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.
Credit: NYU Shanghai
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.
In footnotes or endnotes please cite AIP interviews like this:
Interview of Paul Chaikin by David Zierler on July 22, 2020,
Niels Bohr Library & Archives, American Institute of Physics,
College Park, MD USA,
www.aip.org/history-programs/niels-bohr-library/oral-histories/46756
For multiple citations, "AIP" is the preferred abbreviation for the location.
In this interview, Paul Chaikin, Silver Professor of Physics at NYU, recounts his childhood in Brooklyn and he describes his early interests in math and science and his education Stuyvesant High School. He discusses his undergraduate education at Caltech, he conveys how special it was to learn from Feynman and Pauling, and he explains the fields that would go on to form his area of specialty, soft matter physics. Chaikin explains his reasoning to pursue a graduate degree with Bob Schrieffer at Penn, where he did his thesis research on the Kondo effect in superconductors. He describes his first postgraduate work at UCLA where he developed an expertise in thermoelectric power, and he describes the intellectual and technological developments that paved the way for the creation of soft matter physics as a distinct field. Chaikin explains what it would take to solve the many-body problem of nonequilibrium phenomena, and he describes the delicate nature of collaborating with biologists while ensuring they don’t overtake the field. He discusses his joint appointment with Penn physics and the research laboratory at Exxon, and he explains his move to Princeton, which was just starting to develop a program in soft matter physics. Chaikin describes the famous experiment that discovered that M&M shapes (ellispoids) provided the most efficient and minimal negative space in packing applications, and he explains his decision to join the faculty at NYU. At the end of the interview, Chaikin reflects on some of the remaining mysteries in the field, and he describes his interest in pursing research on self-assembly among soft condensed matters.
OK, this is David Zierler, oral historian for the American Institute of Physics. It is July 22nd, 2020. I’m so happy to be here with Professor Paul Chaikin. Paul, thank you so much for joining me today.
Sure, pleasure to do it.
OK, so to start, would you tell me your title and institutional affiliation?
So, I’m a Silver Professor of Physics at NYU. I’ve had lots of previous appointments as well. I can’t quite hold a job.
[Laughs] How long have you been at NYU?
15 years.
How has the department changed over the course of those 15 years?
Well, one of the reasons I came here was to start a center for soft matter physics, so that’s one way it’s changed, but it’s also grown a lot in cosmology and other fields and recently it has a center for quantum phenomenon. So it’s changed a lot.
I recently talked to Dave Pine. Do you shuttle back and forth between Manhattan and Brooklyn like he does?
No. He’s Chair of Chemical Engineering, which started up at—well, it used to be Brooklyn Poly and it’s now the Tandon School of Engineering.
Right. So you’re home department is physics exclusively.
Yes.
All right, Paul. Let’s take it back to the beginning. Let’s start where it all begins of course with so many people, let’s talk a little bit about Brooklyn and your parents. Tell me a little bit about your parents, starting with where they are from.
OK. I don’t quite know. I think that my mother was born in the lower east side. Parents came from the region which was either Poland or Ukraine, maybe Kiev. They found their way over here and she was born in New York. My father was born in some small town, which was either Ukraine or Belarus or Russia at the time. It switched back and forth. He came over to the U.S. I think in 1911 at the age of six.
Where did your parents meet?
I don’t know, must have been—I mean, I’m sure they met in New York, but I don’t know where or the circumstances.
How’d they get to Brooklyn from the lower east side? Was there a job, was there family?
Again, this is way back. This is well before I was born, so, I think that family spread out throughout New York. When I was growing up, we had relatives mostly in Brooklyn, but also in Manhattan, Queens, Bronx. Our family was sort of spread out.
What were your parents’ professions?
My father was a lawyer and sort of an entrepreneur, he started several companies, but that’s mostly what he was doing.
Did your mom work outside the house at all?
Yeah. Actually, she was an artist who did illustrations for several department stores for their newspaper advertisements. She worked for Mays and C&A Brenninkmeijer In downtown Brooklyn. She did these sketches that you see of clothing, especially women’s clothing.
What neighborhood were you born in?
Crown Heights. Sort of right around the corner—well, a couple blocks from Bedford Stuyvesant.
On the border?
It’d be Eastern Parkway and Bedford Avenue, the area.
How Lubavitch was the neighborhood when you were there.
It wasn’t yet, at least as far as I knew. I mean, when I would visit coming back from, you know, grad school, that’s the first time I recognized that that was happening there.
Uh-huh.
Now the most of the time that I see of it is when I go to Paris, there’s an outpost near where I stay in Paris, there is a Lubavitcher outpost. But that wasn’t a big part of where I was growing up. Certainly not.
Was it a Jewish neighborhood, but more modern or assimilated?
Yes, but there were people from, I mean, I had friends who were—the public school that I went to, I don’t know, probably half and half Jewish and Catholic, I would imagine. I never looked into it very much, but I think that’s probably what it was.
What PS did you go to for grade school?
241.
Paul, when did you start getting good at math and science? Was it even before your formal education at Stuyvesant?
Yeah. I sort of remember in third grade somehow I was reading One Two Three… Infinity, which I loved, right?
Everybody talks about One Two Three… Infinity. It’s amazing.
That was great, but I think that’s when I got interested in science and math and stuff like that. It was just great, right? Up to third grade or something, there weren’t special teachers or special classes or whatever. After I think fourth grade and after, people got segregated. I’m not sure it was a great idea, but they certainly did, right?
And what were the decisions leading to Stuyvesant? Did you just have the grades that made that the obvious choice, and why not Brooklyn Tech?
I don’t know why not Brooklyn Tech. I think I only tested—well, there was a test, as you know, right? So I don’t think it was mostly on grades. I think it was essentially on—let me think about it. I think it was essentially on an entrance exam that we took. And in fact, I think what happened was that I applied for Bronx Science, but the rule at the time was that if you wanted to go to Bronx Science, but you had to pass Stuyvesant on the subway then you had to go to Stuyvesant. And that’s how I ended up at Stuyvesant, right? Because I think that now, Stuyvesant is by far the best school, but at that time I think Bronx Science was. So that’s how I ended up there, by chance, right? I don’t remember why I didn’t consider Brooklyn Tech, but I didn’t.
How was your physics teacher at Stuyvesant?
That’s embarrassing. I’m pretty sure that I figured out in my senior year that I knew more physics than my physics teacher did. I mean, it was a strange situation. He was a strange guy. He was—yeah. And he really didn’t know very much physics. And he had tried, as I recall, to—he started a company that was gonna build a safe that you couldn’t burn your way through with an oxyacetylene torch. So he patented—there was money in it and stuff like that, and then essentially when he had the company going, somebody pointed out to him that if you stuck an iron rod in the flame of an oxyacetylene torch, it would just go right through essentially anything.
Paul, how did you come to know more than your physics teacher? Were you doing a lot of outside reading? How did you learn physics?
I learned physics by reading. Also, we had, at the time, we had a cyclotron committee, a cyclotron group at Stuyvesant.
Really?
Yeah. The guy that taught us electrical engineering sort of. I don’t think it was called that. No, it was called electricity shop. I don’t think it was even electronic shop, right? He started a group that got some funding from the Board of Education or something to build a cyclotron. And so we set out and we actually made a small cyclotron, which I don’t think ever worked during the time that I was there, but did a couple years later.
What were you trying to find out at that early age with a cyclotron?
No, it’s just the idea of building a cyclotron, that you really—I mean, it was right because, you know, you learned a lot trying to build a cyclotron, right? And it actually, you know, it was very good because ever since then, my impression of doing science or doing research, especially experimentally, is that, you know, there is physics in essentially everything you do with it and you absolutely have to understand that.
So if you’re doing an experiment in the lab, you don’t just have to know the physics of that experiment. You have to know the physics of how every single piece of apparatus you’re using works, right? Otherwise you really won’t know what you’re doing. You really won’t know what you’re measuring and stuff like that. And even the little work we did on the cyclotron, right? Putting it together, wiring it up, figuring out how to do it and stuff like that was all that. It was all OK, you thought about it from—or at least I did—from a viewpoint of physics, right? Like, I imagined the electrons going through circuits, I imagined, you know, what the magnetic field was doing, and for each component of it.
And that’s the way I try and tell my students they should do science, right? That if you’re an experimentalist, it’s not just you wanna know the physics through the experiment, you have to know the physics of every single part of the experiment.
When it was time to think about college, were you thinking specifically about physics? You know you wanted to pursue physics as an undergraduate?
Yeah, as a matter of fact, I remember the Caltech recruiter trying to convince me I should do biology. Actually I did some—my earliest research wasn’t in physics. My earliest research was for when I was doing some science fairs was when I was in public school—that’s where I really started this—and I used to work in the Brooklyn Botanical Gardens, they had a research garden there. And we did some experiments—I don’t even remember how I got involved in this—but first I did a science fair project on tropisms, flowers growing towards light and roots growing one way or another.
And I remember doing it with cigar boxes that I used to get and cellophane so you could see where the roots went. And the next year, we did work on growth-promoting substances. And this was done at the research division in the Brooklyn Botanic Gardens and what we were doing is spraying flowers, weeds, beans essentially, mostly beans, with these growth-promoting substances, which, you know, we’d measure how fast they grew with this dose and that dose with this compound and this other compound. And we won first prize I think for that. I don’t really remember, but we won something for it.
But I think what really got me is the stuff was called 245-T.
That’s half of Agent Orange.
That’s Agent Orange, exactly. And we didn’t know it, right? So, essentially we were doing research on how to defoliate Vietnam, right? And the way of course it worked is that they would grow too fast. It kills them, right? When they overdosed it. So that was my real entrance into science.
Oh boy. Oh boy. Paul, where else besides Caltech did you apply?
I don’t remember. I think Yale, a bunch of places. Yale, Colombia, yeah. I don’t remember all of them.
Caltech must have seemed like a million miles away for a Brooklyn boy. How did you even get it on your radar?
I was pretty young and pretty naïve. And in fact, what was really cool was the Caltech catalog—when you open up the Caltech catalog, what they have in there is pictures of the mountains and the beach and surfers, right? So I thought compared to Colombia or something like that, right, this was gonna be a real party school. And it wasn’t. OK? But it was certainly different.
Did you have the sense—
And I did really want to somehow get out on my own, but as I said, I was pretty young and naïve in those days.
And did you have a sense of who people like Feynman and Gell-Mann were before you got to Caltech?
No idea. No idea. I knew Caltech was really good, you know, it was really well known and stuff like that and was really good in science and stuff, but for example, I—no, I didn’t know Feynman and Gell-Mann—the first week I was there—so I took a bus across the country. Went to Caltech and I was there and things were a little bit strange and then I went to—they had a faculty party.
So the house I was in had faculty over for some party right? And what I really distinctly remember is I hadn’t noticed before this that the accents were different, that the California accent was not the same as the Brooklyn accent, right? Suddenly across the room, I hear somebody speaking absolutely perfect unadulterated English. And that, of course, was Feynman. And that’s the first time I met Feynman.
That’s hilarious. He must have learned how to stop speaking Brooklynese at some point himself.
No. He cultivated the fact that he didn’t lose his Brooklyn accent. This is before he had a Nobel Prize, so he was well known to people that knew him well, but not to the general public. So that was an experience.
What were your impressions of Caltech when you first arrived and you realized it wasn’t on the beach and it wasn’t surfing all day long?
