private schools, they're small anyway, but five was really small. And this physics class was taught by a woman who was getting her master’s degree from NYU. I always thought she was getting her master’s degree in physics, but I learned years later, she was actually getting her master’s degree in biology.

So she was teaching our class, and she taught this amazing physics class. Amazing in the sense that I learned nothing that one normally considers introductory physics, but we learned so much about black holes, and timelines, and waves. We learned all these amazing concepts. And I remember literally at one point figuring out, or believing that I had figured out, how the universe began. That the universe had begun with a forward-moving universe and a backward-moving anti-universe, spontaneously created out of the vacuum. And I was all excited, and I went to her and said, “I think I've figured out how the universe began.”

Zierler:

[laughs]

Hellman:

And she just let me go with that. We explored it. We talked about it. And meantime, I was also going to the school in Vermont with the tutors, and the math science tutor there was incredibly excited he had an 11th grader with all these ideas. So I basically got all this one-on-one just amazing physics. I loved it. So I went to college totally convinced that I wanted to do physics, having loved high school physics.

My senior year in high school was academically a mess. The tutoring approach became a full-time school, so it was actually an accredited school for my senior year. Which mostly resulted in my doing absolutely nothing scholastic my senior year. I ducked everything I could, and spent my time focused on ski racing. I almost didn't pass. There wasn’t much standards being set in this first year while they figured things out, but there was some. I did have to pass a chemistry test, which since I had never even opened the chemistry book, was a significant challenge. And I did have to pass it. Fortunately, they gave me three tries to pass it, and by the third time, I had figured out the answers, so that I did pass this chemistry class! So my senior year was an unusual one and not very academically focused.

Anyway, the next year I went to Dartmouth intending to be a physics major, loving physics. I just thought it was amazing. I took a placement test for physics at Dartmouth. Dartmouth is a liberal arts college, so you don’t take tons of classes in your major. But there was an advanced physics sequence, which was called Physics 18/19, and then there was the regular physics. The regular physics sequence didn't start until the winter, but Physics 18/19 started in the fall. And I knew I wanted to ski race in the winter. So I wanted to get into Physics 18/19, not because I had had any particularly advanced physics training, but because I wanted to start physics in the fall, not wait until winter. So I took this placement test.

Which was interesting, because I really had had none of the traditional physics. The person who was teaching the advanced class called me in, and he has my exam in front of him, the placement exam. And he said, “You know, I'm trying to decide what to do with this. Because you didn't do particularly well on the questions, but you seem to have a lot of insight, and I feel like you inherently know a lot. But you didn't do particularly well. So it’s on the edge, and I wanted to talk to you about what you wanted to do.” So I told him I really wanted to take Physics 18. So he said, “Okay.”

This, by the way, was a man named Bruce Pipes, who was a professor at Dartmouth. And he said, “Okay, look, your background is really weak, so I will help you. You can have as many office hours as you need. I will help you get through Physics 18/19.” So I walk into class the first day for Physics 18, and he says, “The first eight chapters of this book should be refreshers for all of you, so I'm going to start directly with chapter nine.”

Zierler:

Ooh.

Hellman:

I think chapter nine might have started with a spring. And so the equation was F = -kx. And I looked at it, and I didn't even know what he meant. I mean, I remember thinking, “How is that an equation?” The only equations I had ever seen were math equations. I went to his office hours, and I said, “I don’t understand. You can’t have three unknowns in one equation. That’s impossible. If you have three unknowns, you can’t solve anything. You can only have one unknown in one equation.” [laughs] In my world of math, which is the only equations I had ever seen, you couldn't have F = -kx. You'd have to have 37 = 28x2 or something like that. You could only have one variable. I just so thoroughly did not understand this. But, he was fantastic and patient and helped me through that first Physics course, and later became my advisor.

Zierler:

Frances, is your sense that—I mean, since you were objectively not prepared for this level of physics, was Pipes particularly encouraging of you as a young woman? Was that part of the equation as well?

Hellman:

I don’t know. He told me he sensed that I was going to be really good at physics. He sensed that I had an intuition and an understanding of physics despite clearly, the background wasn’t there. So I don’t know to what extent being a woman played into that.

Zierler:

How many of your fellow classmates were women at that early stage?

Hellman:

There were I think four, out of 40, which is quite a large number for those days. This is 1974. Bruce Pipes was personally responsible for a fair number of young women physicists coming out of Dartmouth. This is also the days of co-education just starting. I was the third co-ed class at Dartmouth, which was its own story, and quite horrendous in some ways, and bizarre in some ways.

I've had a few places in my career where the only reason I'm still in physics is some person like Bruce Pipes intervened and helped. He was amazing. I was in his office every week with homework problems that I didn't understand. There was so much I didn't understand. I remember one of our very first assignments was on a computer, to create an orbit of a satellite around the Earth. And I didn't even understand what this meant. I thought computers did things like solve for prime numbers. I knew that for a satellite going around the Earth, there was an equation—F=-Gm1m2/r2. That was fine. I knew that. But I didn't have any concept of what it meant to put this into a computer and make an orbit. None of this made sense to me— there was just no connection.

So there were two problems with me and intro physics. One, it was incredibly hard, because I just did not have the background. Two, it was incredibly uninteresting to me. I genuinely did not care how fast a spring oscillated. I did not care how long it took a block to slide down an incline plane. So this was a double whammy. Hard and uninteresting to me. I wanted to talk about black holes and timelines and universes moving forward and backwards and things like that. I went into his office at least four times that first quarter to tell him I was quitting, that I just didn't want to do this. It wasn’t what I pictured. And he just kept saying, “You have to hang in there. This is not what physics finally is. You have to learn all these tools. You will eventually get to where you can do the kind of physics you want to do. But you have to get through this stuff.”

And it’s a message I convey a lot to my students. Like, don’t be discouraged in the early days if it seems hard and not interesting. You will get to the place where it’s about creativity, and ideas, and not just problem sets, which are not very interesting. And by the way, getting back to the question about ski racing versus science—my whole model was fall quarter was going to be about getting school stuff done, and then winter, I was going to get back to my ski racing. And that’s what actually happened. I missed most of my classes that first winter quarter. I was gone, all the time, ski racing. And he still stuck with me, talking me into continuing, and I hung in there.

It was horrible. I mean, there was a moment when I wasn’t doing school well, and I wasn’t doing ski racing well, because they just didn't coexist very well. I could feel that I was not skiing as well as I had, and manifestly not doing as well as I had been, and also not doing a very good job with my schooling. So I stuck that out for the winter, and then decided I would quit ski racing.

Dartmouth had a very equal quarter system, where summer was an equal quarter to the other quarters. So for the next two years, I took winter quarter off and worked at one of these ski academies, Mad River Academy—and coached and tutored kids in a system much like what I had been through. Which allowed me to continue to pursue academics and ski racing, in a way. Athletics has always been important to me. There has never been a time in my life that I haven't tried to have athletics in some form coexisting with my scholastic work.

Zierler:

Frances, at what point at Dartmouth did physics become not an iffy proposition dependent on one professor championing you, but just you finding your groove?

Hellman:

My senior year. There was a moment in my senior year when I was trying to decide what I wanted to do next. At one point, I decided I wanted to get a graduate degree in philosophy. I had this idea that what I really liked to do was talk about physics, rather than actually doing it, which led me to this theory that I wanted to study philosophy. A very unrealistic view, since I had not taken a single philosophy class the whole time I was at Dartmouth. I was also doing a double major in engineering, so I had taken a fair number of engineering classes. So I was also thinking about maybe I would go to graduate school in engineering. And I remember my dad saying to me, “I don’t care what you go to graduate school in, but you have to go to graduate school.” [laughs] So he was very determined on that subject. Not that that had all that much influence on me. I was still in this very rebellious teenager stage of my life.

Anyway, somewhere in there, I went to talk to Bruce Pipes, my freshman advisor. And I said something about how I wasn’t quite sure what I wanted to do. And he said, “I really think you should go to graduate school. I think you really have an aptitude for this.” Somewhere in there, he realized that I was working at Dunkin Donuts. And he said, “Why are you working at Dunkin Donuts?” And I said, “Well, because I need money.” Which seemed obvious to me, kind of the reason most people work at Dunkin Donuts. And he said, “Well, you can come work in my lab.” And it was this mind-opening moment. I said, “You mean somebody would like pay me to do science?” He said, “Yes. You can come work for me in my lab, and I will pay you. You can work here, instead of working at Dunkin Donuts.” So that was a mind-blowing idea. And that’s when I realized, as so many of our students probably still don’t understand, that when you go to graduate school, you don’t usually pay to go to graduate school. You usually are paid, as a GSR, or TA, or whatever.

And somewhere in there I said, “You'd hire me? But there’s so many better people in my class.” I still did not think of myself as one of the stars. There was one particular guy in the class who was always getting straight As. I mentioned this, and I said, “Wouldn't you rather have him?” And he looked at me and said, “I would take you over him every time. You have an aptitude for this. You have a knack for this. I think you will be a really good scientist. You just need to keep doing this.” And so he hired me, in his lab. I was working on semiconductors, on measuring the resistivity of germanium. So that was the first moment where I thought, “Wow, maybe I am actually good at this.”

Zierler:

And Frances, was this formative for the kinds of graduate programs you were going to apply to as well?

Hellman:

Yes. This whole conversation happened at the start of my senior year. I was considering where I wanted to apply to graduate school. And at that point—again, in my inimitable style of prioritizing, I had decided that I wanted to move to the West Coast because I was tired of winters and cold and snow. So [laughs] I literally applied to graduate schools up and down the West Coast. I applied to the University of Washington, Oregon State, U. Oregon, Berkeley, Stanford, Santa Barbara, Santa Cruz, San Diego. I was sick of the cold weather, and I wanted to move to the West Coast. So that’s how I selected the graduate schools I was going to apply to.

Again, I was being a rebellious teenager, even when I wasn’t a teenager. I remember the process of applying for graduate schools—the first thing, there was a GRE. And the GRE test seemed to me so incredibly stupid that when I went to the subject matter GRE, a three-hour test, I had trouble taking it seriously. For some reason, I've never been very good at circuit analysis. I don’t know why. The third problem on the GRE subject matter was a circuit problem. I spent the first hour and a half on that problem. There are something like 90 problems in the GRE, and I spent the first hour and a half trying to solve this one problem. Because I was just so determined that I was going to solve it, and I was so convinced that these were stupid tests. You know, it was sort of, “Let the chips fall where they may, this is a stupid test, so I'm just going to do what I do.”

So needless to say, I didn't do that well on the GRE subject matter test, which mattered for some of the schools I applied to and not for others, for whatever reason. The other thing I remember about applying to graduate school is you have to write an essay, a personal essay. The question was, “Why do you want to go to graduate school?” And I wrote on my graduate school application “I want to do research, and everybody tells me that in order to do research, I have to have a PhD.”

Zierler:

[laughs]

Hellman:

And that is the beginning and end of what I wrote.

Zierler:

It’s very functional. That works.

Hellman:

[laughs] It was. It seemed very functional. It’s like, what more could you say? It seemed very functional. “I want to do research, therefore—” So again, fortunately, some schools were OK with this. It’s interesting looking at where I got in and where I didn’t get in.

Zierler:

In terms of the research, were you well-defined in terms of experimentation and not theory at that point?

Hellman:

For some reason, I've always been drawn to experiment. I've never been tempted to be a theorist. I find some of the theory elegant, and I like that I can do some of it, but I've always been drawn to experiment much more. I liked it—right from the beginning, when I got my position in Bruce Pipes’ lab on measuring resistivity of germanium, doped germanium. So I've never really considered being a theorist.

Zierler:

What was your process for choosing among the schools that you got into?