Well, I think the impression I got wasn’t right away, but the real impression I got over years—which started as soon as I got there—is that there were people that were orders of magnitude smarter than me. Not, you know, just a little smarter than me, OK? I never thought I was tremendously smart. Like, I didn’t graduate Valedictorian or anything at Stuyvesant. I did OK, but I wasn’t the top guy.
But I got to Caltech and there were clearly people, students, that were smarter than me and there were clearly, you know, faculty and not just smarter, but students ten times smarter. And there were faculty than were ten times smarter than that and then there was some faculty that were ten times smarter than that, right? So I really got humbled by it, right? At some point you start thinking, “Well, what difference could I possibly make?” Right? Feynman could do stuff in his head that, you know, would take me a year to do.
Feynman wasn’t the best, by the way. The guy that I was most impressed with who I’m still most impressed with is Pauling. Pauling really, really had—Pauling was 20th century chemistry. And also a very unusual guy. I mean, this guy—one of the things I really took from Caltech is these were really unusual guys and we used to have them over. The way I got to know most of them—Feynman I got to know a wee bit differently—but most of them came to dinner at the house and they’d tell us their stories, right?
So I’ve heard Feynman’s stories about Los Alamos many times. In particular, many times when—what I remember is specifically he used to tell these stories that I’d heard them maybe two or three times by the end I was there, but his wife would usually accompany him, must have heard them thousands of times, right? And she would, of course, fall asleep while he was telling us all these great adventures. But, I mean, Feynman was incredible.
Pauling was incredible. What I remember about him especially is—so I was working, at the time, I joined the lab work around superconductivity and so I knew a little bit about superconductivity. And Pauling came over—and he was talking about stuff, he talked about nuclear physics, he talked about stuff that you wouldn’t expect him to talk about ‘cause he’s a chemist, right—and one of the things he said, “You know, nobody really understands superconductivity except me.” This is, of course, already after the BCS theory and stuff like that. And he started explaining this, what he called, resonating valence bond theory of superconductivity, right?
And, you know, so I heard this and OK, it was clever and stuff like that. I didn’t think that he really had a tremendous insight into superconductivity or at least about superconductivity that I knew at the time, but of course, was it 20 years later, there wasn’t even high Tc superconductivity then, the valence bond theory was resurrected by Fyl Anderson, right? Who actually, you know, gave credit to Pauling for it. So Pauling was really something.
You couldn’t limit him to biology. He was too smart.
You couldn’t limit him to biology or chemistry, right.
What class did you actually take with Pauling?
He taught in freshman chemistry, which was taught by somebody else—he taught maybe a term or a fraction of a term on quantum chemistry. So the first time I saw quantum mechanics was from Pauling. He was amazing. I mean, the thing that got me, which was true throughout his life, is both he and Feynman love numbers. They absolutely love numbers.
So he would come into class and he’d, you know, teach us—he was teaching something—and then there’d be some reaction that he’d write down and he’d write down the reaction. And he’d write down the reactions rates and he would write down the reaction rates for the reaction to six significant figures, right? And it’s like, “What in the world is this?” And then if you go and check, they were right.
Yeah.
You know, the few times anybody checked on it, it was right. And the reason why is essentially both he and Feynman, you know, when they’d see a number, they’d play with it. They loved playing with numbers. And so they’d figure out that, you know, “OK, I can get this number by doing this decomposition of it Divide this by this.” Or, “Oh, it’s the cube root of this plus one.” And you’d ask them about it and they’d give you something like that.
I always remember that and I remember that—let me think how much later—I think 30 years later Pauling gave a talk on quasi crystals at a March meeting of APS, I think it was in Oregon, and he was quoting the ratio of the diffraction spots to get what the increments for the irrationality—or in his case he claimed it was rational—the rationality of the fractions was and he’d quote them just off the top of his head. Not written down. He’d quote them into six, seven places, right? Really amazing.
And it was also interesting because that talk—although Pauling of course is a genius and everything else—his quasi crystal stuff wasn’t right. [Laughs]
It would take Paul Steinhardt a little later to solve that mystery.
Exactly. Actually I knew a lot of these guys. Steinhardt is twice my colleague.
At Penn and where else?
Princeton.
Uh-huh.
So he sort of helped hire me at Penn and I helped hire him at Princeton.
Oh, very cool. You were part of that committee to get Steinhardt over to Princeton.
Well, it was just, you know, they asked—when the cosmology group wanted to get Steinhardt but they walked around, they said, “Oh, you know, here’s this guy we’re looking at.” And I said, “Oh, no. You don’t have to convince me of that.” Right? And so, you know, I helped lobby for him and stuff. But I mean he didn’t even need it.
Paul, at what point did you decide to focus on physics at Caltech? Was that from the beginning or you were on the fence for at least the first year?
From the beginning.
Mmhmm.
Despite the fact that the recruiter tried to get me to go into biology—because essentially, at that time, everybody that was going into Caltech wanted to do physics.
Right.
I shouldn’t say everybody, but the majority wanted to do physics. And so they were trying to get people they [thought] were [a] really good in other fields as well.
And so you entered, what, in 1962?
I guess. Yes.
So what were some of the most exciting things that were going on in the department at that point?
I don’t know. I mean, that was the time when, as you said, Gell-Mann was there working on—I don’t know exactly the history of it, so I don’t know.
But I guess my question is as an undergraduate, did you have a vantage point where you had a sense of what the professors were working on?
Not much until the later years. And, you know, pretty soon I specialized in condensed matter or solid state physics.
But I’m curious, it was called—condensed matter was not a term in use at that point, right?
No. it was super fluid helium. I’ve thought super fluids and superconductors were really cool.
But I understand solid state if the things that you’re studying are solid, but for soft matter, what was the name for that? You couldn’t call soft matter physics solid state because it’s not solid, right? What was the name that was in use at that point?
Well, there wasn’t. I mean, what we call soft condensed matter now grew up sort of in the 70s, the modern version of it. So at that time there really wasn’t soft condensed matter. There were, of course, people doing polymers and there was liquid crystals [from a] century back so, no, goes back much further than that, goes back to soaps, right? With the liquid crystals. And so it goes back millennia. But I don’t think there was the field of soft master and it wasn’t done in physics. It was mostly done, at the time, in the engineering department and the chemistry department. Polymers were big. But it wasn’t really done in physics.
It took probably—well, it depends how you looked at soft matter. I could say it took Perrin and Einstein to get people interested in soft matter physics with the first Brownian motion of colloids right? And atoms in motion. But the modern era sort of was Pierre Gilles deGennes and Sam Edwards, right? And liquid crystals and polymers and [??] matter.
And what convinced you, as an undergraduate, to focus on solid state as opposed to all the exciting stuff that was happening in cosmology, theory, particle physics? What was it about solid state?
Well, cosmology, there wasn’t that much yet because cosmology, if you—I mean, there were some things—but nothing like the explosion now because of the instrumentation that we have now, which really took us beyond. I mean, all that stuff was fascinating, but I actually liked working with my hands a lot, not just doing calculations and stuff. I worked with my hands. And what I really liked is the fact—even early on, you could do this in those days with hard condensed matter physics as well—you could decide, “Oh, I don’t understand something. I’ll go into the lab, I’ll set up an experiment, and by tomorrow I’ll know the answer.” You could do that and you did that.
And it’s a joy working with the experiment—first of all, it’s a joy working with your hands. And then seeing stuff that is amazing right before your eyes. You know, the next day after you design an experiment is just amazing.
The first time I saw super fluid helium, amazing, right? Look in the Dewar you cool down, you go pump it down (to lower the temperature) suddenly the surface becomes absolutely smooth. There’s no bubbling. It’s obvious that the heat transport from the boiling and stuff is now nothing the same. You see fountain effects, you see helium climbing over things. You see it with your eyes and here’s a quantum phenomena, completely a quantum phenomena that you’re just seeing with your eyes. Not splitting atoms, not from a cyclotron, not from particle tracks and stuff like that. You’re really seeing it with your eyes. And doing experiments like that were really cool.
And you had to do all sorts of stuff back in those days, you had to blow your own glassware. So we did experiments—there was an experiment that Feynman—I mean I talked with Feynman several times, but he wouldn’t know me, but he was very good buddies with my advisor, Jim Mercereau who invented the SQUID, the superconducting quantum interference device—and Feynman, even in his lectures, he talks about—in his lectures it’s just a water sprinkler where when you turn on the water sprinkler, it goes one way. And he poses the question, “What if you now put it in the fluid and you suck on it instead of blowing through it? Which way is it gonna go?” And he got interested in this problem. And Feynman was interested in that problem and suggested since there was a low temperature lab, he said, “Well, why don’t you do that with a super fluid and see what happens?”
So there was a little thing called the Kapitza spider, which was you flow by heating it up. You’d flow super fluid through it, and you could make it so that the arms come out like that, OK? Oh, you can’t see because of my (screen) background, but in any event, like that so it rotates with the arms. And so we blew one of those. That is we made the glass gadget and we made something be like that and sat it on a pivot and put it in superfluid and watched it go. That was pretty cool.
And you did all this stuff, you know, currently you have craftsman, right? You had to get things to work. You had to put together your electronic instruments. You had to put together your physical instruments to really get your experiment done. I liked it.
What professors did you get close with at Caltech?
Essentially Jim Mercereau. There were teachers and stuff like that, but my research really started with him in my junior year.
What was he working on? What was his research at the time?
Well, it was, like I said, he took over the low temperature group and the low temperature group traditionally had been the superfluid group, but he started work on superconductivity and his research was mostly on the Josephson effect and on the technological applications of it.
So if you look at the Feynman lectures, I think one of the last Feynman lectures, which is in I think volume 3 of the Feynman lectures in physics, is a whole chapter—I think it’s a whole chapter—devoted to SQUIDS. To the fact that you could make a device that would see quantum interference. Essentially, the SQUID is two Josephson junctions in parallel and it functions essentially like a two slit diffraction experiment.
But what you do in order to get the diffraction is you change the phase, not by your position, but by changing the magnetic field through the apparatus, through the area of the apparatus. And that was an invention of Mercereau, Silver, and I forget the other one off-hand, but anyways, Mercereau was one of the inventors of that. And Feynman loved it, and so that’s why Feynman and Mercereau were big buddies.
Did you have a senior thesis?
I didn’t, I don’t think. I did lots of experiments, but I don’t think we were doing senior theses in those days.
And so when you were thinking about graduate school, you were fully in on solid state soft condensed matter at that point. You knew that—
No, no, no. There was no soft condensed matter.
Right, but I mean solid state, right? You knew you wanted to pursue that.
I pretty much wanted to do solid state, and in particular, superconductivity.
And superconductivity. What kind of advice did you get on where to go to school? Did you think about staying at Caltech?
Somewhat, but essentially what happened is I just asked Mercereau, “Where should I go?” And he said, “OK. Here are some places which are doing really good things in this area in superconductivity, things like that.” He listed them for me. He said, “Where do you want to go?” I said, “I’m thinking of these places.” He said, “OK. I’ll call up my buddies there and get you in.” [Laughs] I think that’s sort of the way it happened.
What year did you arrive? This would have been fall of ’66? You went straight through?
Fall of ’66. Yeah.
And what was your sense of Penn when you got there? What ways was it different than Caltech just generally, in terms of how the physics department operated?