Hellman:

[laughs] Again, continuing my rebellious teenager thing, I didn't actually ask anybody’s opinion. I visited. I had a clear ranking in my head. It came down to Berkeley and Stanford, both of which I had gotten into. I visited Berkeley, and it was [laughs]—I hate to trash the place that I'm at right now, but it was cold and unfriendly, and none of the professors wanted to take the time to meet with me. And there was this reputation they had that half their students failed their prelim exams, and the reputation was because they needed lots of students to be TAs for them, and then they kicked them out. Now, I think all of those are tremendous exaggerations of the reality, but that was the reputation.

And then when I visited Stanford, they had a new building, the Ginzton Lab, and it was bright and lit and friendly and all the professors wanted to meet with me, and the graduate students met with me. So it was kind of a no-brainer in the end that I went to Stanford. It just seemed like everybody was friendly. And Berkeley seemed very unfriendly.

Zierler:

Who ended up being your advisor at Stanford?

Hellman:

It was a mix of Ted Geballe and Mac Beasley. They were sort of co-advisors. You're smiling.

Zierler:

Yeah. I know them both very well. And Aharon [Kapitulnik] as well. But you were there before Aharon got there.

Hellman:

Yes, exactly. So in those days, it was called the Ted Mac Amateur Hour.

Zierler:

Not KGB yet.

Hellman:

Not KGB. In fact, those of us who were Ted-Mac’ers were quite proud of not being KGBers. [laughs] Not that we have anything against Aharon—I mean, Aharon is a great scientist.

Zierler:

Old school.

Hellman:

It’s not about that. It was just, you know, we had the Ted Mac Amateur Hour. We had our t-shirts. We played softball together. There’s an interesting piece of this, which surprises people—Ted Geballe’s wife is a relative of mine. Sissy Geballe was born Sissy Koshland, and my father’s mother, my grandmother, was a Koshland. So Sissy Geballe and my grandmother were I think first cousins?

Anyway, I did not know this. So I did not know that Ted Geballe was related to me. If I had known it, I might have hesitated before joining his research group, because it might have seemed a little odd. But I was sufficiently independent minded that I didn’t ask anyone’s opinion about joining Ted’s group. He had a different name, right? His name is Geballe, not Koshland. So I had no idea that he was related to me. There was a moment where my dad apparently called him, prior to admissions to say, “Ted, do you think you could do something about getting Frances into graduate school?” And Ted—and if you know Ted, you'll know how this would have gone down—Ted said something on the order of, “Well, I'm sure if she’s good enough, she’ll get in.” And my father called me after that to say, “I've done something that is going to make you really angry.” And I'm—“What?” And he said, “You're gonna be really angry, so just—I'm sorry in advance, but I feel like I have to tell you.” And he said, “I called the chair of the department to see if he could help you get into graduate school.” And I said, “Dad, I cannot tell you how angry I am. That is just—you just undermine everything.” He said, “Well, the good news is, the chair was adamant that there was nothing he could do to get you in. And that if you were good enough, you would get in.”

I didn't know the chair was Ted Geballe, married to a cousin of my dad’s. I picked the Geballe-Beasley Group because they were experimental condensed matter, and that’s what Bruce Pipes was. Ted turned out to be a great advisor, so I am glad I didn’t know.

Zierler:

And what was the group doing when you joined? What were the big things?

Hellman:

Superconductivity. A15 superconductors. In fact, literally, my thesis, which was on A15 superconductors, particularly Niobium(3)Tin—I discovered this cool thing about how they grew. My big paper from my thesis, a giant Phys Rev B that was the heart of my thesis, appeared in December of ’85, which is about two months after High-Tc was discovered. So, in my big paper from my thesis, the first sentence of this 20-page-long Phys Rev B is “All of the superconductors with the highest Tcs are in the A15 crystal structure.” And this is in page proof form, about two months—or less than that —after High-Tc has been discovered. And I'm looking at this thinking, “What do I do with this, since it’s now not true?” But it’s in page proof form. It’s not like you can rewrite the whole thing. So I literally inserted a little caret to say, “All of the best-known—

Zierler:

[laughs]

Hellman:

—High-Tc superconductors” or something like that. Something that I could do with a little caret insert. Anyway, I chose to work with Ted because I wanted to do experimental condensed matter. And in applied physics, he and Mac were the two leading people.

Zierler:

What was your sense of the dynamic between Ted and Mac? Was it palpable that Ted was senior, or were they really peers, as far as you could tell?

Hellman:

They were really peers, as far as I could tell. They're quite different, with very different styles. But they acted very much as peers, very much as equals. I was more Geballe’s student, and another student who joined at the same time as me, John Graybeal, was more Mac’s student. In about my third year in graduate school—there was a senior graduate student that I was working with. He was trying to finish his thesis, and I was not even allowed to touch the equipment. And so I was getting incredibly frustrated. And I remember thinking—and not just thinking; writing in my lab notebook—how much I wished that I was Mac Beasley’s student, because I would have had a well-defined project. With Ted, it always seemed like the project was very—amorphous. [laughs] And I remember I thought, “I wish I was John Graybeal. I wish I had a really well-defined thesis that I was already working on.”

And it’s interesting, because that was in my third year, and by my fifth year, I had come to appreciate that Ted was exactly, exactly the right person for me. Mac is very thorough, a great scientist and is the kind of person who designs experiments to tests theories. He’s a superb experimentalist. Whereas with Ted, Ted really is a materials physicist. His knowledge is just vast. And he thinks of things starting from the materials, and then designing experiments using those materials. Which maps well onto my own approach, or perhaps he created that approach in me!

And so in the end, I got to design my own thesis. For better or worse. There were times when—late in my thesis, Ted suddenly came up with completely other ideas. Once, near the end of my thesis, I was already supposed to be at my postdoc position at Bell Labs. Ted had a conversation with Phil Anderson about strong coupling superconductors, and this idea they had, instead of the electrons and the phonons and a weak interaction between them that led to superconductivity, the whole problem needed to be flipped the other way around. The interaction was the dominant thing. And so you had to recast everything, you didn't start with separate excitations and then a perturbation. You had to start with the coupling between them.

The bottom line is all this led to a proposed experiment involving niobium(3)antimony which was a material I had never made, and XAS measurements, which is a measurement I had never made, to test this hypothesis. And this was in February or March of 1987, and I was already supposed to be at Bell Labs for my postdoc. I was trying to finish writing my thesis. And I looked at him and I said, “I can’t do that. That’s completely outside my thesis.” And he looked at me and said, “Well, what is your thesis anyway?” [laughs] And I remember thinking, “Wow, he doesn't actually know what my thesis is!” [laughs]

Zierler:

That’s great.

Hellman:

But he was exactly the advisor I needed. He gave me the space to ask my own questions and to design my own experiments. I have so much respect for him. He is in some ways the epitome of the scientist—this encyclopedic knowledge, an expansive vision, and a willingness to let people try things. There was also a moment in graduate school where I decided I wanted to quit graduate school and go be a scuba diving instructor. And Ted was amazing – instead of freaking out, as many advisors would do, he said to me “isn’t it hurricane season in the Caribbean right now?” And I said “yes”. And he said, “so, maybe you want to wait a few months and then see if you still want to do this?”. And that was such a good point, that I ended up not going at all! Ted was just so supportive of me. I have been blessed with great mentors throughout my career. I am sure I would not be in physics at all if not for them. And I love being in Physics, so I am super grateful for their encouragement.

Zierler:

To go back to your comment about Mac’s theoretical sensibilities, where was theory at that point in superconductivity? Was it relevant to experimentation? Were the theorists in close consultation, or not necessarily?

Hellman:

They were pretty relevant. There was a lot of things going on. Remember, this is before High-Tc. So what did it mean to have a strong coupled superconductor? How did that change its properties? How did you think about it? How did you know what its coupling was? There was quite a bit of theoretical work around tunneling spectroscopy and, well, like this XAS experiment that Ted and Phil Anderson devised.

Zierler:

Did you have a theorist on your committee?

Hellman:

Conyers Herring. I chose to have Conyers on my committee. And Conyers was—I don’t know if you know Conyers, or if you ever interviewed him—he’s an amazing, just an amazing man. I mean, the depth of his knowledge, the depth of his understanding. I chose first to have him on my qualifying exam, where I was talking about granular superconductivity. And I remember he asked me a question about the entropy [laughs] of granular superconductors, where I just looked at him and I thought, “I have absolutely no idea. I do not know even slightly the answer to this question.” And then I looked at both Mac and Ted, who were in the room, as well, and I realized they don’t have any idea, either, so I was not alone in this. [laughs]

Conyers was remarkable —the depth of his questions. I mean —he was very polite, but you'd be in a seminar, and he would raise his hand and he’d ask this question, and you could just see the panic on the speaker’s face. Because what he’d be asking was clearly not actually from left field, but it was from left field because nobody had ever thought about it before. So I asked him to be on my qual exam and then to be on my thesis. And I remember my fellow students were, “Are you crazy? He asks such hard questions.” And I thought, “But I want him to ask hard questions. That’s a good thing. Then we figure out the answers—” So I viewed that as a good thing. And so yeah, so Conyers was my theorist.

And what I ended up doing my thesis on—it was on the A15 superconductors, but really at its heart, what it was about was that I discovered a new type of surface segregation phenomena that affected them. I was forced into this from the experimental results—the things I was measuring, I realized I could only explain them via materials science. It’s related to surface segregation but a new version of it, which happens in a polycrystalline film. That you get what I called a surface-to-surface segregation, where instead of just segregation from the bulk up to the surface, there would be segregation from like a 1:1:1 surface over to a 1:0:0 surface. Actually, you go the other way around, from 1:0:0 to 1:1:1 in the material I was looking at. Chemically segregated. And that was a new idea. Conyers was very excited by the idea. It was not something he had seen before. And so we interacted a bit over that. And it was quite fun to have thought of something that he had not explicitly already thought of. I ended up writing one of my few sole-authored papers on this, with Ted’s encouragement.

Zierler:

Did you apply widely for postdocs or was Bell Labs set, and that was a very easy decision?

Hellman:

The latter. It sort of fell into my lap. I knew some people at Bell, and they asked me to apply. And I went and did an interview, and they offered me the job working on magnetic thin film materials, and I took it. So I didn't apply anywhere else.

Zierler:

Who was the point of contact? Was it a Stanford connection?

Hellman:

No. Well, a little bit. I was a postdoc with three people, one of whom was a Stanford person, which was Bruce van Dover. Bruce van Dover had been a Stanford student with Mac Beasley, primarily, maybe four years older than me. And then there was Mike Gyorgy who was my primary postdoc supervisor. He was great. Mike always insisted that he—something that I still believe and still say to my students—that I did not work for him; I worked with him. He was very adamant about that.

There’s kind of an interesting story around how I ended up working in a magnetism group. I did my thesis on superconductivity, and at one point I went to the D- and F-band superconductivity conference, which includes more exotic superconductors. That includes the A15s, but also various other heavy metals and things like that.

So I went to this D- and F-band superconductivity conference, and from there, I went to a conference in Erice, Italy, on superconductivity. And at that conference, and at the D- and F-band conference, I heard about magnetic superconductors, which was a big important topic in those days. And again, an important topic again now, with High-Tc. But in those days, it was about these heavy fermion superconductors and the coexistence of magnetism and superconductivity, which everybody thought couldn't coexist, but it turns out they can. Different parts of the Fermi surface, things like that.

I heard a talk by Brian Maple, who later was a colleague of mine at UC San Diego on magnetic superconductors. And I loved that idea. It was such an exciting idea, that they didn't have to oppose each other. And in my rather pragmatic approach to things, I decided, “Okay, I've done my thesis work on superconductivity, but now I have to learn about magnetism. Because otherwise I can’t work on magnetic superconductors —I have to know about both.” So I decided that I wanted to do my postdoc in magnetism, and specifically in magnetic thin films.