Interesting. Well, actually one of the reasons I went I guess is because Bob Schrieffer was there, right? So the connection with superconductivity was really good, but the professors there—aside maybe from Schrieffer—weren’t as well-known as at Caltech. But I mean it was a really good climate and they were really interactive.
And the first year or so, mostly I was involved in courses, and some of the courses were great, right? There were people like Herb Callen, who was teaching his thermodynamics course. I got thermodynamics purely from the thermodynamics viewpoint from him. There were some teachers that weren’t very good, who I won’t name. [Laughs] But in any event, overall it was a good experience and actually—what was surprising to me is that the Caltech education I guess was pretty good because I skipped some classes and so I took—at Penn, supposedly you had to take courses for two years and then you took a three day exam. And they let me take it after the first year. And so I finished my course work in the first year and then could go on to research, which was nice.
But afterwards, I took courses that weren’t required courses and I had some really good people. I always remember Doug Scalapino’s course and Brook Harris’s magnetism course and there were, you know, lots of really good people there who I got to respect and who many scientists who, you know, I still look back at their papers and what they’ve done since, what they did before and what they’ve done since.
But I started working for Tony Jensen, who was a young assistant professor who was doing really cool stuff. And he was really amazingly lively and stuff. And he was big buddies with Alan, with Heeger. So I got to know these guys. They were young. That was really cool because the young—Tony, younger than Alan was—but they were, you know, sort of bright rising stars and it was interesting to see how science was done, how really new, important discoveries were made.
Penn was definitely in growth mode during these years, right? It was really rising in its stature.
Yes, it was. Penn has always had really good people. I don’t know why, but somehow some of them go, some of them stay. Penn’s always been really good. I mean, it’s never been a weak department. It’s always had really good people and some really excellent people as well, right? I mean, Santa Barbara grew from Penn, if you want, in the Physics department. But the faculty at Penn is really great and when I was there on the faculty, the people I also interacted with were great and maybe even be stronger now than ever at Penn. In soft matter, they’re incredible.
And you mentioned earlier that the term and the field soft matter condensed physics really got started in the 70s. Was that starting to come into focus during your years at Penn or would that be afterwards?
You mean when I was a graduate student at Penn?
Right.
I was a graduate student in hard condensed matter. So no. There was essentially—I don’t remember anything like soft matter when I was a graduate student there. I mean, the closest you got to soft matter was conducting polymers or conducting organics, right? So they were organic, but it was the electronic properties that were interesting, quantum properties that were interesting, not the mechanical properties. Not the soft dynamics.
And what were you doing in the labs, as a graduate student? What were some of your projects in the labs?
So the first thing I worked on—well actually my thesis was on the Kondo effect in superconductors. It was based on the premise—which I think was sort of my own, but may not have been—which was that the best understood many-body problem at the time, maybe even now, was superconductivity. So why don’t we use superconductivity in order to study other many-body problems. So I think the title of my thesis is Probing Many-Body Problems With Superconductivity.
And so one of the easy ways of doing that was it was the Kondo effect, which was not yet solved at that time. It was known what it was. It was known that the resistance went up and why, Kondo’s work, but it wasn’t anywhere near solved and so it was lots of activity there. So a lot of my research was on trying to understand Kondo scattering, Kondo impurities in superconductors, which you could essentially see by several ways. One of them was by how much the transition temperature of the superconductor was reduced when you added the impurities, you could find out what the Kondo scattering was doing to the superconductivity. And I know Penn’s learned something about what the Kondo effect—the state that the Kondo effect was going through was.
And then while I was there, at Penn, I spent a year at San Diego. Originally I got sent out by my advisor. My advisor, Tony Jensen, was one of Bernd Matthias’ students. I don’t know if you know Bernd Matthias, but Bernd Matthias was a very well-known physicist, material scientist. At the time, he had three positions. One was at Bell Labs, one was at some national lab, Los Alamos I think, and the other was UCSD, San Diego. And he was my grand advisor, one of my grand advisors.
So my advisor sent me out to do a Kondo experiment, a simple susceptibility experiment on an apparatus that one of his fellow graduate students, Deiter Wohlleben, had at San Diego. And I got out to San Diego and I was supposed to be there for three months and I ended up spending the year there. Partly because—this was I think my second year at Penn or maybe my third year at Penn—partly because I discovered San Diego, actually there was some good science going down there and the weather was fantastic. Best weather I’ve seen any place in the world.
But really what happened is I did these susceptibility experiments. They were incredible because you’d have to get on top of this apparatus, climb on top of the magnet on top of the dewar, and sort of adjust the apparatus for every measurement that you did. So you’d go up and you’d adjust things with a wrench and you’d go down and you’d take the measurement and then you’d climb back up and you would adjust it and you could take the measurement, one of those things.
But then I went—somehow I made contact, I forget how—with Ted Mihalison who had a job at Gulf General Atomic which was down the road, and he had set up just what we really needed for the experiments on Kondo effect, he had set up an apparatus for doing low temperature tunneling experiments with a magnetic field, and so somehow we made contact with him and so I got supported as a consultant for General Atomic and I was one of, I think, three consultants that year.
One of them was John Bardeen and one of them was Hans Bethe, and those guys got, I think, $3,000 a day and I got, what I figured out—I got like 25 cents an hour.
That seems about right.
It seems about right in terms of the contribution, absolutely. But anyway, I’d spent the year there and then that’s when I got involved in proximity effect tunneling and things like that. So that was actually very good because it broadened what I was doing from simple measurements, like resistive measurements, to susceptibility measurements and transport and tunneling measurements and things like that. Actually it was very good, it really broadened the experiments I could do and a lot of them I continued doing essentially until I went into soft matter. Well, essentially until ten years ago.
Out of all of this work, Paul, how did you settle on a dissertation topic?
Well, as I say, I saw what was going on. At that time, Alan Heeger was somewhere in his Kondo effect days, he was still making progress on that, as was Tony Jensen. And I wanted to, you know—I don’t think I was actually given this particular project by them, but as I said, I got the idea that superconductivity was the best understood many-body problem. And so since they were interested in Kondo physics, we sort of used superconductivity to study the Kondo effect and that’s the way it came up.
And then when I went to San Diego to do these experiments, the first was a pure Kondo experiment, and then later we were doing superconductivity with Kondo and then it became tunneling in induced superconductors with Kondo impurities, and that’s sort of what the thesis developed into.
Do you remember who else was on your committee?
Actually, I don’t. I can look it up. I mean, my advisor actually died before my thesis was done, so Alan Heeger is the one that signed my thesis.
What was it like to work with Alan?
Alan was terrific. Even before my thesis was done, Alan had decided to go into this new area of organic conductors. He was influenced a lot by a postdoc that he had, Tony Garrito. Tony sort of brought the problem to Alan and Alan sort of developed it into some exciting, and controversial, science which is now a well-known field that stands on its own. But we didn’t know anything about it. I mean, Alan, of course, knew more than anybody after he started looking and he understood that there was really interesting, deep, good, many-body physics in it.
And so for I think a year and a half or maybe two years before my thesis, Alan organized the group to meet every week at night and educate ourselves in the physics that we needed in order to attack this problem, the problem of conducting organics. And in those days, conducting organics were conducting organic crystals. So they were molecular crystals, which had some slight conductivity depending on how you doped them and things like that or what materials you made.
Alan was terrific. It was really terrific because Alan let everybody talk. So he would suggest papers and you’d go back before the week came when you were to present and you’d research it and you’d figure it out and you’d explain to everybody in the room, I guess the group was maybe ten people, something like that, and you’d explain to everybody what you thought was going on and other people would chime in and say, “That’s not right,” or “This is the way you should look at it,” and the thing that was really, really good is there was no such a thing as a stupid question. Everybody really learned because you had the opportunity to ask anything. You wouldn’t assume that people understood everything and if you said something that people didn’t know, they would just stop you and you’d have to explain it.
So it was really good—both from learning from what other people did and from the way to do science—which is—it really should be there are no stupid questions—and the fact that you really tried to explain this so that everybody would understand it. And of course the only way you really learn anything is when you have to teach it. And so it’s a terrific experience.
And so for at least a year before my degree, and then for a half a year when I was a postdoc with Alan, we had these group meetings that were terrific. And they were always preceded by going someplace for dinner. So it was great.
Did you do a postdoc or did you join UCLA direct as a member of the faculty?
So what happened is I couldn’t graduate until—I couldn’t get my degree until December of whatever year it was, ’71 or ’72, I forget, because I just turned 25 in November of that year. So I got my degree in December and I already had the offer from UCLA, but I stayed on for another six months or so working with Alan.
Were you on the market? Were you looking for opportunities or they just sort of recruited you?
Well, I interviewed at a couple places. I mean, Alan helped a lot. I mean Alan was, by then, well-known for many things. And so I interviewed at several places. It was really strange. Anyway, I interviewed at—the two places I remember best are I interviewed at Santa Barbara and at UCLA and again, I’ve been naïve all my life, but anyway, I went to—Santa Barbara was just starting there. Santa Barbara—the physics department was in a trailer ‘cause they hadn’t built a physics department yet.
I had two great people that were recruiting me which were Doug Scalapino and Vince Jaccarino both amazing physicists. And I looked at Santa Barbara and the campus wasn’t done yet and stuff like that, but I looked around and I said, “You know, this is just an absolutely beautiful place.” You know, you have the ocean right by. You’ve got the mountains. I said, “It’s an absolutely beautiful place. I could probably live here for a weekend without going crazy.” So I said, “If I’m gonna go here, I’m gonna spend my time driving down to the biggest city around,” which was Los Angeles, and I had an offer from UCLA, so I said, “Well, I may as well go to UCLA then.” And that’s the way it is, because unfortunately, being raised in New York, I just find it hard to be in a place that isn’t New York. Or isn’t enough like New York that I can take it.
And, of course, Santa Barbara became an absolutely incredible place, largely it turns out due to those guys, especially Doug. Anyway, one of the reasons that I ended up at UCLA is that Alan had a big buddy there and I had to be with him. Anyway, we got off track.
Well, my question was, everybody talks about the value of a postdoc before joining a faculty, right, but did you feel like, in a miniature version, you actually did that postdoc at Penn by staying on?
No, what I actually feel is that—as I said, my thesis was on superconductivity, but since Alan organized this group to go into this new field, I also started doing experiments in that field. And so I really felt that the year and a half or two years before I got my PhD were more like a postdoc because I really got into a different field and I started doing different experiments.
That’s very interesting. So you kind of built a postdoc right into your graduate work.
Yeah. I actually remember one of the things I had to present was - somebody in Russia—Shchegolev—was sort of Heeger’s counterpart in Russia for a while, had done an experiment on thermoelectric power, the Seebeck effect. And I had never heard of this and nobody else in the group had, and the claim was that he was measuring some stuff that was interesting and so I think Alan assigned me—either Alan assigned me to give a report on that paper or somehow it came up and I decided to do it.
But what happened is I read the paper, it sounded interesting, and so instead of just presenting the paper, I just said, “Oh, this looks like a really easy experiment.” And I went into the lab and I set up the experiment. And indeed, it was an easy experiment. And I got the results that this guy showed and actually made it—did the experiment a completely different way than it was done there and it turns out that’s what led to—I used to be known for thermoelectric power for maybe, I don’t know, ten, 15 years, 20 years, whatever. I was considered one of the world’s experts on thermoelectric power and it all started with that. That’s as good postdoc training as you can imagine.