So the opportunity arose —Bruce van Dover was working on magnetic thin films, and he was part of this group, including Mike Gyorgy. Mike Gyorgy was an older guy who had been working at Bell back in the days when magnetic bubbles were a possible recording technology. He was one of the main people in that whole field. So, I went to Bell Labs, in order to work with them.

Zierler:

And this was very much still heyday Bell Labs?

Hellman:

Oh, yes. Although Judge Greene’s decision had already come down. And so a lot of people asked, “Why would I go to Bell Labs when it was clear the heyday was over?” And I said, “Well, it’s not over—for a postdoc, the decision doesn't matter.” It didn't affect me as a postdoc at all.

Zierler:

And the funding, the emphasis on basic science, none of that changed as far as you could tell?

Hellman:

Not in the timeframe that I was there. Not at all.

Zierler:

And what years is this? When are you at Bell?

Hellman:

1985 to 1987.

Zierler:

Okay, so really like that’s the tail end of the good times.

Hellman:

That’s exactly right. So it was a great time to be there. Mike Gyorgy was probably the best collaborator I've ever had. He was infinitely patient and willing to work complicated questions through— I would just come into his office, and we would scribble on the board. He always had time. He never said to me, “Look, I'm sorry, I can only give you half an hour.” He seemed to have as much time for me as I needed. We worked all kinds of science problems through together. He was just an amazing man.

I then applied for jobs, after my postdoc—1987 was a good time to be applying for jobs, so I applied for assistant professorships all over—I had something like 14 job offers. I had one big advantage—everybody else who was a postdoc was working in superconductivity, and I wasn’t. I was working in magnetism. [laughs] And so I stood out. It was kind of ironic. I was pretty much the only person who used to work in High-Tc superconductors. Like, I used to work in High-Tc back when they were A15s. But I think I stood out as a result. I mean, that’s not quite fair. I did a little bit of work in High-Tc while I was at Bell Labs, also.

Zierler:

Were you publishing a lot at Bell Labs?

Hellman:

A fair amount. I was only there for two years, but I got a fair number of publications. And a couple of patents. I was on a patent for levitation, which was fun. [laughs]

Zierler:

What was the intellectual value of the sequencing of superconductivity and then magnetism as opposed to theoretically magnetism and then superconductivity?

Hellman:

I don’t know that anybody but me thought of it that way. I had this idea that I was now going to be in a position to work on magnetic superconductors, but in fact I've never worked on magnetic superconductors. So really the value was that I had both backgrounds. I think that was viewed as a value. And the value of magnetism—I mean, magnetism was interesting fundamental science but also with applications. And I've always liked that aspect of condensed matter physics, that there’s an applications part of it, but there’s also a fundamental science part of it. I can tailor projects for graduate students depending on their long-term goals that are either more fundamental or more applied. I'm not a big fan of the picture of there being a spectrum—it’s not a spectrum. It’s much more holistic and integrated than that. But nonetheless, I like having both aspects to my work.

Zierler:

So between the engineering component of your undergraduate, your dad and the entrepreneurial thing, did you ever consider going into industry?

Hellman:

I totally considered going into industry. When I was in graduate school, I was completely convinced that I wanted to work in industry. And the reason I was completely convinced is I didn't think I was any good at multitasking. I had it very clearly in my head, “I do not do well multitasking.” As I used to joke when I was in college, I'm the kind of person who, when you hit exam week, all of a sudden the bathroom has to be cleaned. Like the bathroom that has managed to be filthy the entire quarter, all of a sudden, in the middle of exam week, is the time—“Oh my god, I can’t stand this any longer. I have to go clean the bathroom.”

Zierler:

So at this point, let me refer readers back to the beginning part of our conversation, when you told me all of the titles that you currently hold. [laughs]

Hellman:

Yes, exactly. And in a way, I get a lot done but I'm a procrastinator, and I'm deadline driven. And if I'm going to do something hard, like writing a paper, I have to create this big block of time. Because I will procrastinate. I mean, I'll do everything else that has to get done, like cleaning the bathroom, until finally the only thing left in front of me is this big project that I have to tackle. I was always convinced that I am not a multitasker, never going to work for me to multitask. So, industry seemed like the right solution. I'm not sure I can actually defend that statement, even—why industry is less multitasking—but I had it in my head. So I was convinced I wanted to work in industry, and that’s partially why I took the job at Bell Labs. Now, of course, Bell Labs is hardly your typical industry.

And I will say along the way—this is another diversionary story— maybe my fourth or fifth year in grad school—I decided that what I really wanted to do was something like a Peace Corps job. I wanted to do something totally not physics. And I had this idea that I was going to go to China for a postdoc position. I didn't want to do a normal Peace Corps job, because I still did want to do science. But I thought, “I could go do a postdoc in China.” That was my solution to this dilemma. If I did a postdoc in China, I would be living in a really foreign country, Third Worldish—keep in mind when this is; this is 1982.

So in 1982, I went to China [laughs] on an alumni trip with my undergraduate institution. But I also tacked on a week at the end where I went back to the Institute of Physics and spent a week living in one of the friendship hotels and visiting the Institute of Physics, and meeting with people, and talking with them, and exploring the idea of doing a postdoc there. I was totally convinced that’s what I was going to do. However, I realized at some point that it wasn’t going to work the way I pictured, at all. You were very, very isolated. Like you don’t actually live in a foreign country; you live in a friendship hotel with lots of other foreigners. And I realized this wasn’t going to meet my goals at all. So I bailed on that idea.

So mine was an equally naïve idea of industry, that Bell Labs was going to be industry. Which of course, it wasn’t. But I was convinced that I wanted to work in industry, because I didn't do a good job multitasking. Then while I was at Bell Labs, a couple of things happened. I was doing all this magnetism work, but I also had in my mind that I had to learn how to do ultra-high vacuum depositions. I mean, I’d made thin films but I had no ultra-high vacuum thin film experience. I mentally had this checklist of what I needed to learn before I went on. I needed to learn how to do ultra-high vacuum (UHV) experiments. So I found a temporary stint in UHV with Ray Tung working on silicides, both cobalt and nickel silicides. And learned to do ultra-high vacuum stuff, and learned to do UHV type techniques and characterization tools and so forth. RBS and XPS and LEED —Low Angle Electron Diffraction, and things like that. So really good experience.

But at the end of it, my immediate supervisor—not Mike Gyorgy but my department head, said if I wanted a job at Bell Labs, I was going to need to continue with this specific field. And I'm like, “I don’t want to—I mean, it’s fine, working on silicides. I like this. I've got a couple papers. But I don’t want to spend my whole career working on silicides.” So I realized this wasn’t going to work. And so I ended up deciding to apply for assistant professor positions.

Zierler:

All in physics departments, or even at this point, it’s interdisciplinary?

Hellman:

So for my postdoc, I only applied for one position. And graduate school, those were all physics departments. Assistant professor however I applied broadly, to a few chemistry schools, some electrical engineering, some materials science, but then ended up deciding that I really liked doing physics, or liked teaching physics, specifically. And so that was the decision. I was offered a position at UC San Diego, which has a magnetism effort —specifically a magnetic recording center, and so that was attractive. And the funny thing is, I multitask like crazy now. Like that’s just part of my life. I still have to clear everything away before I'm ready to tackle the next big, hard thing. But I love academia, and I'm really glad that’s where I've ended up.

Zierler:

Was San Diego still in growth mode at that point in terms of its graduate program?

Hellman:

Not so much. UC San Diego has an interesting history. It started as an entirely graduate program and did not add undergraduates till somewhat later. It also started with a strong emphasis on student athletes, BUT the idea was not focused on intercollegiate athletics but instead on students having intra college athletics programs, with as many as possible participating. Also, they launched the physics department with a particularly interesting breadthwith respect to disciplines, specifically condensed matter physics and biophysics. The traditional great physics departments—the Princetons and the Harvards and the Caltechs and Stanford—prior to the 1960’s they did not have much condensed matter or biophysics. The 1960s was really when all that started changing. So some departments like UCSD capitalized on this by developing some of the very early, high profile biophysics and condensed matter physics.

Zierler:

Well, it didn't have any of that institutional baggage.

Hellman:

Correct. And so they started from scratch, and managed to hire some really great people. They also of course had some great particle physicists and astrophysicists. But they did hire some really great biophysics and condensed matter people at UCSD, in the 1960s. Stanford Applied Physics — my graduate program – also did this. That’s when Geballe and Beasley were hired. But no, I think UCSD, by the time I got there in 1987, I think it was pretty well formed.

Zierler:

And who was dean that would have been encouraging these developments at this point?

Hellman:

Oh, gosh. Who was the dean then? I do not remember. I remember meeting the dean when I applied.

Zierler:

Who was department chair?

Hellman:

The chair position rotated every three years. Roger Dashen was chair at a particularly important point of my career as an successor. Roger was an amazing chair—oh, my gosh, talk about breadth and interest. Roger would walk around the hallways and he’d stick his head in my office and just say, “Tell me what’s new. What are you learning?” He was the kind of person who had the breadth to think about almost any aspect of physics. When you wrote a qualifying exam or the prelim exam, you could hand it to him, and if he couldn't finish the 3-hour exam in well under an hour, it was too hard. He was this amazingly broad, amazingly deep and engaged physicist—he was somebody I looked up to so much. So later, when I became department chair, I tried my best to honor the tradition that I had been set. He made things happen. If there was something that was needed, he’d find a way to do it.

Zierler:

Now, did you start up a lab right away? Did you have the resources to do that?

Hellman:

Yep. Kind of. The physics world was in a transition where assistant professors were starting to get more real startup packages, but they were still kinda small. So I didn't really have enough. And in particular, I didn't have enough to actually build the lab. I inherited a lab that wasn’t even empty. My lab literally was full of stuff. The first thing I had to do was get rid of all the stuff that was in my lab, keep what I wanted. Literally the first thing I had to do, was clean all the person’s stuff out of there, that had been in there before.

Somewhere in there, as an assistant professor, I wrote some grant proposals, and got enough money to put together my first deposition system. But there was no money for the electrical power and the cooling water and things like that. And I remember going to the vice chair for facilities, and he said to me, “You're asking for the wrong amount of money.” And I said, “I don’t know what you mean.” And he said, “It’s too much for the department, but it’s too little to go to the university.” At the time, there was a senior person who had just been hired who had gone massively overbudget, and they had managed to accommodate it. And I just looked at him and I said—I think I actually said something like, “This is”—well, let’s just say—"bull-pucky.”

Zierler:

[laughs]

Hellman:

Something like that. And I marched out of the office. I said, “This is ridiculous.” I was going to leave. There was no point in my staying. I couldn’t ask NSF for these infrastructure needs—I think I needed like $50,000 or $70,000 to put a new transformer in the building to power the needs of the deposition system. And NSF is not going to give you $70,000 for a transformer for the building. So I went to the department chair, which was Roger. And he just said, “We'll figure out how to do this.” And he did. I still don’t know how he did it, but he did.

Zierler:

Between the opportunity to build a lab and the fact that the resources were so slim, that really requires you to focus on what’s most important experimentally at this stage in your career. So what was it?