Yeah. It was as much a learning experience that you realized you could just jump in and figure something out and do it.
Yeah, and also I had no experience before with these materials. They were ridiculous materials if you grow up doing superconductors and evaporated films and stuff like that. These are tiny little crystals, right? You can barely see them with your eye. In order to make contacts with them, you really need good hands. Anyway. And in those days actually—well, I don’t think I did at that time—but in those days you used to be able to smoke in buildings and even in laboratories. And I remember one time, what I actually did is I used to smoke a pipe. One time I mistakenly mounted up a piece of the tobacco ash as a crystal, but it didn’t conduct.
Good thing. [Laughs]
I read the paper and I just did the experiments. Then I was sort of hooked. “Hey, this field looks easy.” It wasn’t.
Now, did UCLA give you the support and funds you needed to set up the labs that you wanted to set up when you got there?
So, I never complained. Actually no place I’ve ever been has given me lots of money to set up a lab, but it’s always been—I’ve never hurt for funds to do what I wanna do, but it’s usually because I really like to make things. And so it’s typically, you know, I can live with what’s—if there’s something around, I can usually use it. I can build on it. My start-up funds at UCLA I think were 30 kilobucks and there’s enough to buy a couple nanovolt meters, a couple of function generators, a couple of lock-ins maybe and that’s it. I inherited a lab, though. There’s a lab from Hans Bummel who moved to Konstanz as, I think, maybe even head of the University of Konstanz, and he was a famous man for doing ultrasound experiments. And so I got some dewars and pumps and stuff like that.
One thing that worked, I remember the first year I was there, most of the time I spent fixing pumps, soldering together apparatuses, stuff like that on my own. Actually I really loved that. The problem with growing old, as I am now, or even growing older, is you just absolutely don’t get in the lab as much anymore. You feel guilty, or at least I do, feel guilty going to the lab because I figure I should be doing something else. I should be figuring things out, my job should be figuring things out, writing proposals, and doing stuff like that. You know, making sure my students are doing well, but I really, really love going to a lab.
Sometimes I’ll go into a lab—I’ll decide, “OK, I’m gonna do something,” and I will just go into the lab and I love working with my hands with tiny things, with microscopes, with, you know, binocular magnifiers or stuff like that, figuring out—as I say, you really have to understand physics even to figure out how to mount things. To realize what all the thermal conductivities are, what the impedance is, where your energy’s gonna go, where the heat links are gonna kill you, everything like that.
And so you have to do it all the time and you have to figure out clever ways of doing it and I love it. And so the first couple of years, when I was an assistant professor, I had built every single apparatus in the lab, I had built it myself. I soldered things together. I did leak checking, I did everything. And I got them to work and they were pieces of junk. They just barely would work ‘cause I didn’t care about how beautiful they were or whatever like that.
And then when I got it to work, I’d give it to a student thinking, “Oh, this student’s gonna make improvements. This is going to be much better next year or the year after.” And six years later I’d go back and the apparatus would still be working and nothing would have changed on it. They hadn’t done anything. And so I have racks and racks—I still have them in my lab—racks and racks of things that I welded and soldered and stuff like that and I absolutely used to love doing that. And I probably could do it now, I’m not sure, but I would consider it a waste of time and I would hate myself for doing it because I’d be having so much fun and if you say, “Jeez, I’m just having fun doing this,” I should be doing something else. I should be figuring something out.
And so Paul, in those young days at UCLA as you were setting yourself up, what were the big research projects you wanted to get involved in? As you emphasized, what was most fun to you at that point? Was it thermo power first or that came later on?
That came—well, I think I started doing it immediately when I went there, but I was doing superconductivity also a lot. I mean, this idea of using superconductivity to study other phenomena, I thought, was a really good idea. So for example, I mean one of the reasons I got this idea is that I always thought the real proof of the theory of superconductivity at BCS were these experiments that I think—just a sec—anyway, these tunneling experiments where they could unravel the phonon spectrum and you could measure the phonon spectrum then better by doing superconducting tunneling than you could by neutron scattering.
And the fact that from the spectrum that you calculated from it, you could feed back into the Eliasberg equations and exactly get things to work was essentially, I thought, the best proof that the theory of superconductivity was right. And it also gave you the best information about the phonon spectrum and the electron phonon interaction and other materials properties. And so I thought, “That’s really cool. Why don’t we take normal materials and induce them superconducting and study their electron phonon’s interactions.” And so that led to lots of work that we did on the proximity effect. And so that was another big thing that I was doing, but it all was under this rubric of using superconductivity to study many-body of things.
But the thermo power was also really neat because—so UCLA was a terrific experience. , Fyl Pincus was there—the real reason I got a job at UCLA is that Fyl Pincus and Alan Heeger had always been buddies. They’re still buddies. As a matter of fact, I think Heeger lives in the apartment above Pincus’s now in Santa Barbara. But they were buddies since Berkeley, since they were graduate students at Berkeley. And so Fyl is really the one that got me to go to UCLA. And Fyl is a little bit nuts, but he’s a great guy to talk physics with and he knows a lot and stuff like that, and he sort of wouldn’t let me get away without—he’s a theorist—and he wouldn’t let me get away without doing some theory.
And so this was terrific because I learned that you don’t have to—I shouldn’t say this—you don’t have to be a genius to do theory. It’s just a matter of having an idea and doing the best you can to figure out where it leads you. And so when I got to UCLA—no actually, sorry, I’m wrong. Actually I did theory with Alan before that.
One of the things I did—I did a couple of papers when I was at Penn with Heeger on theory. I forget who else, but in any event, it was on excitonic polarons and the guy—so that’s another reason I ended up at UCLA—the real hero of polarons was Ted Holstein, who you may or may not know, but he formulated the Holstein-Primakoff transformation, but essentially he was sort of the senior theorist at UCLA. And his work on polarons inspired what Alan and I had done on excitonic polarons when I was at Penn. And even Fyl, and a little bit Ted Holstein, sort of gave me this impression that it was OK to do theory, right? That you could do stuff like that.
And so I did some stuff with Fyl’s graduate students on thermo power and those papers have actually lasted, they’re cool and still cited. Anyway, so UCLA was really good because it was really an amazing group. All the condensed matter people were on the same floor in the same corridor. And Holstein would, every morning, go from one—work his way down the hall, go into everybody’s office, ask what you were doing, and then let you know whether he thought it was worthwhile. Whether he thought it was complete nonsense, or whether what you were doing was interesting. So it was really terrific.
And Fyl, of course, had already almost made the change when I got there. He was still doing some electronic physics and still interested in it, but Fyl was the disciple of Pierre Gilles de Gennes. He and de Gennes were big buddies. And so Fyl, I think, is the guy who really brought soft matter to the US from France, from Europe, and Fyl tried for 10 years to get me interested in soft matter. And I didn’t budge. I just didn’t do anything with it for sort of the first ten years.
But nonetheless, I had a great time at UCLA and I learned a lot and it was a terrific time.
Did you continue on with the Kondo effect? What was the status of that field by the time you had gotten to UCLA?
By the time I’d gotten to UCLA, well, it hadn’t yet been solved, but it was being solved.
Right.
By Wilson and Anderson. So it was still interesting, but I mean—and there’s lots of work that was done after that—but somehow I wasn’t—I don’t remember why I wasn’t particularly interested on it anymore. It seemed to me that this was a field—so I got more involved—most of the time I was at UCLA, the electronic physics I did was on the organics. And if you like, the reason that I found that much more interesting is this was a field that was absolutely just opening up. It wasn’t like, “OK, here’s the Kondo effect. We have 10 years’ worth of research and we understand this and we understand that and here is actually a theoretical solution to it. We now know what it’s gonna do.” It was, “Well, we have all this vast class of new materials, they’re showing different effects.” This is when charge density waves first and then spin density waves were coming in. There was always the possibility of them being superconducting, but they weren’t.
And the field actually rested on this idea by Bill Little, which never worked out, that the organics may give you excitonic superconductivity at room temperature and above. So that’s what propelled a lot of the field and the science was fantastic on its own with the discovery of new effects.
So I think it was the fact that it was a new field and there were actually completely new things to be discovered that took me away from Kondo and maybe even superconductivity to the organic conductors.
And what were some of the technical advances in organic conductors that you saw really providing a lot of opportunity to move the field forward?
Well, I mean, part of it, of course, was the chemistry and part of it, where almost everything in condensed matter, was new materials, right? When you discover you have a new set of materials, you have a new playground. And unless something is wrong there, there has to be some really interesting new physics almost anytime you get new materials.
So there was that. There was the observation of new phenomena, and the fact that a lot of it was new physics in the sense that there were lots of nonlinear effects and there were things which were not equilibrium effects that came in. But I think there were all sorts of questions that resonated with new models and new physics, with dynamical models, with questions of dimensionality and incommensurability with the density waves, with the dynamics that you get from nonlinear effects and couplings. So I think that’s what made it really interesting.
And, of course, then, the idea—the thing that really made the field blossom much, much more so than it had before—was the discovery of conducting polymers, right? Because the conducting crystals were essentially really, really interesting physics, but nobody ever imagined that they would be applicable practically, that they would be of any industrial use whatsoever because they were fragile, they were small, they were temperamental. You never imagined that you could process them easily. At one point they weren’t even superconducting, then they became superconducting and interest rose again.
Then it turns out they showed everything. So, in fact, I think that the organic conductors and the studies that all of us in the field did through the mid-80s, early 90s, were, in terms of fundamental physics, as interesting as anything that’s been done on high temperature superconductors, maybe more. We essentially saw all the effects that they’re seeing. We saw competitions between spin density waves, charge density waves, charge order and spin ordering. The organics even had Quantum Hall effect in them. They even had new versions of Quantum Hall effect in them. The only thing was the highest Tcv was 12 degrees. The highest superconducting transition was 12 degrees. If it had been 120 degrees, we’d still be studying them and everybody’d be excited, but at 12 degrees, nobody cared. But the physics was really great and remains great.
Paul, were you thinking at all about some possible commercial opportunities or aspects of your research?
Well, yes, I thought about it, but for instance, I was much more interested in the science. If I thought there some possible application, I would do it. I spent some time thinking about whether we could do anything with thermoelectricity, you know, for power or cooling and stuff like that?
Yeah, I’m curious about that. With thermoelectricity, I mean, would people in industry approach you about possibly pursuing commercial viability plans?
Yes. There were people that—I’ve been approached on that several times. I looked into it and I saw what had been done and what the highest thermo power materials were and how they got them, and I didn’t—I saw one possible way that we might increase the efficiency of them, but I didn’t think it would make a big difference.
The main thing we found in there is something that—as with almost everything I’ve done—the final result is actually trivial. The final result for thermo power was we said essentially, “OK, if you’ve got something where the electrons repel one another so they stay away, so they keep their spin, then there’s gonna be a contribution to the spin entropy and that should show up in the thermo powers ‘cause thermo power is entropy for carrier. So you should get KB over e, log2 term.” Boom. And in fact, we found it and other people found it and it tells when electrons are strongly correlated, that’s a paper which is cited ‘til today. It came out in like the mid-70s, right?
So essentially, all I’ve done, usually, is find some effect and reduce it to something that’s trivial. That’s trivial, but on the other hand, it’s interesting. And so I was thinking, “OK, we might be able to improve the efficiency by making the spins not interact, by making a material that has a large Coulomb repulsion but it turns out that would give you 60 microvolts per Kelvin and—give you an additional 60 microvolts per Kelvin—and people were already up to millivolts per Kelvin, so it doesn’t make much difference.” But I looked into it. And I looked into other ways that you might do things, and I couldn’t find one.
So you would have pursued it if it was viable, you’re saying. That you weren’t against the idea of thinking about commercial applications, it was just not viable.
No, as a matter of fact, I even have some amazing patents. Amazing in that they sound ridiculous, but in any event [laughs].
So again, I keep coming back to your entrée into soft matter, in what ways did advances in microscopes help you with this transition?
No, what got me was—you can ask me about why the field took off or you can ask why I got into it, and they’re two completely different reasons.
All right, so let’s set the stage then. Let’s talk about the field and then talk specifically about you.
OK. So in my viewpoint, recollection, there are many—there had been lots of activity, but not by the condensed matter community in soft matter for a long time, in particular on polymers and liquid crystals. Somewhat on colloids, but not as much, but typically not by physicists. Typically by other departments. I mean I think that I would say Sam Edwards and Pierre de Gennes and Masao Doi were people who were physicists, real, you know, card carrying physicists. They had real insights into other problems and stuff like that who realized that you could treat phenomena in soft matter with the same rigor, with the same interest, with the same prospect of discovering new phenomena and new physics. In other words, making contributions not just to understanding materials, but to understanding physics by going into those fields.
And so they showed that - of course, before that, what predates them of course was Landau. Landau theory applies to lots of things and it’s of course useful in soft matter, but in particular, —one of the things that de Gennes showed— that the paths that electrons take and the quantum mechanical paths that electrons take in one dimension essentially maps onto a polymer problem is an interesting insight from de Gennes, right?
And that showed lots of things and the fact that you could do Landau theory on the phase transitions that you see in crystals. Those are real insights. And the fact that when you—and of course it’s realized now, you know, some of this stuff which Landau did became the Phi four field theory, which influenced all of physics. The insights that you have from constructing order parameters for soft matter systems or phase separation, which is essentially now part of soft matter as well.
So I think the field really—and I may be—that’s my impression, that it came from a couple of people who were card-carrying physicists realizing that here is a field where you could really not only have impact on the field, but have the field have impact on Physics. So deGennes sort of did superconductivity for a dozen years or so, did magnetism, did superconductivity, did fabulous work in it. And then he said, “I’m not doing this anymore, I’m doing liquid crystals. I’m doing polymers.” And started making contributions in those areas and a lot of it was mapped directly onto what he was doing in superconductors and phase transitions and magnetism and things like that. So that’s my viewpoint of where the field came from and—
And then specifically on you, where did you get involved?
So what happened is—like I say, Pincus had been trying to get me for years to do it and I never really thought it was that great or that interesting, stuff like that. So I took a sabbatical at Orsay in Paris and I was supposed to work with two groups, one on superconductivity on proximity effects of superconductors, and the other on organic conductors. But my host was Etienne Guyon who was one of deGennes buddies and disciples in Paris dating back to when they were at Ecole Normale, and he put me in a room with his graduate student Pavel Pieranski.
And Pavel is an amazing guy. Everything he does is not only beautiful physics, but beautiful, and he walks in one day to the office, and he says, “Regardez ca”, “Take a look at this.” And I look at this little vial and it looks like liquid opal. And I said, “What’s that?” He said, “Oh these are colloids, these were charged colloids, they are charged and they form a Wigner crystal, they’re crystalline, and what you’re seeing is the Bragg scattering.” I said, “What?” And I took it and I just played with it and it was just amazing because what you could do is discover things by shaking and watching the colors disappear and oscillating to see the resonant modes in the cylindrical bottle—and this was just within 10 minutes—he was showing me stuff and I discovered stuff on my own—you could essentially twist it around your hand and you could melt it. So you could see the difference between the liquid and the crystalline phase immediately.
And then you’d say, “Well, OK, if it’s crystal, then it’s got rigidity.” So there’s a shear modulus, so there must be sound waves in it. Maybe we can look to the sound waves. So what you do is you just take this thing and you wiggle it back and forth and you see the colors flicker back and forth and that’s a standing mode of the sound inside, of the transverse sound. And so you could do all this stuff with that sample and I just said, “I’ve gotta work on that stuff. Let me work on that stuff.”
And so part of the time I spent there, I spent with Pavel, under his microscopes, looking at these guys, doing theory and how you could play with them, how you could do different things with them. I knew from electronics the idea of a Wigner crystal and nobody had ever seen Wigner crystal in those days. And here, in front of my eyes, was a Wigner crystal, right? It was classical and I could see the individual particles. And the interactions weren’t quite as long range, but nonetheless, they were long range enough. The crystal structure was what you’ve expect from a Wigner crystal. I was just blown away. And from then on, I was really sold on doing soft matter.
Before, it was like, as I said, what turned me on to experimental condensed matter was the fact that you could come up with an idea, you could make an apparatus, you can go in the lab and you can get the result. Here, you could do it in a half hour, right? Here you could say, “Oh, that looks interesting. Let’s see.” And there it is.
And, of course, they’re amazing phenomenon. Everything that you could look for or not, everything except quantum effects, that you could look for in hard condensed matter you got in soft condensed matter. My impression has always been the following; the reason that there is condensed matter physics is the many-body problem, right?
Paul, can you explain that a little? Why is the many-body problem so fundamental to soft matter physics?
All of it’s fundamental physics including the condensed matter. So when you’re learning physics for the first time, you get Newton’s laws and you get the other laws and you find that there are basic forces and stuff like that and you think the problem is solved, but as everybody knows now, that means you’ve solved the problem for one particle, and in fact also you’ve solved it for two particles because you can go to the center of mass. But even for three particles, you don’t have a general solution to the problem because you can have chaotic orbits, you can have periodic orbits, quasi-periodic orbits things like that.
And so even for three particles, you can solve the problem in that you can write down the equations, but you don’t know how to solve them. And in general, you don’t have solutions that even necessarily converge. So the one of the real—if you wanna say that physics is understanding nature, it’s understanding how particles interact and the phenomena you can get from it—you’ve only solved a very small part of physics when you know the fundamental forces.
And the business of condensed matter physics, which is the physics of many particles, is to figure out this many-body problem. And it’s not a trivial problem. There’s a start to the problem, is just the statistical mechanics, which is that you can, on average, find out how things interact and stuff like that, but the statistical mechanics alone is not gonna give you a superconductivity. It’s not gonna give you cooperative effects and things like that. It may give you crystals, which are one interesting many-body phase, but crystals don’t have all the degrees of freedom that you usually have, and in many phenomena, in almost all phenomena that are interesting, the system is dynamic, it’s out of equilibrium, and the many-body problem is absolutely essential to understanding that problem.
You can treat things on the average, you can use mean field, you can do things like that, but at the heart of it—and you can make progress that way—but at the heart of it, the individual particle interactions will play some role. And a role that’s a different for different systems. You only really touch a tiny bit of physics if you’re just doing single particles, and the rest, you know, physics, chemistry, biology, are essentially all the many-body problem. Cosmology, astrophysics, everything like that are essentially the many-body problem. And that’s why I think it’s fundamental and it’s really exciting.
And there are quantum many-body problems—you can learn from those—and there are classical many-body problems, you learn from those, and they inter-feed one another. And that’s why I think it’s fundamental. And that’s why I think the real business—the contribution to physics of condensed matter—is the many-body problem. And it’s nowhere near solved.
There’s some examples, superconductivity is one where you couldn’t get to that solution, even perturbatively things like that. That was a real inspiration getting away from electron physics and there are things like that to be done in both hard and soft matter physics. So I think that that’s really it. I think nonequilibrium phenomena, the many-body problem—which is a big part of that—is what’s exciting.
Paul, when you say, “It’s not solved yet,” what would solved look like? How would you know that it’s been solved?
Let’s see. Well, if I could tell you whether guppies can mate in turbulent pipe flow, I’d say that I made a—I really understand a many-body problem. Or I could say, “OK, I have two things and I want them to come together by their swimming motion in turbulent pipe flow.” Then I’ve understood something about what turbulence is. I’ve understood the interaction of living things with the forces that they generate and stuff like that with the turbulent flow and I made a prediction that is useful, at least to the tadpoles.
That’s an example. That wouldn’t be a complete solution. If I could say something more than what I said already about turbulence, that would be fantastic. People are making headway on that problem, on how you essentially go from putting in energy at a large scale to devolving it down to small scales. Or if I can take a system of active particles and get the opposite energy flow where I’m putting in energy and maybe correlations at a small scale and getting macroscopic correlations in energy collected, right? Things like that. If I could do that, I’d say part of the problem’s solved, but those are hard problems.
Paul, talking about guppies, it sounds like you need more biologists in soft condensed matter physics.
I mean, I think that we need collaboration with biologists. I think it’s probably good for them. If you mean having biologists come in and take over the field, I think no. And I think that soft matter is different from biophysics. I think biophysics, at least the way it is now—and this probably will get me in lots of heat—is mostly using physics to explain biological phenomena.
You mean as opposed to the other way around?
Well, as opposed to—I think what soft matter does is it essentially asks, “Well, OK. What phenomena do we have with complex systems? How can we understand a phenomena or the physics of what goes on rather than to explain why this works, rather than figure out why, let’s say, microtubules mechanically grow and then explode, can disassemble themselves immediately.” So figuring out what the physics is in the molecular assembly, specific molecular assembly for those molecules coming together.
If you find in general what causes structures to be stable or unstable, you’re doing soft matter. If what you do instead is you say, “OK, I’m gonna figure out how a motor protein works by essentially looking at how the ATP is used in a molecular motor to get the action,” the mechanics of that, the chemistry of that, I would say—this is really gonna get me in trouble—I say that’s biophysics.
If I want to know something like how, in general, do I imagine molecular machines working? What’s the overall operating principle by which they work, I’d say that’s soft matter. So between those, for example, is one of the examples I always like is the ratchet model, the athermal ratchet model that Jacques Prost made many years ago for trying to describe the commonality of molecular motors. That, I’d say, is soft matter, whereas if you’re looking particularly at how myosin works, it’s not. Seeing what happens in general when molecular motors assemble or what they can do or how they assemble in general, I’d say that’s soft matter.
An individual one is biophysics. Wow, this is really gonna get me in trouble. OK. [Laughs]
Please.
You want me to get myself more in trouble?
Absolutely.
Well, I don’t know what else I can tell you. I mean, biologists— to me—have two tasks. One is to actually figure out how biology works. They don’t have to make a contribution to physics, right? They have to figure out how biology works and in some sense, their responsibility is how it works microscopically and on another scale, you know, how it works globally.
How it works as what are principles of organization of biological systems, which are specific to biology. And I think—I hope—that the condensed matter of physics can also do some of that, but in general, condensed matter physics is trying to figure out how things organize on a macroscopic scale from the microscopic scale. They’re trying to figure out condensed matter, soft matter is about how that operates not only for biology, but in general. Biology is part of it. I don’t like the idea of mimicking life so much because—although everybody tells you that biology is optimal—what I’ve never seen in biology—what I always say is I’ve never seen anything in biology make an I-beam. And an I-beam is structurally much better than anything else for what it does. So it’s not that optimal is the answer. Actually my opinion has always been all biology has to do is work. It really is not required to do anything optimally or otherwise. If it does, that’s fine, but I don’t know whether that’s a general principle.