Hellman:

I was doing the multitasking thing. I had two priorities, for better or worse. My thesis work had involved what we called microcalorimeters, which were these little silicon and sapphire chips that we used for calorimetry on relatively thick films, but nonetheless thinner than commercial calorimeters can handle. Micron-thick films, using these little silicon and sapphire chips that then you wire-bond leads out to the side. You don’t work in an adiabatic limit. You're working in the time domain, a technique called the small delta T method specifically. You actually do the equivalent of—it’s much like an RC circuit. You put power in to a resistive heater, you reach an equilibrium, where the power flowing in through the heater is equal to the heat flowing out through conduction or radiation. And after that equilibrium is reached, you then turn off your heater, and measure the relaxation. And that relaxation time is the specific heat over the thermal conductance of the link. So much like measuring a capacitor’s capacitance, by measuring the time decay using a known resistor. You'd have to know the resister, then you would measure the time constant, and that would give you the capacitance. The small delta T calorimetry technique is much the same. You have to measure the thermal link, which, without going into all the details, is a hard thing to do, but doable. And then you measure the time decay, and that gives you the heat capacity.

I had done that in my thesis work, and so when I got to San Diego, my goal was to set up two things. One, I needed to be able to make thin films, so I needed a thin film evaporation system—you know, a high vacuum or ultra high vacuum physical vapor deposition system, multisource, things like that. And two, I was going to design the next generation of calorimeters. I looked into this, and I realized that there was a new MEMS technology that was membrane-based. People were making these amazingly strong, thin amorphous silicon nitride membranes. And you could make them pretty large, but thin, and still strong enough to stand up to being handled. People were using them for all sorts of things. They were using them for x-ray masks, for sound velocity measurements, pressure transducers. I realized I could use them for heat capacity, because the thermal conductance is so low through the membrane that I could put a sample in the middle, and the membrane would thermally isolate the sample area to allow the small delta T technique to work.

So I set out to do two things. One was to build a way of doing thin film evaporation. And the other was to design this next generation microcalorimeters, membrane-based. A calorimeter on a chip, if you will. It’s a completely self-contained calorimeter, using MEMS technology to make these membranes. And that, interestingly, involved coming to Berkeley. [laughs] I looked around, where I could make these membranes. It turned out Berkeley had what they called the Nanolab at the time. So I visited and talked with people at Berkeley, and they welcomed outside users. For the next five or six years, my research group—pretty much my students—would every summer rent an apartment in Berkeley, and they would fabricate devices.

It took us about five years to get this to work. It was very close to not happening in time for me to get tenure. That was pretty darn risky—as somebody said to me at San Diego, “You've put an awful lot of eggs in that basket.” And I said, “Well, I don’t know what else to do. That’s what I want to do. So I have to put all of my eggs in that basket.” It worked, but barely—the paper, the key paper, to getting me tenure, came out pretty much exactly when it had to come out so that I wasn’t denied tenure. [laughs]

Zierler:

Phew.

Hellman:

Yeah. It was a little interesting. It never crossed my mind not to do that. I really had this feeling that if this didn't work, so be it. “This is what I want to do, and if I can’t do it here, I'll go somewhere else, or I'll do something else. But this is what I want to do, so I'm just going to do it.”

Zierler:

Did you take on graduate students right away?

Hellman:

Yep, yep. I had two, in the early days. The first student was a person who was relatively senior. In fact, he was older than me, slightly. He had had a whole career before coming to graduate school, then had started in one group, then was leaving that group, so joined me. He did not work on the heat capacity development. He worked on getting our deposition system set up. I had another student who worked on the calorimetry, the microcals, as we called them then. So yeah, I had students pretty much right away.

My early days as an assistant professor were challenging – people say things like, “Oh, it’s not worth it to me to write a grant proposal if it’s only for $20,000.” And I'm thinking, “Hmmm. When I was an assistant professor, $5,000 was enough.” [laughs] Like anything was enough. There really was no lower limit. I would have written a grant proposal for $1,000 if there was one out there. I wrote grant proposals all the time. I wrote them for whatever amount they had. And I just kept writing them. I think the least I ever did get was $5,000, but if there had been an $1,000 grant proposal, I would have written it. There was no such thing as a threshold. Maybe if it involved like 30 pages of writing, maybe I would have said, “This isn’t worth it.” But that’s not the way I thought about it. I just thought: “I have to write these proposals, because I need money to support my work.” So I just wrote lots and lots and lots of proposals. And fortunately had enough success, although barely. It was all very much right on the edge of not working.

Zierler:

Given how precarious the timing was between the publication and tenure, I'm going to assume that achieving tenure did not necessarily make you any more adventurous in your research. That was always your sensibility.

Hellman:

I guess I've always been very stubborn about what I want to do. People at the time, some of the senior faculty at San Diego, suggested I collaborate with them on some more surefire project. But I always had this feeling, “But that doesn't make sense. If I collaborate with you, then it’s your project.” So it just never made sense to me. It was kind of them to offer, but it didn't seem like that would work. And in a way, I don’t think it would have. You have to do your own work as an assistant professor.

Zierler:

At this juncture in your career, what are some of the advances technologically in the field that are relevant for you?

Hellman:

Well, the low stress silicon nitride membrane was a big deal. That was a technology I found in 1987 that had been only very recently developed —like within a couple of years. When people talk about the MEMS revolution, this membrane is a particular aspect of the MEMS revolution. It’s not the silicon cantilever MEMS revolution, but it is this incredibly strong nitride membrane. It’s remarkable. We make these membranes that are 30 nanometers thick and two millimeters on a side, and strong enough to be handled— you could pick up the chip and stick it in my deposition system and grow on it. So that was one of the key technologies, with perhaps the biggest direct impact on my career. Beyond that—somebody once said to me—the biggest asset in your lab is always your brain. It’s your ideas. That is the most important thing in your lab. Which I believe and agree with. But you do have to have the tools. So my deposition system was very key to everything I did. I pieced that funding together and it was a really important step in making things work. I've moved but I still use that deposition system. It moved with me here to Berkeley and is still one of my key tools.

Zierler:

Of course, you're firmly rooted in academia. I wonder, though, in terms of your overall research agenda, if you segmented projects according to basic science interests and possible applications.

Hellman:

I'm always cognizant of possible applications. Both my magnetism work and the work on amorphous insulators have both basic science aspects and possible applications. Let me use the example of my DOE-funded work on magnetism. DOE is quite clear that we are supposed to be doing basic science —the magnetism work is funded through the Basic Energy Sciences directorate. They're very clear that we are not supposed to be aiming at applications. We're not supposed to be thinking about can we make a better device, like for magnetic recording, in this case. Instead, we are supposed to work on the underlying fundamental principles that might be limiting the application.

The way DOE might put it is, anything that has a yes/no answer, like “Can we” is not the right question. The example they like to use is if you're working on photovoltaics, it’s not about “Can we make a more efficient photovoltaic?” The question should be, “What limits the efficiency of photovoltaics?” Do we understand what limits it? Can we address those limits? There’s nothing wrong with the work having an application; it just can’t be the immediate driving factor. And so in all my work, I pretty much think of it the same way. Like, why do we work on spin electronics? It’s likely to enable more efficient, lower energy consumption, with higher density bits. But if your goal is higher efficiency, higher density, more efficient, there’s a lot of approaches you would use. Engineers are very innovative, and they come up with lots of new approaches. So aiming at the technology is the wrong thing, in my opinion, to do, as a physicist.

I don’t really know how the engineers manage to do it, because industry changes constantly. People in industry and engineering are so creative. They find something that we call a fundamental limit, like the size of a magnetic recording bit, and then they find a way around it. And if there’s some limit, there’s some assumption that we're making that we should try to understand. One of my favorite professors of all time was Dirk Walecka at Stanford. He was an amazing teacher. Sandy Fetter, also my professor at Stanford, was also pretty amazing, but Dirk Walecka, even more so. Dirk Walecka would do some incredibly complicated derivation in field theory or in quantum mechanics, both of which I took from him. And then at the end of it he’d stop and say, “Okay, what assumptions did I make?” And as a class, we’d have to sit there and go back through and think what assumptions he had made. I love that approach. It’s not about the proof; it’s about what assumptions did you make in deriving the proof. I love that.

And so in industry, yeah, they're making advances all the time. And if you just focus on that outcome, of like a faster or more efficient computer or a higher bit density, there’s many ways to get there. So for me, that’s not the right problem to work on as a physicist. We have to work on a fundamental understanding of some phenomena that then may lead us to a better device. I like knowing about what’s going on in industry, but then I follow up by thinking about, what’s the underlying physics? How do I understand what’s going on and what’s limiting it? And underpinning all that is what assumptions are getting made. When they think they're about to reach the fundamental limit of some process, what assumptions are they assuming, and that if we got rid of that, we could open up a whole new approach?

Zierler:

This begs the question if there’s a Moore’s law for MEMS devices.

Hellman:

Mmhmm. You know, Moore’s law is such a funny one, because of course it’s not a law, at all.

Zierler:

[laughs]

Hellman:

It’s an observation. It’s a goal. Well, it’s a past observation and a future goal. But there’s no particular reason it has to be true; it’s just a goal. Everybody would like it to be true. Although, I'm not positive I even want it to be true, because I'm sick of having to buy new computers. [laughs] I’d be perfectly happy to just stick with my existing computer. I don’t need more memory. I just need my computer to keep working. But anyway, leaving that aside.

So yeah, Moore’s law is a great example. There are places you're going to hit so-called fundamental limits, where you're not going to be able to get beyond that. But in reality, there’s an assumption somewhere earlier, and somebody is going to find a way to make that assumption not right. Make the material granular, or move into the third dimension. Use light instead of electrons—various things, that you can imagine, that make your assumptions go out the window. So yeah, I'm much happier trying to understand the underlying principles of something, with the hope that it will be relevant to practical things.

Zierler:

Now it’s in the mid-1990s that your interest in amorphous and crystalline thin films really comes front and center?

Hellman:

Oddly enough, I've been working on amorphous materials for a long time, starting with my postdoc position. My postdoctoral work was on amorphous terbium-iron alloys, which were used then for magnetic storage. The very first rewriteable disks were amorphous terbium-iron type materials. I learned a lot about amorphous materials in making amorphous terbium-iron. That then evolved into an appreciation of the flexibility and power of amorphous materials. So in the 1990s—maybe even very early 1990s—I started working on amorphous silicon doped with rare Earth ions. And that actually has two interesting pieces in its history. One, it started as an undergrad project. I've always worked with undergrads. I love working with undergrads. I love working with undergrads in part because there’s so little pressure. Like I can just say, “Let’s try making this material.” And it’s just an idea, and often it’s just an idea I've had for a while but we haven't tried. But with a graduate student, you have this pressure of a thesis.

Zierler:

There’s a career. There’s funding. Right.

Hellman:

There’s funding. There’s a career. There’s a thesis. Sure, there can be negative results, but you kind of hope there isn’t, because it’s hard to get a job with negative results. So there’s lots of pressure. But with undergrads, as long as you've given them a good experience, they're happy. It doesn't have to work. Nothing has to work. You just have to give them a good experience. COVID has threatened that particular approach, because it’s hard to even give them a good experience. But in normal times, you just have to give them a good experience. They have to have tried some things, had the experience of having ideas on their own, designing an experiment. All of these things are what matter.

So, I had read about how silicon at the metal insulator transition has magnetism come into it, and it is a highly correlated electron effect. And when I say silicon, I actually mean doped crystalline silicon. But it was also known to happen in amorphous silicon. I mean, we could do a whole session on the beauty of correlated electron physics—but the whole idea, in my mind—and again, with an experimentalist lens on the metal insulator transition—is pretty remarkable. Let’s start with why does the theory of metals work? Landau Fermi liquid theory is all about the fact that by the time you're done accommodating all of the many electrons you have to accommodate—the totally identical electrons in their non-equivalent states—you have so much kinetic energy that the potential energy is kind of a perturbation. Again, that’s my hand-wavy statement about why this works. More or less true. You have 7 eV worth of kinetic energy, and the potential energy is less than that.