But I’ll leave that to the biologists. I think that we can really impact on that, but the role that soft matter has or condensed matter has in general is figuring out cooperativity and maybe using it as well ‘cause a lot of things that we do are useful. A lot of the things that we discover about cooperative things, how you build things, how you assemble things, what you can make and what you can’t make and how to activate them, that’s all in the purview of condensed matter and especially soft condensed matter.
Paul, let’s talk a little bit about your move back to the east coast. What was the circumstances? Did you accept a joint position at Penn and Exxon or did they happen separately? How did that work out?
It worked out in a funny way. What I thought—I was at UCLA. I was fine, I liked it there and everything, and so I looked at Penn and I looked at Exxon and I said, you know, they’re both really good. They’re both essentially each individually not necessarily better than what I have at UCLA, but if they were combined, it would be definitely better. [Of course the number of people, the contacts and stuff I would have with really good people that I really respected and stuff like that would be higher.
And so I told people at Penn and Exxon this, and they were going to try and give me a joint appointment, figure out how to do a joint appointment. And I was guaranteed that they could do it, well of course they couldn’t do it. It turns out when they finally got together to try and do it, they said, “There’s no way that the intellectual property da, da, da, da, can possibly by handled. There’s no way we can do this.” So what they arranged is that I would be a non-regular employee at Exxon at no more than 49% and I would have a regular position at Penn. And so that’s what it turned out to be. That was the arrangement.
And were you in touch with people like Harry Deckman before or you only got to meet him when you got to Exxon?
I spent a couple of summers before I moved there. I spent a couple summers at Exxon, so I met Harry there. I met Harry—I forget how I met Harry. I think I met Harry just before or after I met Dave Weitz. What I remember is I think I must have met him after. Well, I could have met them at the same time. I remember going and giving a talk at Exxon when it was in Linden, I think, New Jersey. Anyway, and they used to be there before they moved out to Annandale.
So I remember giving a talk there and I remember going out for dinner with Dave Weitz who didn’t like it at all. And I remember going out to dinner and we got talking about something and we decided to go back to the lab and do an experiment after dinner. So we went back to the lab and we did an experiment and I think Harry was there then. I think Harry was in the lab then. He may have gone to dinner with us as well, but I think that was it.
But I really liked this because it was just—I’d been visiting lots of places, in no place that I’d gone to they would come up with an experiment and say, “Yeah, let’s just go to the lab and do it.” At night. I said, “OK, this looks like my kinda place.”
Were you following—did you recognize that Exxon was becoming quickly a center of excellence in soft matter research?
OK. So it wasn’t the first couple times I visited during the summer, but then they sort of decided that they were gonna become the Bell Labs of soft matter. It wasn’t called soft matter then, so complex fluids maybe or in any event, whatever the field was going to be named, but then it was clear that by the time I went there, it was clear that they wanted to be the world leader in that research. And indeed, they became the world leader in that research, without question I would say.
What kind of expectations did Penn have? Did they just want you to do the best possible research or were they looking for you to teach classes and take on graduate students in a more traditional, full-time role?
I was absolutely in a full-time role from day one at Penn. I taught, I had graduate students, I had undergrads. What I had when I was at Penn and Exxon, I spent between four and six days a week at Penn. I think I would go up on Fridays. I’m known at Exxon as showing up—if it’s Friday, I’m there. Or they could tell it was Friday because I was there. And I think that’s still true at Exxon.
So I had one day a week—which you’re always allowed to do consulting at a university—and I would spend that at Exxon, and if there were exciting things going on that I was involved in, I would go up on weekends. And otherwise, on the weekends, I’d be at Penn. So I had a regular position at Penn. They expected me to be a regular professor there.
So essentially as far as they were concerned, anything you were doing at Exxon was extracurricular?
Yeah, absolutely.
This must have been an insanely busy time in your career.
It wasn’t that different. [Laughs] It doesn’t matter if I’m working at two places or one place, I still tend to do the same amount of stuff. No, that’s not right.
Did you live in central Jersey? Were you looking for a place to live where you could equally commute both ways?
Well, that’s one of the reasons I went to Princeton.
Right, that makes sense.
Yeah, the commute—it was an hour and 10, an hour and 20 minutes each way to get where I lived to Exxon from Penn. And for about a year, it was a little bit longer because I would stop by and pick up Dave Pine and he and I would drive together to Exxon.
So five years was enough for you living closer to Penn while you were trying to get up to Exxon.
Yeah. I had no problems whatsoever with Penn. Penn was a terrific place. And my interaction with Tom Lubensky was amazing. He’s an incredible scientist. They’ve built up a lot in soft matter since I was there. So they’re much stronger in soft matter than when I was there.
And so in 1988 you reached out to Princeton or how did that come about?
No, I didn’t reach out to Princeton. It wasn’t on my mind. They contacted me and they said that they were interested in materials and they knew that I was interested in materials as a physicist, an interest in materials, and they wanted to start a multi-disciplinary program in materials. As well as—no, but in fact, I should say one of the people that really recruited me for Princeton was Phuan Ong. Do you know him?
I don’t.
Phuan is really an excellent hard condensed matter guy. I think he just retired from Princeton. But I had known Phuan had worked on low dimensional conductors we were still working on the organic conductors, and charge density waves. I knew from—he was at USC at the same time I was at UCLA. So he was the cross town rival, but we both did a lot of work—he was the guy that discovered charge density waves, sliding charge density waves, which is a field in its own.
So he is the father of that field. And he was at USC for a long time and when we were together, we collaborated on several papers and we fought on others. We were at odds on other papers, but when he went to Princeton, and some years after he went there, he contacted me and said, “OK, what about coming to Princeton?” And so I went up and saw what they were doing and what they were gonna do and stuff like that and it was closer to Exxon. There were some really good scientists there, of which you certainly know some.
And how strong was Princeton at that point in soft matter?
In physics, it was zero. Even more zero than it is now. Well, I could also take it back. There were already two biophysicists there in the physics department. Bob Austin and Saul Gruner were there as biophysicists. They didn’t have any soft matter physics, but they have a good chemical engineering department and I had known or I knew of—from their research—several people in chemical engineering at Princeton.
And when I went to meet the president, who was trying to recruit me, I remember I was very impressed. It was Bowen, and Bowen was really impressive. I mean, he obviously did his homework and he’d do things like, he’d say, “OK, so we’re building up material science, we want you for that.” He said, “My top priority for the next month is getting you here.”
That must have been nice to hear.
Yeah, that’s always nice to hear, right? But he obviously knew his job very well.
And did he also appreciate, Paul, that Princeton really lacked any expertise in this area?
Well, but like I said, they had chemical engineering. People there were really good—
But I’m saying for physics.
Did I dislike the fact that they—OK. So I’m fine if they weren’t in soft matter. I was also doing experiments—so half of the things I was doing at that time were hard condensed matter. So I knew I would be interacting with Ong, with Phuan, with Fyl Anderson, with the other people I forget because their—Princeton is now really, really strong in hard condensed matter and they have a little bit of soft matter in physics, but not much. Actually only one person who’s sort of soft.
So no, I wasn’t particularly worried about it because throughout my career, I have always collaborated with people in other departments. So it’s always been, you know, with chemists, with chemical engineers, pretty much with chemists and chemical engineers. To me, what’s amazing is chemical engineers really, really know a lot of physics. Really know a lot of physics. Absolutely impressive. And that was certainly true of the people at chemical engineering at Princeton.
So I knew Bill Russell and Bill Russell’s really excellent. And he and I started a group together at almost maybe the first year I was there—maybe before that maybe or after, but anyway, he and I had a group together essentially as long as I was at Princeton.
I’m trying to think who kept telling me so many wonderful things about Bill Russell. I want to say it was Norman Wagner if that would make sense.
It might well be. I mean, Bill Russell’s a fantastic person and a fantastic scientist.
Yeah. What kind of work did you do with him?
We set up a group on colloids and we did lots of stuff on colloids. Some of it—somehow, early on, we got into doing stuff in microgravity, so we did several experiments that were flown on the space shuttle. We actually had some nice results from them. We did a bunch of things on—it was all work on colloids. So work on colloids. Colloids and interfaces, microgravity, colloidal crystals.
We developed a couple of techniques that were kind of neat, like using—when you take colloids that are hard spheres, the only control parameter is the density, so we came up with a way of controlling the density just with an electric field so you can take one sample and do an entire range of densities, you know, span the entire phase space for the system. Hard sphere colloids which were just coming in at that time. I don’t know, lots of things. Colloids.
And did you set up another lab at Princeton or between Exxon and what Bill had already put together, that was enough for you?
No, at Princeton alone I had three labs most of the time. So I had my low temperature lab in physics in the basement of Jadwin Hall and then I had a lab in what became the Bowen Building, which is the Materials Research Center. I had a lab on one floor with Bill that did colloids and I had a lab up on the floor above, with Rick Register which was on diblock co-polymers. So I had three different labs there with different students.
Paul, did you take students with you from Penn to Princeton?
I think, yes. I took at least two. I think. Yeah.
And then as soon as you got to Princeton, I assume you had new students right away?
Yeah, by right away you mean within the first year?
Yeah.
Yeah - some of whom then accompanied me to NYU. That was much later then, though. But also, what we did is we—I had students from both Penn and Princeton at Exxon. Actually, at Exxon, I started off with students—so I started soft matter when I was at UCLA. I started doing research in soft matter. I continued the stuff in hard matter, so when I moved to Penn and Exxon, I took my soft matter students to Exxon and my hard matter students to Penn.
And then I had soft matter—when I continued in soft matter, my soft matter at Penn and—that’s not true. At Penn, most of my soft matter was at Exxon. When I moved to Princeton, I had the two labs at Princeton and another lab at Exxon with soft matter. And the Princeton students—some of them came to Exxon and did their thesis there, right? So I had students—at different times—students from UCLA and Penn and Princeton in the Exxon lab.
Wow. That’s a lot to manage.
No, it was different times. So I usually would have one, two, at most three students at Exxon. And/or postdocs.
And just in terms of compartmentalizing all of this in your mind, did you keep these worlds separate in terms of the kinds of research projects you were doing or was there overlap?
I mean, they were pretty much separate, but of course there was overlap in the ideas and there was transfer of ideas one to the other. What we found in one, we used in the other. But the experiments were separate experiments.
So to the extent that you could separate them out for the purposes of explaining, what were some of the most important experiments you did during you time at Exxon and at Princeton?
Well, I know the only one that I’m ever gonna be really known for is the one that was done with undergraduates at Princeton. Well, it was started with undergraduates at Princeton, and then I had a graduate student, which is the packing of M&Ms. You may not know this. So the only thing I’m known for is we showed that M&Ms, that is to say ellipsoids, pack better than spheres, to a higher density than spheres do, both randomly and crystal.
You mean like an M&M, the candy, is an ellipsoid?
Yes. So what happened is, I mean, there’s this problem people have studied for literally thousands of years - the packing of spheres. It’s an ancient problem and it’s been solved by grocers because when you go to a fruit stand, you see that the oranges are stacked up in a lattice and it turns out it’s a Face Centered Cubic lattice and people have known that and stacked them that way forever, since there was fruit, and it’s been—this thing is also known as the Kepler conjecture because Kepler raised the problem of the densest packing of spheres, and again, Kepler conjectured that it was this FCC packing, cubic packing, but it’s only been proven mathematically that that’s the densest packing maybe 15 years ago. So mathematically, we now know that’s the densest packing.
But it turns out that if you just take spheres and you pour them in a container, they don’t make a crystal. They make something that’s disordered, which is like what happens if you go to your grocer and they don’t stack them up, then you just see the oranges all over the place. And it’s known and it’s been known for a very long time that that doesn’t pack as well and instead of going to 74%, which is crystal packing, it packs to something like 64%.
74% means that there is 26% empty space?
Yes. So for example, if you fill a box with marbles and you put them exactly in an FCC lattice and then you pour in water and you drain it out, the water that’ll come out is 26% of the volume of the box, just as you said. OK. So if you do it randomly, that is you just pour them in so they go everywhere and they don’t form this crystal, they fill space to 64%, which is less.
So it turns out that problem—everybody sort of knows that answer, but nobody has done anything in order to prove or even show mathematically or even define the problem mathematically. And the reason is because nobody knows really what random means. That’s really why it’s a real hard problem. People think it’s an undefined problem.
On the other hand, they know something about it. So one of the things—this is kinda cool so I’ll tell you, but I’ll make it short ‘cause I could go on forever on this. So Maxwell made the following observations; suppose I know nothing else about its system except that it’s stable. Except that nothing is moving in it, it’s sitting there and it’s stable. I can still say something about it. Suppose I have particles there and these particles can have any arrangement you want with them. But nothing is moving so everything’s stable. So he learned something just from the fact that it’s stable.
What I have to do to have it stable is that I take a particle and I know that it can move in three—let’s say it’s in three dimensions—it can move in three dimensions. That means I have to specify—for it to be stable—the three coordinates, X, Y, and Z for this particle, right? OK. So I have three numbers for that particle, three variables. Now if I have N of them, I have lots of them, N together, that means I have 3N variables. And for this to be stable, I have to specify them.
And we learned in public school that the number of equations has to equal the number of variables if you wanna solve it, right? So I need 3N equations. But if all I know is that these guys are particles and they could be touching or not, the only way I can get an equation is from a contact. So I know that there must be three N contacts. And since a contact has to have two neighbors, this is a neighbor of that and that’s a neighbor of this, so that means on the average, every particle in this random pile has to have six neighbors, which is really clever, right? And that’s from Maxwell.
And the idea was if you’ve got something that’s not, you could have more than that, more neighbors but you can’t have less than that. You can over-determine the problem, you can under-determine. So people have studied this problem and indeed, people have found that when you have marbles randomly thrown in a box, there’s six contacts, six neighbors, and the packing fraction is .64 and that’s all that anybody had done for thousands of years. So I said, “Well, what if you use a different shape?” Some reason I got in my head what if you use a different shape? What’s gonna happen? OK. And you know for instance if you choose cubes, you can completely fill space. Of course if you pour them in randomly, you don’t know what they’re gonna do. What you can do is you can just pack face-to-face and it’ll completely fill the space 100%.
So I said, “Well, what’s the simplest object other than the sphere?” Physicists always use spheres. It’s the simplest object, right? So what’s a small perturbation on a sphere? The answer, it’s an ellipsoid. OK. So why don’t we try packing ellipsoids? So now the question is where do you get lots and lots of ellipsoids all the same shape and all the same size? Answer; M&Ms. OK? So—
Now Paul, did you have an amazing, comic book moment where you were eating M&Ms and you looked at it and you said, “Ah-ha”?
No, it was actually was more like what I told you. It turns out though everybody thinks that that’s the case. Everybody thinks that I got it that way because for many years before this, my lunch was always coffee and M&Ms. Everybody at Exxon knows that, everywhere I go to lunch, the only thing I would have for lunch is one bag of M&Ms and coffee. So everybody thinks that that’s it, but in fact it wasn’t that. It wasn’t an “ah-ha” moment, it was, “Well, what can I use?” And I was thinking maybe I can flatten some spheres and I can make some, da, da, da, and I was looking around and it’s like, “Oh yeah. I can use M&Ms.”
And then I went and I had somebody measure and it turns out they’re all the same size and they’re all the same shape to better than half a percent and we did all sorts of studies on it. And then I got somebody to do the experiment. I got an undergraduate to do the experiment. And then—so I got one guy to do the experiment, he did it, and he found that they pack better. It packs more densely. It packs essentially to 70% instead of 64%.
So then I had what I considered this lightning moment, which was OK, the interesting thing is why does it pack better? I mean, first I didn’t believe it. I told—I had a student do it and then I had another student check it and then I did it myself because I said, “What the hell?” I know that I can get from a sphere to an ellipsoid just by squashing.
So if I take a box full of spheres and I squash it and nothing changes, the packing’s gonna be the same because the volume—well, anyway, the volume of an ellipsoid is 4/3 pi abc, instead of 4/3 pi r3. In any event, I get the same volume fraction – filling factor. So I didn’t believe it and then I had a couple more students to get them to do it and then? Why better? And that’s the time I had the Eureka moment, if you like, which is I said, “Oh, it’s because when I have an ellipse - if I have a sphere, all I have to tell you is X, Y, and Z. But if I have an ellipse, especially and ellipsoid of revolution, there’s a direction along the axis of rotation and that has two angles associated with it and that means that the number of degrees of freedom I have to specify is XYZ and two angles.
So instead of having three equations per particle, I need five equations per particle and therefore instead of having six neighbors, I have to have ten neighbors. And if I need ten neighbors, I have to pack it denser than if I only have six. Does that make sense?
Yes.
So then I got a summer student who was from another university to sit there, pour paint in, drain the paint out, dry it, take the things apart, and see where—on the M&Ms—how many patches were there with holes in the middle. So this guy spent the summer doing that. And there should be—instead of for the spheres he counted—and as you expect, there were six—that’s also what other people had counting and with the M&Ms, it was 9.86. So essentially ten. So that was really, really, really cool.
And so there’s a picture of me—I think there is with Sal Torquato, somewhere standing with Sal who’s someone who used to be a chemical engineer—he’s also a chemist and a physicist now—in front of a big barrel of M&Ms. And so that’s what—probably actually if you just Google my name, that’s one thing that’ll come up. So that’s essentially what I’m known for.
Well—
And it’s actually kinda cool. And the thing to me that was great, I mean I really—when I figured this out—actually when I figured it out, I knew that that was the answer and the experiment was just the thing at the end, but the thing that got me is I just went around for weeks after that thinking, “Jeez, I just have the best job in the world. This is just the best job in the world. I get paid for figuring out that the reason that ellipsoids pack better than spheres is that they’ve got 2 more degrees of freedom. It’s just amazing that people pay me for this.” So I was just ecstatic with this.
Were there commercial applications for this? I mean, are people in the packing and shipping industry, are they interested in the significance of this finding?
So far, I think the answer is no. And the people that actually should be interested in it are people that make ceramics. The reason is with a ceramic, you know, are particles in a solution and they mix them around and they put them in a container and they mold them and the strength of the ceramic—comes from how many contacts you have and the density.
And it turns out if you use spherical particles or something like spherical particles, the best you can do is 74% and the number of contact—actually the best you can do is 64% if you just pour, and the number of contacts you get is 6. Whereas if you take ellipsoids, not like M&Ms but with three different axis, the best you can do is like 77% and the number of contacts you get is twice as many, it’s 12. So you should be able to make much better ceramics that way.
Now, the interesting thing though is that the candy company which makes M&Ms, they didn’t know that M&Ms pack better, and of course that’s exactly what they don’t want, right? Because if they pack better, that means if you sell a bag full of M&Ms, you have to put more of ‘em in to fill it than if they were spheres. So in fact, what would be the best for [the] Mars [Company] is having spheres in there or something even worse that would pack less densely. Then you can sell for the same price fewer M&Ms.
Anyway, that’s sort of I think the only thing I’m gonna be remembered for. I don’t mind because it was a cool experiment and a cool idea. And it was something where it was really—when I got it, it was like, “Wow, that is cool. I figured this out and now I really understand.”
That must have been so satisfying to just have like a complete understanding of something.
Yeah. It was like—I’ve only had that a couple times in my life.
Some scientists never experience something like that.
Maybe. Hopefully they do.
Paul, can you talk a little bit about your decision to join NYU? How did that come together?
Well, they called me up for a recommendation for somebody, which I gave them, and I said, “But you’re in New York, you should take advantage of it.” New York is the best place in the world. Anyway, so I told them that and so I get a call back and they say, “Listen, why don’t you serve on an advisory committee?” They said, “We have lots of retirements coming up, we haven’t hired people for a long time. Why don’t you come and serve on our advisory committee for another year, couple of years, and tell us what you’re gonna do?” Or, “What we should do,” I should say. “Tell us what we should do.”
So I joined it and I came for a couple years and they had some plan where in fact they had—they were gonna have 12 positions open to fill in the next four years or something like that, and so I listened to their plans, I met with lots of people, and their plans was gonna be, “Well, we’re gonna replenish this group or we’re gonna replenish that group and we’re gonna replenish this group, maybe we’ll start something over here.”
And I just said, “No, this is just ridiculous.” I said, “If you wanna build something up, what you wanna do is you wanna build something up which uses your positions in a smart way and makes you instantly, or in a very short time, one of the top departments in the world, or the top department in the world in this area. And you don’t do it like that. The way you do it is you take your 12 positions and you pick one or two fields and you just put all your positions in those fields.”
And so for a year or two, this thing went around and they said, “Well, what areas are open?” And so I said, “Well, you should do—“ I forget what I said. I think something like, “You should decide. You should either do quantum computing or soft matter. Something like that. Those are fields that are opening up. There is nobody that has the dominant thing in that. There are groups everywhere, but you should take advantage of it and do one of those two fields. It’s up to you which you wanna do.”
So after a while, they said, “OK. Why don’t we give eight people in soft matter and you do it.” So, eight people is a lot of people. And I talked around with some people and I told them I could get other people to come and I talked with Dave Pine and I called him and I said, “So when are we gonna form a soft matter group at NYU?”
[Pause]
OK I’m back. Where were we?
Starting the group at NYU.
Yeah. So I also knew that Dave loved New York. And actually what I was hoping for, which unfortunately hasn’t come to pass, is when the university said they were gonna really make an effort at NYU in physics and the sciences and stuff like that because they weren’t particularly known for that.
I mean wasn’t physics really in a pretty deteriorated state at NYU?
They were except that they had been building up—they had a new chair who had actually spent some time in building up cosmology, astrophysics, particle physics. And so they had been building up for a while. But the president seemed to be going strong. It was Sexton, Sexton was president when you were here as well?
Yes, he was.
And Sexton is a real character, you know, right?
And I mean he was on a mission to just build NYU all over the place.
He was. He got in a lot of trouble for it, I mean, but I thought it was pretty good. He had a vision, which a lot of people don’t. And his vision sounded good for NYU. Anyway, what was in my mind especially was I thought, “Oh, the only place I know of that has gone from zero to first rate is Santa Barbara. In my lifetime, anyway.” So what I thought is, “Well, OK. So half the world wants to be on the beach, half the world wants to be in New York. I’ll take the people that wanna be in New York.” But the difference is that, Santa Barbara had the fiscal backing of the state which NYU didn’t. But that worked for Pine because I knew that Pine really liked New York.
And he came the same year as you, right?