Then the interesting thing is, if you have fewer electrons, of course you have less kinetic energy, but you also have less potential energy. And it turns out the low electron concentration is where the potential energy dominates, and the high electron concentration is where the kinetic energy dominates. And you can see that in a sense, it’s almost the definition of metals. Metals are all about kinetic energy. Metals are all about electrons that are free to move. They release themselves from their being trapped on individual ions, and they move freely through the underlying material. So that’s a kinetic energy dominated limit.

In insulators, on the other hand, electrons pretty much stick back where they started. They may be in bonds or bands and lots of complexity around that statement, but they are, in some sense, dominated by potential energy. And, the interesting thing is at the metal insulator transition, their potential and kinetic energy are close to equal. And so it’s not surprising, any time you have energies competing with each other, then small perturbations matter a lot.

So, one of the things that happens at the metal insulator transition is the whole principle of condensed matter physics with its orbitals and wave functions break down. I don’t know if you've thought about this, but the term “orbital” is a very funny one. What happened to the word “wave function”? Why are we suddenly talking about orbitals instead of wave functions? And the answer is, because that’s the one-electron solution. So we're using the one-electron solution, and then sticking all our electrons, all our 1023 electrons, into the one-electron solutions, as though that had any meaning. It’s a crazy idea. Why do I not have to re-solve the entire problem? I interactions between the electrons; I should have to create all new wave functions. Why am I still sticking them in the original wave functions, as though the other electrons didn't matter? And the answer is just what I said before—that you do, but the kinetic energy dominates so it doesn't matter. It’s only the electrons at the very top of the Fermi surface with the most kinetic energy that matter.

Anyway, the point is, in the limit where the material is about to become an insulator, the potential energy terms matter, and what happens is, in a hand-wavy sort of way, electrons want to stay away from each other. As soon as they want to stay away from each other, then you get single occupation of states. And once you have single occupation of states, the spin is visible. Normally, you have double occupation of every state. Every state is occupied one spin up, one spin down. That’s the lowest energy state. One spin up, one spin down. No magnetism. But, near the metal insulator transition, highly correlated electron phenomena start coming in, creating what are called Coulomb gaps due to potential energy effects. And, the bottom line is electrons don’t want to be in the same state anymore. So they singly occupy states. They therefore have a magnetic moment. And that magnetic moment now can interact with other magnetic moments. Which they do. So they start having a magnetic susceptibility.

I thought that was fascinating, that something purely non-magnetic, like phosphorous-doped silicon, at the metal insulator transition, has a fairly significant magnetic susceptibility. And I was like, “That’s weird. It has this big magnetic susceptibility.” So my idea was I wanted to stick Rare Earth atoms in. The thought was, “Okay, we have this little tiny, dinky magnetic susceptibility here, but I'm going to put these big magnetic moments in, and they will couple to it, and then amazing things may happen.” That was the thought process that went into this project for this undergrad. And it turned out to be completely true. Huge effects happened. Huge. And all of this was enabled by the amorphous state. You can’t put that much Rare Earth into crystalline silicon. It’s not soluble. You end up with just a screwed-up mess of nanocrystals and segregation and stuff like that. But you can make amorphous Rare Earth doped silicon pretty homogeneously.

So that was the first time I deliberately made an amorphous material that wasn’t already a known material. The amorphous terbium and iron, people had known about for a long time. And I discovered some things about it, but it was a known material. The amorphous rare Earth silicon was the first time I deliberately made something amorphous, because I couldn't make it crystalline, and I wanted to understand how this worked. And we found all these amazing properties. And then somewhere in there, as a background [laughs] to the amorphous Rare Earth silicon materials, their heat capacities specifically, I needed to make an amorphous silicon background, with no Rare Earth in it. And that’s when we started discovering things about amorphous silicon, and the remarkable effects of how we grew it.

And that has then led into this whole project that I have spent a lot of years—maybe the last ten years—researching, which is this idea that by vapor deposition, you can make your way to a state in the energy landscape that is lower in energy and lower in entropy than what you will ever reach by liquid quenching. Liquid quenching, you fall out of equilibrium. Vapor deposition, you can grow at a temperature that is low compared to the glass transition, so there is no mobility at all in the bulk, but you have surface mobility. And that surface mobility, let’s say at 0.8 of the glass transition temperature, Tg, you can grow at 0.8 Tg and still have a lot of surface mobility, so atoms on the growing surface can move around and find lower energy states. And these lower energy states turn out to have a whole lot of interesting properties. They have low mechanical loss. They don’t have all the excess heat capacity that amorphous materials usually have.

And this is all connected to the relatively old idea of the ideal glass. We're at least approaching the theoretical idea of an ideal glass by growing materials near but below their glass transition temperature. And that’s where my work on LIGO mirror coatings and other amorphous thin film insulators came from. I still do a lot of work with amorphous magnetic materials, but this ideal glass concept, and trying to figure out when this works and when it doesn't work is now a huge part of my research. We discovered this in amorphous silicon. Amorphous silicon is great for some things, not great for other things. May be great for future LIGO; it’s not great for present LIGO. So we're trying to find out, how do we generalize this idea? When does vapor deposition enable an ideal glass, with low noise and low excess heat capacity, and when does this not work?

Zierler:

Transferring into the 2000s, you become full professor. Were you recruited at Berkeley? Were you looking for a change of scenery?

Hellman:

Okay, that’s a complex question. Let me try to do justice to it. San Diego was great for me. Launched my career, helped me get started. But there is a way in which I was always an assistant professor in everybody’s mind. It was kind of ironic, because by the end of my time there, I wasn’t, of course. I was a full professor. I was on all kinds of high-level national committees. I was on the Board of Physics and Astronomy, including their Executive Committee. I had chaired and been on Advisory Committees at NSF and at National Labs. I had already been chair of two of the large units in the APS—a division and a topical group. So I had held pretty high-level positions nationally. And yet somehow San Diego still thought of me as really young.

I once had a conversation with the chair of the department. After 17 years of being at San Diego, where I had never been asked to be on a search committee. As I was leaving, I said, “You know, I just want to point out that I've been here for like 17 or 18 years, and you never asked me to be on a search committee.” And the chair, who was a very, very lovely man, looked at me and sort of paused, and he said, “Well, I guess we just thought you were very young, and we were trying to protect your career.” And I'm like, “Excuse me, but the world does not see me as young. I'm a full professor. I'm on all these high-level national committees. So it’s really only you here at San Diego who still think of me as being young and not quite ready for these responsibilities.”

Zierler:

There might be the gender thing going on here, as well.

Hellman:

I've always kind of assumed it was the gender thing. I don’t think any of it was intentional. In fact, I'm sure it wasn’t intentional; it just was. But at any rate—when I moved to Berkeley, everything flipped. I mean, I came to Berkeley, and I was only here for two years when I was asked to be the department chair. It’s like, when I got to Berkeley, everybody saw me as a senior professor and somebody who had gravitas.

At the time I moved to Berkeley, I was married to Bob Dynes, who got chosen for the presidency of the University of California. So he was moving to Oakland. I didn't want to live in San Diego and have him living up in Oakland, so I started exploring also moving. And the obvious places to move were Berkeley and Davis. Davis came through very quickly with a blanket statement that they would give me whatever I needed, and just let them know if I was interested. And I said, “I appreciate that, and it’s nice to have that as an option, but let me first see what Berkeley can do.” Because the idea of living in Davis while my husband lived in Oakland still didn’t seem like a great idea. So I said, “Let me see what Berkeley can do.” And so Berkeley was happy to make me an offer. And so that’s how that came to be.

Zierler:

Just being in the UC system, did Berkeley have a cachet for you, just in your mind? That may have been unconnected to the actual reality there. Did that influence your thinking at all?

Hellman:

Indeed, I was already in the UC system. I went to grad school at Stanford, but—

Zierler:

And negatively defined, because you specifically found Berkeley to be cold.

Hellman:

Yes, exactly. Which is an interesting point. But the reality was, for various reasons —because my ex was going to be a professor at Berkeley, and because Berkeley had such great faculty and students, Berkeley seemed preferable. I think Berkeley was trying to grow their materials group, so in a way I was a natural addition to them. They came through reasonably quickly, and so then I didn't go further than that.

Zierler:

Did you find Berkeley to be warmer? Had it matured at this point at all?

Hellman:

They were amazing! It was exactly the opposite to my experience when I was a graduate student. Yes.

Zierler:

Maybe that’s just about being a well-respected senior professor at this point and not a prospective grad student.

Hellman:

Yes, it could well be. So yes, Berkeley was incredibly welcoming, long before I even got there. As soon as the offer got made, I heard from every one of the condensed matter faculty in the department saying, “Oh, can’t wait for you to get here. We're going to collaborate on all these things.” They were incredibly welcoming. I never felt so welcomed in my career, as I was when I got to Berkeley. It was fantastic.

Zierler:

So when you become chair, how important for you is it to do some tone-setting?

Hellman:

Say more.

Zierler:

In the sense that, you had this very negative reaction as a prospective graduate student. But it was entirely different, but perhaps that was only because you were a well-respected senior professor at this point. And maybe the next Frances Hellman as a prospective graduate student really should have a much better reception.

Hellman:

So, it’s interesting—I wasn’t that conscious of it, but it is true that as chair, I paid a lot of attention to our graduate student processes, and the prelims, and the quals, and all of these sorts of things. One of the things that I have been very conscious of —both as chair and now as dean—is the dropout rate. The leaky pipeline. I don’t actually like the term “leaky pipeline,” but everybody knows what I mean by it. The fact that we differentially lose women, and we differentially lose people of color is disturbing, and it’s not okay. We can rationalize all we want about how we're trying and we're not racist and things like that, but it’s not okay. So I have spent a lot of time, both as chair and as dean, trying to improve the climate —recognizing that there’s a wealth of literature that suggests that if you make a university more welcoming to everybody, you will differentially impact the non-mainstream people, meaning the women and the people of color. You will differentially impact them. So you don’t even have to single out women or people of color; you just have to be more welcoming in general.

My experience of not coming to Berkeley because of the feeling that it was unfriendly is important. I think it was unfriendly. But I think it impacts differentially—I think women are often more sensitive to that climate. That you're not as much part of the mainstream, so you're more likely to be sensitive to an unfriendly climate and decide not to go to Berkeley. I think of that as a general theme. How do we make processes more transparent? What is the qual exam? What is the prelim? How does this work? How do you find an advisor? Making things more clear and transparent is helpful to everybody, but they differentially help women and people of color and others who are not as mainstream. In my role first as chair and now as dean—I've got a few things I'm working on now, in the final two months in the office—to try and impact this.

So yes, in a sense, I do think, as chair, one of the things that I tried to do was to make it feel more friendly. I've always thought it was important, not just to the students, but for example, to the staff, and to the junior faculty. There’s an annoyingly large number of senior physicists who somehow regard themselves as better and more important than the staff. And to my eyes, that’s just not true. We have different roles. But I am not better than the staff. I'm not more important. Everybody is important. We're all working to support the research and teaching mission of the university, and we're all important in that mission.

So, I think you asked early on about one thing I did not say early on but should have—one of the things that is a high priority for me as president-elect of APS, and is very important to the current president as well, is the culture of physics. To have a culture which is more respectful of people. I think that’s the underlying word that one needs, more respectful. More respectful of differences. More respectful of where people are in their career, how much they know, how much they don’t know. But just generally, more respectful. Also, a more welcoming culture overall. And I’d say I wanted to add that to what I said way back at the beginning, because I forgot to say that. [laughs] So yeah, I do think that was important to me as chair.

Zierler:

Berkeley physics of course has a very strong legacy, and that cuts both ways. In terms of your agenda, what you wanted to do for the tone, what were some positive aspects to build on, and what were some challenges to overcome, right from the beginning?

Hellman:

At Berkeley, you mean?