We came together. Yeah. We both decided that we were coming and that’s what started it, essentially. Actually David Greer had come the year before. I think we were the ones to negotiate—Pine and I are the ones that negotiated the soft matter center. So the thing that I didn’t realize—I still think it’s true that half the world likes the beach and half wants to be in some place like New York—the problem is that it seems, not surprisingly when you think about it, that the University of California has a lot more resources than New York University has. And that’s why they could do it and at least we haven’t succeeded in doing it yet.
Yeah. Did you see this as an opportunity, since you were essentially building something from the ground up, to change your research focus in any way or did you essentially bring all of your Exxon projects and experiments with you?
I think that about the time I went—well, I actually change research directions a lot. I mean, actually I got—I always change my research directions. OK. So we were working on colloids and stuff so I was gonna stick on doing some stuff that I knew, but I really got convinced, if you like, or convinced myself, that the future was not in building stuff that’s static or that’s just studying the property of colloids. I thought it would be really cool to try some things that are challenging.
Actually, there was a party for a 60th birthday thing, it turns out, for Tom Lubensky and I, which was held at Exxon. And Weitz, Dave Weitz, organized this pretty much. He said the only thing that we had to do—well we did that—is to present a list of the ten most important problems in soft matter to do that. So I actually thought about it and discussed it with Tom and stuff like that, and essentially, I forget how we did it, but our list came up with something that was essentially creating something that can self-reproduce. That would be the top goal. But of course I was the only one that took this seriously. Nobody else gave it a second thought.
So sort of from that time on—that was just the year before or the year that I came to NYU—I sort of decided what’s really interesting is the dynamics. Some of the dynamics we solved. You’re not gonna learn anything new from thermodynamics. You may see a couple new phenomena and stuff like that, but most of the new stuff is nonequilibrium, had to do with well, how does nature do stuff? What’s the physics of life, if you like.
So we started working—from that point on, essentially since I came to NYU, most of the work has been on active matter of one form or another. I have a group with Ned Seeman here who’s in chemistry. The guys I work with are great. All the guys I’ve worked with through the years are infinitely better than me. All the guys I’ve worked with have really done something, like Bill Russell, for example, he’s really amazing and stuff like that.
Ned Seeman, who I work with now at NYU, is the father of the field of DNA nanotechnology. He is the first guy—I don’t know if you know him—but he is the first guy that ever came up with the idea that, “Oh, you know this DNA which recognizes stuff and everything like that, you can actually use that as a material and build things with it.” He founded this field “DNA Nanotechnology” with maybe 300 institutions in the world now that work on it and stuff. Anyway, so I collaborate with him. He’s fantastic.
And so first the idea was we wanna take particles that recognize one another and so this had been done before, but we were the first to do it with colloids and Pine was part of this and other people were part of this. So we just took colloids and put DNA on them. And now you can make it so that this colloid will only stick to that one or only stick to that one or stick only to these two or whatever. And now we’ve designed all sorts of things where you can put the colloids together.
And then the real thing I wanted to do because of this list that I made, that Tom and I made for Dave Weitz, is to make an artificial system which would exponentially grow, which would replicate, which could exponentially grow and could mutate, right? So I wanted something that could show that we knew enough about this to actually design a system that could do elementary life, essentially. If you try and think of what life is, nobody of course knows how to define what life is, but one of the components that usually people associate with it is self-replication, exponential growth, the ability to evolve.
So we did it. We have a system that does that. It’s made of tiles that are made from DNA, but they don’t use the DNA replication. They just use the DNA so that they can attach to a specific other tile and they do that. So most of the stuff we’ve done since I came here is active matter. We did all sorts of stuff with active matter, like matter that you shear. We made particles which are swimmers and the swimmers swim, which means motile, they have a velocity going forward. When you turn on light they swim and they stop swimming when you turn the light off, and they do all sorts of interesting—show all sorts of phenomena.
We have particles which we can spin and those guys are really interesting. They’re the first example ever of a system that self-assemble from hydrodynamic interactions alone. So the only interaction between these particles is through the fluid. Essentially you have a particle that’s rotating that creates a flow field that’s rotating and that advects other particles and it turns out they set up a flow field so that they clump together and roll as what we call a “critter”. And that’s the first time that people have ever assembled things just from their flow field.
So we’ve done all sorts of things that are actually really kind of neat which are, you know, in this newer field of sort of active dynamic soft matter. But that’s the way the whole field is going, I think. I mean it’s not just us. So it really did change. I mean, what’s interesting is—on the other hand, I have—but it’s really I had—, I had a low temperature lab at NYU up to maybe four years ago. Probably five years ago. But probably my last hard condensed matter Phys Rev Letter was 2016. So I carried on the hard matter stuff through all of this stuff as well.
But in recent years it’s been almost exclusively soft matter.
Yeah, in recent years it’s been almost exclusively soft matter. I’ve only had one, maybe two—two graduate students at NYU that did hard matter.
So Paul, I think that’s a great segue to my last question since we’ve worked up to the present in terms of your research and that is it’s easy in particle physics or in cosmology to talk about the big unknowns, you know, dark energy, dark matter, if we ever build a new accelerator, what particles are we gonna find in there? What’s the analog in your view in condensed matter? What are the ongoing great unknowns as you look forward and what would it take, both in theory and experimentation, to push the field forward towards those discoveries?
OK. The big problem is the many-body problem, but—
It’s where we started at the beginning.
You’re not gonna solve that, right? You’re not gonna solve that and close the book on it ever, I don’t think, right? But I got pushed by NASA to hold—so I ran a workshop 17 years ago on what was hot in soft matter, essentially, right? So we ran a workshop in March on “grand challenges in soft matter”. And I just wrote sort of the first little synopsis of it. So I can tell you what I decided from that, which is taking together the ideas from everybody in this workshop. There were maybe 130 people that were in this workshop, which was all Zoomed. We scheduled this for the March meeting but the March meeting was canceled, so we had it on Zoom.
But essentially what comes from it, which isn’t just my ideas, but sort of by reading of community is indeed nonequilibrium phenomena are what we should be working on in the 21st century. We want to know the organizational principles for things that are living or life-like and how to make them. So we want to know how things that are life-like work. What do I mean by that? So one of the big things is suppose I have—what we’re used to in soft condensed matter is how do you self-assemble things? You take things and you have forces acting between them and you can self-assemble them and you can make them in different shapes and stuff like that. OK. But that’s old. Don’t tell everybody that that’s old because there are a lot of people that are still working on that.
But suppose instead of doing that, you make a little machine. So instead of using a colloid, you take a colloid, which also swims, which also recognizes another colloid because they fit together or because they do this, that, where you can turn the swimming on and off, where you can make this essentially a little machine. So you take a little machine and now, instead of using your individual little colloids to make something, to assemble something, you assemble it out of machines. So you have a machine that’s made of machines.
Well, as you’ll immediately see, everything that’s alive is a machine that’s made out of machines. So this is sort of asking the general question, how do you do that? How do you take something which has active particles in it, particles that can sense, that can change, that can move, have them interact, and figure out what are the problems associating with that system? And it’s a nonlinear system. You have feedback in that system. It’s a complex system. And it’s a fundamental question in physics, how do you take a dynamical system and will it always go chaotic? Is there any way of controlling it? How will it organize? You know, these are basic fundamental problems.
And so I think that for example, nonequilibrium phenomena, dynamic self-assembly, something which has an amazing amount of complexity in it, how does that work? How can you do that? How can you figure something like that? How do you figure out its stability? And it’s sort of a fundamental question on life, on biology, but it’s also very physicsy, a question on the essence of biology, right? How do you get something like that?
Now you can also ask, on a different scale—but it also sounds very related—how can you make an ecosystem? In other words, how can you build something where although they’re still not connected together, you know, like a machine, like building something like that, but you have lots of pieces that are, themselves, acting cooperatively and now you have something else is that—how can you get those to be sort of an ecological system so that they feed back off one another? It’s another level of complexity - stuff like that and it’s a problem that’s a general problem in society as well, or for instance for anything you’re doing or for any way that societies or biological systems or active systems in general are gonna interact.
So that’s one of our grand challenges. Grand challenge? Let me see for a second. I forgot what I put down for this. I mean one of the grand challenges, although it’s not written as such, is artificial life, but that’s something else. So the third one that we came up with or that I tried to distill from this workshop is active meta-materials. That is to say you can tell all of this has to do nonlinearity, activity, and being out of equilibrium, but meta materials, you know, that have been made so far are static. They are inanimate and things like that.
And so the idea here is you can make, for instance, fluids where the fluid element—the particle the fluid is made of—has its own energy source and its own activity and hence, it’s not something where you act on it and you dissipate your energy in it, but it has its own behavior and it can give energy out to a system rather than simply absorb it. And making more and more sophisticated machines like that.
So essentially, all of these things tie together into something which essentially means that the problems we have involve many-body problems, but now not just with static forces, which you’d have if you were just doing the four elementary forces, but now you’re looking at things which themselves are active and can do things and can recognize and can compute and can learn and everything like that. That’s where I think the field is.
And it sounds like you and your colleagues are going to be busy for some time to come.
You bet. And if we do this thing right, then what we’re gonna do is we’re going to make some soft matter constructs, which will figure these things out for us, right? [Laughs]
This is like artificial intelligence meets condensed matter.
You got it.
Wow.
I mean the idea was essentially, when I wrote it here, was OK, so we’ve got things that self-assemble. We already have made things, individual things, that self-assemble, that move, that change shape, that sense light and temperature and all the things like that. Why not also make them think?
Are you waiting for computers to get more powerful to make these things a reality or that’s already there?
It’s time to start now.
Yeah.
One of the things that we wrote here is that, you know, it looks like—in this field—especially since becoming more complex and dynamic and stuff like that—things like machine learning and stuff are gonna be—are starting to play a role and they’ll play a bigger role. And it’s also probably that—I mean, so people are starting not only to use machine learning essentially to analyze data and stuff like that, they’re using it to do experiments.
So you do your machine learning with the experiment as well, right? And that’s a new way of doing experiments. So you let the program determine which way the experiment goes and have it converge or diverge or whatever. And of course in doing this, when you do the experiment, you do that—you are also learning something about how machine learning works and feeding back on it. But indeed, we expect that there’s gonna be stuff feeding back.
There are people—I’m not one of them yet—who are doing things like—you can make robots. They’re not quite the size that we wanna do, which is micron scale, tens of micron scale, they’re on millimeter scale, little robots that signal to one another and you can just put them down and turn them on and actually have them move to their power source and have them do things. Like for instance, since they’re powered by light, they wanna go toward light.
And so you can have a simple program and a chip in them that wants to find light. But now you have many of them that are fighting for the same light. And now you can also put in them that—and this is not science fiction, this is of course real experiments that people are doing right now—you can put little LEDs on them so they signal to one another. And now you see what they do cooperatively, and typically a system with even those few features is complex enough that you have no way of predicting what they’re gonna do. You essentially have to do the experiment to find out or put the program in with machine-learning and have it feedback and see what it does. But as soon as you get something that’s even that simple, yet complex, the problems are more than we know how to handle. So there is a lot to be done. There’s a lot to be figured out. Just knowing the forces are not enough.
On that note Paul, there’s a lot to stay tuned for. I would like to thank you for spending this time with me, it’s been quite a ride listening to all of your stories and experiences throughout the decades so I really appreciate it and this is gonna be quite a special edition to our collection. So thank you very much.
OK.