Zierler:

Right.

Hellman:

Well, there were a few specific things that I took on. Berkeley physics Department does have a remarkable legacy, and I was very honored that I was asked to be its Chair, but you can’t rest on your laurels or assume that just because you are a great department, everything the department does is great. The oral prelim is probably the most dramatic example. There was a very long tradition of an oral prelim, which was very inconsistently applied, is the best way I can describe it. It was very contingent on who gave the exam. Each oral prelim was given by two faculty. It depended on who you got, what that looked like. And that inconsistency was something I found unacceptable as a chair.

There was a specific example where the inconsistency was just too much. There was a person who was about to get kicked out of grad school, having passed with one committee, and failed with the other. And somehow the average was getting taken. [laughs] And it’s like, “Why would it be the average?” If they had taken the two exams six months apart, either order would have worked. Either order—if they passed one, and they failed the other, it wouldn't matter. If you passed the first one, you'd be done; you wouldn't take the second one. If you failed the first one, but then passed the second one, you'd have passed and you'd have stayed. But somehow the fact that they did it on the same day meant we were going to kick them out, because the average of these two was below passing, as though the average of two exams meant something. [laughs] How can you have an average of some point system with no clear standard of what the points mean. I spent a lot of time talking to people about what their goals were with the oral prelim, and could we meet them in another way, and could we create something that was a little more consistent, and—well, friendly is another part of it. And ultimately, the department voted to eliminate the oral prelim.

One of the things I did when I first started as chair was a very important way to have started. When the dean asked me if I would be the chair, I said, “That’s crazy. I don’t even know everybody’s names. I mean, I just got here. I've been here less than two years, and I'm still building my lab.” Although at that point, I was actually kind of done building my lab. I asked for a week to think about it. And I went away, and I realized that was a solvable problem, not knowing everyone’s names.

Since then, I've appointed many chairs in my role as dean. I have five departments. Every three years, there’s a new chair, so I've had quite a few new chairs. And I have this conversation with them about—there’s things you may know or not know, but those are fixable things. And so the first thing I did as chair—in fact, even before I became chair—is I went and met everybody in the department. I went to their office and talked to them, each. In their office, not my office. I made the effort to reach out, and to ask them what they thought was important in the department. What was going well, what was going badly, what did they need. And that stood me very well going forward. The fact that I had reached out to ask people. There was a real lesson there for me, that people, by and large, just want to be heard. People want to know that they've been listened to. It doesn't mean you always agree with them. It doesn't mean you always do what they wanted you to do. But you've heard them, and considered what they had to say, and incorporated what you could out of what their opinion was.

So I think that was part of my success as chair and as dean. I end up having to make a lot of decisions, and I think of myself as pretty incisive and thoughtful about the decisions, but I am also very proud of the fact that I listen to people and am very willing to change my mind when I’m presented with facts or ideas I had not thought of. I've also always been very respectful of staff, and I think that has served me well over the years. At one point, I was given the title of honorary staff member, and I was given a key to the kitchen. [laughs] Because the kitchen was supposed to be for the staff, and they gave me a key, because I was an honorary staff person. And I think that’s because I've always been respectful of staff. I really value what they do. I truly believe I'm just in a different role than them. I mean, yeah, I have more learning. I've been to more years of school. I've discovered more things in science than they have discovered. But that doesn't make me better; it’s just a different role. And I'm very appreciative of staff. Staff who try hard. I mean, there are staff who are phoning it in; that’s a different story.

Zierler:

As chair, you have a unique vantage point to gauge Berkeley’s strength vis a vis other top physics programs. So I wonder, in terms of recruitment, where were you seeing, within the department, in terms of the various subfields—where was Berkeley winning out against Harvard, Princeton, MIT, Caltech, that echelon—and where was it losing out?

Hellman:

Well, recruiting, and just in general supporting junior faculty, is one of the best things about being chair. And I loved being chair. I really valued it. I love the sense of being really grounded in the research and teaching missions of the university by being directly connected to the faculty, staff and students in the department —a little bit in contrast to being dean. As chair, I felt very grounded in the department and the needs of the faculty, staff, and students, and their aspirations and their needs and how to support them. Wait, where was I going? What was your question? I just got off track there.

Zierler:

Recruitment. Where was Berkeley doing well?

Hellman:

Oh, yeah, recruitment, right. I helped to create a strategic plan for the department. About my third year as chair, we went through a whole strategic planning exercise including prioritization. I asked the question in everybody’s minds, what were the most important places that Berkeley Physics could do well in? Not the most important areas in all of science, but where can Berkeley Physics be particularly impactful, where we should hire another person? We created a strategic plan out of that. And out of that came our recruiting goals. There were very few recruitments that I lost. Like, really few. It was one of the places that I was particularly determined and adept as chair.

Zierler:

These are good years for Berkeley physics, is what you're saying.

Hellman:

Exactly. It was good years for Berkeley physics. And I think Berkeley has a good reputation for the physics that we do, and the graduate students and things like that.

Zierler:

To go back to my very first question, did you come to Berkeley dual-hatted? Was the affiliation in both departments from the beginning?

Hellman:

Almost. Officially I'm 100% physics and 0% materials science. And as I'm fond of joking, 0% is infinitely different than none. [laughs] And I mean something quite specific by that. Zero-percent faculty—I don’t teach in materials science, but I am a full voting member of the department. I can take graduate students. I can be the sole supervisor of graduate students, and have had quite a few materials science students. So it is literally infinitely different than no affiliation. But I am 100% in physics. And also LBNL - actually the LBNL one came first, oddly. I was allowed to be an LBNL faculty staff scientist even while I was at San Diego. So that one actually happened first.

Zierler:

What did you learn about your insistence on maintaining your research agenda as chair, in terms of balance, in terms of what you could take on, in terms of what you had to say no to?

Hellman:

I have always been convinced that it is possible to do both. And I point to some very high-profile examples. Whenever anybody would ask me about this. People would routinely say, “Oh, that’s impossible to do.” And I said, “It can’t be impossible. Bob Birgeneau does it.” Bob Birgeneau kept a research effort going even while he was chancellor. Paul Alivisatos, while he was director of LBNL and then as Provost. They both kept—not just research, but world-class research efforts. I was never in charge of Paul Alivisatos’ merit cases, but I was in charge of Bob Birgeneau’s merit cases, and I got to see, first-hand, the quality of research he was continuing to do, while he was chancellor. So it is possible to do.

You have to be pretty organized. I was very structured. I split my day in half. I spent mornings in my research office. I had a separate phone number, a separate email, separate calendar. And then in the afternoons, I would go to my chair’s office, or then my dean’s office. I really attempted to divide my day. I’d meet with all my graduate students in the morning, so that they get priority in the mornings. And then the afternoons were for my dean’s or my chair’s meetings.

And I think there was tremendous value in my doing this. Because my feet were on the ground. I knew what the faculty were experiencing—when they were unhappy about facilities problems, or problems with central administration, HR, any of those things, I knew firsthand what they were experiencing. It also left me in this position, which is sort of an interesting one—I can step down at any time. I've always been able to step down at any time. And I feel like that gave me a power. Like I could speak truth to power, without fear of retribution. Because they're not going to fire me from my faculty position. So the worst they can do is tell me I can’t be chair any longer, or I can’t be dean any longer. And there’s times that it seems like that would be kind of a plus! [laughs]

Zierler:

“I dare you.”

Hellman:

Exactly! It enabled me to speak truth to power. It enabled me to speak out when I think others couldn’t —if you're worried about advancing your career, if you're on an administrative track and that’s an important thing to you, I think it constrains you from being able to speak truth to power. So, to my eyes, I think continuing to do my research has been important professionally as well as personally—it gives me a balance in my life. There are times when some of the administrative stuff is just—oh my god, it’s so tedious and/or frustrating. And then I can remember, look, why I really do this. At one point, I had stopped teaching classes, and I was sitting in a meeting thinking, “Oh my god, I hate this. I am so sick of these meetings. I'm so sick of everybody talking.”

Zierler:

“What am I doing? Why am I doing this?”

Hellman:

“Why am I doing this?” And I got a message come through on my email, which I was reading at the time, in this meeting, asking for freshman seminar volunteers. And I thought, “I can do that. I can do a freshman seminar.” I love to teach. And I love that sort of informal teaching —it’s 15 undergraduates, they're freshman. It’s a one-unit class. I can teach them about magnetism, which is what I did my freshman seminar on. I can let them explore. They can do a project in magnetism, there’s lots of possible projects that would appeal to a wide range of students. And it will give me some contact to these undergrads. So I did that pretty much the whole time I've been dean. I love the reality of still having my research and a little bit of my teaching going, as well as the mentoring of the students in my research group. It leaves me grounded. It leaves me better connected.

There’s one aspect that I have not mentioned this whole time, which I probably should—I'm a scientist in residence at the San Francisco Exploratorium. Right before I became dean, I did a six-month sabbatical, one semester, at the Exploratorium. And I had the experience of helping to build exhibits. There’s a way in which the Exploratorium, it’s my parallel universe life. Like if I had not gone down the academic track, I'm convinced I would have ended up at the Exploratorium. I love that outreach to the public. I love that sense of trying to—not teach the public, but to engage with the public. Help them to see the scientist in them, that they can experience that. So I love that aspect. I did that for most of the time that I was chair and dean—well, the dean time—it’s still on my calendar every week. But I got overrun with meetings, so it got less and less often that I went. Of course, in COVID, I haven't gone at all. But in reality, once the APS commitments took off, there just was not room to also do the Exploratorium. But I love the things I was doing at the Exploratorium and I need to not let go of those things.

The administrative side of things gets tiresome, at times. So I really value having that grounding of being still a researcher and teacher and exhibit designer, including the humility of trying to make an exhibit that people can use, and the moment you realize, they're just not seeing what I thought they would see. They're not doing what I thought they would do! They're doing something completely different! Like: “No, you're supposed to press this button. Here, it says ‘Press the big red button,’ not that little black button over there. Why are you pressing that button?” But you have to get over that! Because that’s the way this works—no one is wrong at an exhibit. There are no grades. So, as the exhibit developer—if they're pressing the wrong button, it’s you who have made the mistake, not them. I've loved that experience of trying to figure out how to get people to love science the way I love science.

Zierler:

Frances, I'm not hearing a narrative trajectory that ends with you stepping down from chair and becoming dean later. How does that happen?

Hellman:

Okay, well, there was a sequence. I was chair for six years, which is sort of two terms. And I felt like I needed to do that, because the first three years, you're just learning how to do things. So I chose to do a second three years, to get some stuff done. But then it was clear to me, at the end of those six years, that I had to make a choice. That my research needed attention—oddly, my research productivity was the highest in the sixth year, of any of the years, but that’s not coincidence. You're writing up the papers. In reality, the research had reached its natural conclusions. The apparent productivity was the highest, but the actual productivity was about to fall off a cliff. I was out of experiments, and I needed to shift directions. I needed to write new grant proposals, create new areas of research focus. So I stepped down at the end of six years as chair, because otherwise my research was going to be done.

And then the dean said to me, “I'm going to be stepping down in another year or so, and you are just a natural person to be the next dean.” And in a completely unpolitic moment—sometimes I talk without thinking quite hard enough—I said, “Oh my god, that is the last job on Earth I’d want.” [laughs] And I literally said that to the dean, which is not a very friendly kind of a thing to say. Like, “Yours is the last job on Earth I’d want.” But I meant it. I liked being chair, I liked having my feet grounded in the department, and I had no ambition to be an administrator.

Zierler:

It’s also discipline-specific, though. Dean is like—it’s everything.

Hellman:

It is. But it is Math and Physical Sciences, so it’s not that broad. My deanship is—it’s math, statistics, astronomy, physics, and earth and planetary science. It’s close enough—I mean, I have to a little bit wave my hands when I'm talking about math. The research they do is kind of esoteric compared to what I understand. But it’s still not infinitely far from my knowledge base. Anyway, I did not think I wanted to apply for dean. And I didn't apply. They started the dean search, and I did not apply. I had a whole sabbatical planned, like I mentioned. I was going to do a scientist in residence at the Exploratorium, and I needed to kick start some new directions in my research. But then they came to me and said, “You should apply. We need you to apply. We want you to apply.” “We think you'd make a great dean.” And I said, “Well, I'm sorry, but I've got this whole Exploratorium thing planned.” And, “I can’t do it now. But maybe in a year.” And they said, “Well, then, why don’t you just apply, and then we'll worry about that when we get there?” So they talked me into applying. And then they offered me the position. And I said, “Well, but I told you I couldn't start for a year.” So that’s when we got into the negotiation, and we compromised on six months. So that is how that came to be—yeah. [laughs]

And, you know, being the dean—the part that is facing the departments is much like when I was chair, and I love that part. I feel like I'm still supporting individual faculty and staff and students. I love Berkeley’s approach of prioritizing hiring assistant professors. I just love that approach. Sometimes I've referred to it as like the Oakland A’s. You start with raw talent, so you have to recognize the raw talent, but then your goal is to turn them into top players —you're the one who helps them launch their careers. And I love that aspect. Instead of the Harvards and the Stanfords of this world, who steal our faculty away after they've become successful! I really, really value this focus on assistant professors. So that part of my job is great, as dean. The part of my job that is rewarding —the rest of it is less good. [laughs]

Zierler:

So just to bring our narrative up to the present, back to the science, the past five years, what have been the scientific topics that have been most compelling to you?

Hellman:

Well, there has really been two, and they're both related to the amorphous materials idea. One of them is this ideal glass, and the hope that we have found a way to make an ideal glass by taking advantage of thin film growth techniques, which gives the ability to have significant atomic diffusion at the growing film surface despite being well below the glass transition temperature. The pragmatic aspect of that could be the next LIGO mirror coating. I had a high profile article in The Atlantic, which was very cool, although a bit scary, because you don’t control what writers write. I got more credit in that than I was totally comfortable having. And some of my LIGO colleagues were a little bit taken aback that I was the one interviewed—I had to send them a note saying, “I'm sorry it sounds like I thought I was doing everything, but really that isn’t what I said.” And the few of them said, “Well, you should have demanded to see the page proofs.” And then others said, “Come on, guys, you don’t get to see the page proofs.” [laughs] “That’s not the way this works. This is a news article. Lay off of her. She just interviewed, and they wrote what they wrote. And you don’t get to see page proofs; that’s not the way it works with The Atlantic.”

So, one of the things that I'm the most excited about is this ideal glass and its connection to low noise — this has potential to be the greatest overall impact in my career. I am however badly behind in writing papers, which is one of the things suffering under my overcommitment level. My students may kill me if I don’t get caught up—I have papers sitting on my desk for months, which is terrible, and very unfair to them. But I can only do what I can do. I think the idea of this ideal glass and its connection to the excess heat capacity and excess mechanical losses in amorphous materials—or to making those go away, to be exact—that the ideal glass can make those excess excitations go away—I think that’s one of the big things that is happening in my career right now.

And then the other one, which is somewhat related, but it’s the idea that there are properties that people think only happen in crystalline materials, but that’s not true. There are local versions of things like Berry phase curvature, which is a hard-core condensed matter property reliant on crystal structure mathematics —Berry phase curvature is a gradient in k-space. So needless to say, since k-space is not a good variable in amorphous materials, how do you have a gradient with respect to k. But there is a local analog to this, we have shown this both experimentally and theoretically. We're seeing some really interesting effects, like skyrmions, enormous anomalous Hall effects, enormous spin orbit torque effects, in amorphous alloys, whose underlying physics is that of a crystalline material, but it’s happening in an amorphous material.

We're struggling to get a paper published right now on an amorphous topological insulator, which is interesting, because when we started trying to publish it, there was only a couple of theoretical papers saying this was possible, and no experimental paper, and now, a year later, there’s about ten theoretical papers saying it, and still we're the only experimental paper, but the reviewers are telling us “this is kind of obvious. There’s all these theoretical papers. What’s the big deal?” And it’s like, “Well, come on. When we started trying to publish this, there was one or two, and most theorists thought it was impossible.” So I think this idea that there are local real space descriptions, and that we're going to understand the phenomena better, for seeing that they happen in amorphous material, is going to be a very important advance.

Most theoretical physicists don’t like amorphous materials. Every textbook that you ever pick up about condensed matter physics, chapter one, or latest chapter three, talk about symmetries and periodicity, and derive all the properties from that mathematical construct. Those are the underpinnings of condensed matter physics. And, that is mathematically true. We don’t have another mathematical framework. But it’s a little like—I'll go all the way back to Zen and the Art of Motorcycle Maintenance, which is dating me—this is when I was in college, right? “The map is not the territory.” The math is not the physics. The math allows us to predict things, and it may be a completely accurate mathematical framework, but it is not the territory/not the physics. And so I think the fact we can find these exotic properties in amorphous materials gives us insight into the physics. So that’s the other side of my work—the magnetism side- where I see impact.

Zierler:

For LIGO, this is a totally different world, even universe, if you will, thinking about quantum gravity, astrophysics, cosmology. What is your expertise going to do, best case scenario, to help LIGO do Advanced LIGO? To do what it hasn’t been able to do so far?

Hellman:

Distance. More. Right now, the leading source of noise in LIGO, and Virgo and other detectors as well, is due to mechanical losses in the mirror coatings, the dissipation of energy in the coatings. The test masses have Bragg reflector coatings on them. Low-index, high-index alternating layers. That’s how you reflect the laser light. But in the coatings, these little, tiny motions of atoms, because they're amorphous, leads to noise. So, if we can beat that down by making a better amorphous material, we will be able to detect fainter signals coming from further away. The word “better” is kind of a cool one. People think when I say “better amorphous” that means more disordered. And it doesn't. It’s the opposite of that. It’s less disordered. It’s still amorphous. It’s not crystalline. But it is actually a more ordered state — a lower entropy and lower enthalpy state. In fact, the ideal glass has zero entropy. In the energy landscape, it’s the lowest energy, and it is a unique state. With entropy approaching that of the perfect crystal.

The only thing that is completely disordered is an ideal gas. In an ideal gas, the positions of the atoms are random. There is a hard-core — or a hard sphere center, but that’s negligible compared to the distance between them. Glasses are not even slightly random. The atoms are next to each other, and there’s quite a bit of short-range order. We know they are not random. So the idea that you make them “better” by making them not more disordered, but less disordered, I think is a very cool concept. It gets at what does one mean by “disorder”. If you think in structure terms, you're going to bog down in what you mean by disorder. But if you think in enthalpy and entropy, meaning if you think in thermodynamic terms, it’s very well defined. It’s a lower enthalpy, lower entropy state.

If you want to think in configurational space, like in thermodynamic theory, it’s a unique state, the state with the lowest energy other than the crystalline state, for some N atom material. The ideal glass is a unique state in configurational space, so it makes sense it has low entropy and low enthalpy. And these are not just hypothetical ideas. You can make thermodynamic measurements of the liquid state and of the glassy state and show that your vapor deposited glass has lower enthalpy and lower entropy than if you quench the liquid fast, for example.

Zierler:

So I'm doubly tickled, first because this is an opportunity to emphasize that after the Nobel and this misperception that LIGO has done what it set out to do, that there’s so much more exciting things to do.

Hellman:

Oh, there’s so much more.

Zierler:

So that’s the one part. The second part is, I love how your decidedly Earth-based expertise is really, best case scenario, going to help penetrate the deepest mysteries of the universe.

Hellman:

Yep! I do, too.

Zierler:

That’s just fantastic.

Hellman:

It is. And one of my favorite talks, the public lecture kinds of talks, is “How does condensed matter help LIGO do better detections?”

Zierler:

I can’t think of a better summary explanation for why you and a Kip Thorne or a Nergis Mavalvala would belong in the same department. It’s perfect.

Hellman:

I know. And you know, just to loop all the way back in my history to when I was an undergraduate—Dartmouth, it’s a liberal arts college, so I only had to take eight physics courses. It led to some problems when I got to grad school, by the way, that I had only taken eight physics courses. But, I got to do physics reading courses. I did two reading courses when I was an undergraduate, one on superfluid helium, and a second one on black holes. And keep in mind, this is 1977 or 1978, my senior year in college. I read this article—Cyg X-1 had just been discovered. Prior to that, everybody thought black holes were just a mathematical abstraction, that they were actually impossible. Everybody assumed there was some way around the math that led to this mathematical insanity of a singularity. But Cyg X-1 had just been discovered and was already pointing pretty strongly at being a black hole. And Kip Thorne wrote the article on which I did my reading course.

So at Dartmouth, in 1978, I did a reading course using Kip Thorne’s Physics Today article on black holes. I wrote this whole report about Cyg X-1 and black holes and Kip Thorne. And then graduate school, postdoc, and I go to UCSD as an assistant professor, and Kip Thorne is coming to give a colloquium. I read this article by him when I was an undergrad, about black holes. And now he’s coming to UCSD to give a colloquium, and I'm totally starstruck, and was like, “Oh my god, I get to meet Kip Thorne.” And I meet him! And he’s only like ten years older than me!

Zierler:

[laughs]

Hellman:

[laughs] It was this just completely shocking moment, where I realize this. He’s not some old guy. I mean, I'm 30, and he’s probably 40 at that point. [laughs] And he’s a totally lovely person. All three of them are.

I got to meet Rai Weiss later. The paper that led me into being in LIGO was a Phys Rev Letter showing how we could make the losses go away in amorphous silicon, completely reproducibly, by raising the growth temperature. Not crystallizing them, but just making a more ordered but still amorphous material. It was a fairly high-profile paper, particularly in the LIGO world, because it was the first time anybody seemed to have a systematic way of making these losses go away. These losses are supposed to be universal glass properties. That means you can’t do anything about them. But that turns out not to be true. You can reduce them.

So Rai Weiss had not yet won the Nobel Prize, and he was visiting Berkeley to give one of our colloquia, and his home institution MIT was completely up in arms, because the Nobel Prize was about to be announced. Rai Weiss is visiting Berkeley, and the Nobel Prize was to be announced that night, and he’s in Berkeley, not MIT.

Zierler:

[laughs]

Hellman:

So they're not happy about any of this. And Rai Weiss—I don’t know if you know Rai or not—

Zierler:

I do, I do.

Hellman:

Yeah, so you can just picture him. He’s like, “I don’t care that they want me there. I'm not going to change my plans because I might win the Nobel Prize tonight.” And in fact, he didn't win it that night. It was a year later that he won it. But he was like, “I'm just not going to change my plans because I might win the Nobel Prize and the PR people want me at MIT.” He is visiting Berkeley and he has a whole set of meetings scheduled. And I'm on the meeting schedule, or maybe I'm not on the meeting schedule. I was going to introduce him later, I think because I was dean, and I wanted to make sure he knew that I was going to be introducing him, and that I was now being invited to join LIGO. And I wanted to not surprise him with that information.

So I walk in the office, and he and Saul Perlmutter are having a whole talk about Nobel Prizes. You know, everybody really thinks he’s going to win that night. And Saul’s offering him all this helpful advice as to what you do when you win the Nobel Prize. So I stick my head in a little timidly and say, “Um, Rai, I just wanted to say hi to you quickly before I have to introduce you later. And just want you to know that I'm going to be joining LIGO”—so, this whole LIGO thing. And he looks at me, and he goes, “Are you Hellman?”

Zierler:

[laughs]

Hellman:

And I'm like, “Yes.” And he goes, “You're Hellman! I need to meet you!” And he kicks Saul out. [laughs]

Zierler:

Wow.

Hellman:

Because he needs to meet me, because he needs to talk to me about my materials. [laughs] It was so funny. So anyway, that was my meeting of Rai Weiss -- “Are you Hellman?” And I'm being so timid. I've got these two Nobel laureates in the office, and I'm interrupting their conversation about winning Nobel Prizes to talk to him about my materials for mirror coatings?

Zierler:

Well, that’s exactly the point. Of all the Nobel Prize winners, I can’t think of two people less affected by winning the Nobel Prize because of their commitment to the science.

Hellman:

I know. I know.

Zierler:

I mean, has Saul Perlmutter slowed down since the Nobel Prize? No way.

Hellman:

Not at all, no. And in fact, it’s interesting—my husband Warren and I, we get to meet lots of Nobel Prize winners. Warren’s not an academic; he’s got a degree in biophysics, but he’s an engineering type and designer and manufacturer of precision stuff, and he runs a machine shop. He is a really great partner to me, so supportive, interested in what I do, and doing such interesting things himself. And one of the things we talk about – I'm going to make an obvious comment – Nobel Prize winners put on their pants one leg at a time just like everybody else. So there’s people that you are just thrilled you know them because in addition to winning the Nobel Prize, they're such nice people and they're supportive and wonderful, and welcoming. And then there’s others that are not. [laughs]

Zierler:

Yep, just like anybody else.

Hellman:

Yep. And Saul, when he won the Nobel Prize, he was all expansive, and he brought literally everyone—his whole team— into the room, and he ended up sitting on the table. Everybody else had a chair, and he hopped up on the table, because somebody else had taken the chair that he was supposed to be sitting in. And you know Rai, Kip, and Barry—the first LIGO conference I went to, right after the Nobel Prize had been won. Everybody is “Rai” this and “Kip” that and “Barry” that. And this is undergraduates and technicians and staff and faculty. And everybody’s on a first-name basis. I had never experienced anything like that.

And the three of them were like, “We tried to give the Nobel Prize back, because it should have been for the team, but we realized that wasn’t really going to be helpful, to give it back, so we decided to accept it. But it really is all of yours.” And it’s just this giant love-fest. Not all of science is that way. There’s a lot of very competitive science out there, where everybody’s not all chatty and happy and passing credit off to everyone else. So it was a real delight to get to join the LIGO folks.

Zierler:

Frances, for the last part of our talk, I’d like to ask two broadly retrospective questions, then we'll end looking to the future. So the two questions – we'll do one on the science side, and one on the sociology side. So we'll start on the science side. Amorphous materials has been the through-line for you, really from Stanford all the way to now. What is your fundamental curiosity that keeps you so connected, and what remains unknown that keeps the field productive, and there’s always the need to do new stuff?

Hellman:

So, I have a two-part answer to that. One is that amorphous materials are very underexplored compared to crystalline materials. There’s a richness and an availability. I can make materials that nobody has ever made before. Sometimes, like my example of amorphous gadolinium-silicon, there is no crystalline analog. You simply cannot make Rare Earth doped crystalline Si with enough Rare Earth in it to be interesting—and the Rare Earths are the only ones that have this giant magnetic moment. There is no crystalline counterpart. And so making it amorphous has opened up a whole set of materials that have never been made before, with all new possible properties, and that’s exciting. It’s a very unexplored area.

So that’s half of it. It’s a much more wide-open field. Very few people like working in amorphous materials. Because it’s harder to do structural characterization. If you have a crystalline material, you can characterize the degree of crystallinity. The only thing I can do is try to characterize the entropy/enthalpy, and that involves pretty high-temperature thermodynamic measurement. It’s harder to directly characterize the materials. So, one motivation is just, it’s a wide-open field. And I like working in wide-open fields. I like having parameter space to explore, where I'm not directly competing with anyone else. I get to go try things and see what works. So that’s part of an answer. The other part is—it’s maybe even a little bit mischievous. I like the fact that the theorists don’t like amorphous. I mean, really, literally, theorists say this to me all the time.

Zierler:

[laughs]

Hellman:

“Oh my god, I hate working on, I hate thinking about amorphous materials.” Because they don’t have a framework. They don’t have a mathematical framework. Not all theorists; I mean, there are theorists who are comfortable with the approach you have to take to study amorphous materials, but far fewer. So it’s partly a little bit mischievous. I also like the idea of forcing an expansion of the theory underpinning phenomena. I go back to the superconductivity example—we understand superconductivity better for knowing that it happens in amorphous materials as well as crystalline materials, sometimes even at much higher temperature. That we are being artificially limited by thinking it can only exist in a crystalline material. I like the idea that I'm expanding our understanding of condensed matter physics by forcing people to recognize that the periodic lattice is a convenient mathematical tool, but it is in general not a necessary tool. I like that expansion, that we're broadening how we look at condensed matter physics. I think those are the two big-picture overviews of why I like working in amorphous materials. To be clear, I don’t exclusively work on amorphous materials—I've made plenty of epitaxial films in my career. But I do like the power of amorphous materials and the fact that the field is so wide open.

Zierler:

On the sociology side, to come back to what you quite rightly emphasized as front and center for your incoming APS leadership, and that is the urgent need to expand diversity in the field. The way I’d like to frame the question is, there’s diversity within diversity, and within underrepresented groups, the lived experience of a white woman is obviously going to be very different than that of a Black man, for example, even though both of you are historically and wrongly underrepresented in the field. To what extent is your experience and your perspective universalizable in thinking about making the field across the board more diverse? And where is there a need to sort of stay in your lane, because what you've experienced, for good and for bad, cannot be extrapolated to other people’s experiences? How might you take both of those perspectives into what is going to be a fundamentally critical point in the field as you take on this leadership position?

Hellman:

Well, the short answer is humility. [laughs] Like, don’t assume that you know what you don’t know. Something I believe quite profoundly, in all things. My experience—I don’t think my experience as a white woman is generalizable. I recently watched the Picture a Scientist movie and had one of those moments where you realize you really don’t know everything. I started by watching and thinking, “I've seen this. I know this. What’s new here?” I don’t know if you've seen that movie, but the whole beginning of it, it’s just a horrifying story. And I'm thinking, “Well, this is a horrifying story, but I have heard this story before, and I'm not sure there’s anything new here. It’s just another horrifying story.” And then we suddenly hit this moment where there was an aha moment. And this is an aha moment that I've had on a couple of occasions. It’s related to the fact that I am not a representative white woman. I'm the white woman who made it through all of this. Most don’t.

So my experience is not typical. My personality is not typical. There’s a story I often tell of early in my career—of one of my colleagues overhearing that it was my birthday [laughs], saying to me—it was in a public setting; there were other people around— “Oh, it’s your birthday! Do I get to spank you?” I was an assistant professor. And I had what I can only describe as the perfect reaction to this. He was old, and pretty overweight, not particularly attractive. I was 30 and could not have been less interested in this person. And my reaction to him was like, “Oh my god, no!” I was appalled and recoiled away from him in such a visceral way that—you know, it was I'm sure quite embarrassing for him. I was just horrified by the idea. It was revolting. And the perfect way to respond to this ridiculous comment. I've told that story before. And one time, when I told the story, a young woman came up to me afterwards and said, “You know, Professor Hellman, I appreciate the story you told, but there’s a way in which it’s kind of disempowering, also. Because we can’t all do that. We're not all that person. And when you tell that story, it just makes me feel like it’s hopeless for me. I'm not that person, and I can’t do that.” And the thing I realized in that moment—

Zierler:

That was eye-opening to you.

Hellman:

Totally eye-opening. And I had a similar experience in Picture A Scientist where a similar thing happens. There’s a jerk of a guy going around tossing his room key down next to attractive women in this room, at a conference. And they're all throwing his room key back at him. Which is funny, but also very not funny and the response of throwing the key back at him is not enough. What I wanted to say about this is, I realized—you know, bully for me. It was a jerky thing, this guy wanting to spank me. And, great. I was able to laugh this off. It was just a jerky thing he did, and I was able to laugh it off. But as a senior woman in this field, I have a deeper responsibility than just telling that story. My deeper responsibility is to make it so that women who don’t have that core ability can succeed —you shouldn't have to be that person, who can just laugh off some jerk, to make it in physics. You should not have to be that person. And so I have a responsibility, as a senior white woman in the field, I have a responsibility to not just generalize my own experiences. And by the way, another part of this story I've also realized—that’s the story I tell. There are other stories that I don’t tell, where I don’t come across quite as well. Where I've been too chicken to call somebody out. So that story is a great story, but it’s also just one story.

I have to be careful to not generalize my experiences, even to other white women. My goal is to make it possible for women who are not necessarily strong and secure, and don’t have the ability to just laugh something off—I have to make it possible for them to survive. Otherwise—there’s a reason we only have whatever we have right now, 15% or 20% of women physicists. I mean, we're picking out the ones who can survive. We have to do better than that. Black experience is quite different. Their challenges are quite different, and their solutions are probably pretty different. And I think there’s a wealth of other disadvantages —I think you're hard-pressed to make it if you're overweight. I think you're hard-pressed to make it if you have a deformity, if you have a speech defect. We have to find a way to be more inclusive, generally. I guess my mantra is I have recognized that my experience as a white woman is not even representative of all white women. We just have to do better. I'm supervising right now a Black graduate student in my group. And watching him go through the challenges of this past year is pretty heartbreaking.

Zierler:

And being isolated, on top of everything else.

Hellman:

Yep. There’s a way in which we all were affected by the death of George Floyd and the trial of Derek Chauvin, but we were affected by it; we weren’t gut-punched by it. We weren’t gut-punched by the fact that while that was going on, another Black person was shot by police. I don’t know that we can know what’s going on for everybody. But we have to find a way to get people to accept that change is needed —that the fact that they don’t think they're racist doesn't mean they're not racist.

Zierler:

Well, Frances, let’s go back to a happy place. Let’s end with the science, for the future. It’s a simple one, but it’s also complex at the same time. What’s the breakthrough that you're most excited about? With advances in materials, with advances in ideas, with advances in collaborations. Of all the things that you're working on, when you have time to carve out in the day, for the real science, the real research, what are you most excited about, and what can you accomplish in however long that might take you?

Hellman:

Well, I do think I'm going to end up circling back to what I've already said, about these amorphous materials. I think I'm really poised to make a really profound impact here.

Zierler:

I'll ask like this, then: how will you know when you get there? What’s the feedback?

Hellman:

Ah. [pause] I'll get these papers off my desk! [laughs]

Zierler:

[laughs]

Hellman:

I have half-written papers sitting on my desk with data that I don’t yet have a model for. I don’t just crank out papers. Most of my papers are unique. I have very few multi publications on the same idea. Very, very few. I tend to wait until I really have something thought through. Right now, we are sitting on a really huge result. I've talked about it at conferences, but I've not published it, because we don’t understand it. And I'm very conflicted as to whether I'm just going to go ahead and publish the data, with whatever understanding we have right now, and just let it be, or am I going to keep trying to find an answer to why the data is the way it is. And I don’t know the answer to that yet.

So, how will I know? How will I know? I'm very excited by the ideas we've already got! We have to get them out into the literature. I'm frustrated at my own inability to get these things done.

Zierler:

Well, on that note, Frances, let me let you go and get back to the papers. It has been so fun spending this time with you. It’s a great example of why what I do is really not work at all, which is why I'm so happy to do this on a Sunday evening with you. Thank you so much for doing this. I really appreciate it.

Hellman:

Thank you. It was lovely getting to meet you.