Steven Chu - Conversations with History

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments
Captions
(electronic music) - Welcome to a Conversation with History. I'm Harry Kreisler of the Institute of International Studies. Our guest today is Steven Chu, who is a Nobel Laureate in physics and Geballe Professor of Physics at Stanford. He is the 2004 Hitchcock Lecturer on the Berkeley Campus. Steven, welcome to Conversations. - Thank you. - Where were you born and raised? - I was born in St. Louis, not raised there. I only spent a few years there. My parents then moved to Long Island. So I was raised in a suburb of New York City, Garden City. - And looking back, how do you think your parents shaped your thinking about the world and about science? - Oh, that's a tough call. (Harry laughs) Because most of how all parents shape people is very under-the-table. You're not really aware of it. I was aware of many things, though. My parents came from China. They came here as graduate students to go to MIT, and, like many Chinese in higher education, there was sort of a reverence for education and that they communicated very clearly to me and my two other brothers, that one of the highest things you could aspire to is to be a scholar, and to be a scholar just for the scholarship's sake, not so much as a stepping stone into some other job, and that was communicated very well. And while we were young, they would always say read. They didn't care that much what we read as long as we read, and there again was something that I felt. - And you didn't necessarily take to school like a fish to water. I read a short biography that is at the Nobel website. I mean, you got interested and were a good student when you got interested, I guess is the way to summarize it. - Well, you have normalize this to what was happening. I had an older brother who was an excellent student. He was two years older than me. We were in a school, Garden City, and it was a very good public school, excellent public school, and he went through this public school setting the highest cumulative average in the record of the school. - I see. - And so I follow along two years later, and the teachers would say, oh, you're Gilbert's brother. We expect you to do just as well. (Harry laughs) And so that was hard to really live up to. So while he was setting records, I was kinda coming from behind. I was an A minus student. By my family's standards, this was appalling. (laughs) He was very good. He was very structured and he would study the things he was supposed to study, and he was fundamentally a very good student, and I would get very interested in one thing and let something else lay, and something like that, and it wasn't really until I went to college where they didn't hear my older brother. (both laugh) - But in high school, you were turned on to science. How did that come about? - Well, there were two things. One is I had a fantastic physics teacher as a junior. - [Harry] In high school? - In high school. And then the same teacher as a senior in high school, and this is someone who is nationally recognized. He was winning prizes for being an outstanding science teacher. - [Harry] And his name was? - Thomas Miner. I had two excellent mathematics teachers, one in ninth grade in geometry and then a calculus teacher in the 12th grade, and there the mathematics was different than the other math. In the other type of mathematics, you learn how to do algebra and trigonometry and things like that. In those two subjects, it was mostly about logic and thinking and putting together logical arguments, and it was very different. And the physics and the logic slash mathematics courses I got very excited about. - And you also early on were something of a person who liked to build things and do experiments and litter your mother's living room with projects. Tell us a little about that, and then how that ultimately contributes to what you become. - Well, I don't know what it was, but since I was very young, I loved building things. I loved building things with my hands. I would be given a model set for Christmas, model airplanes or boats and things. I loved to put them together. I would ask my parents for things like erector sets. These are little pieces of metal and screws and unlike Lego blocks, you actually have to screw something together and it wasn't all pre-designed to make a boat or something like that. And I loved doing those things. In that respect, it was somewhat different than my two brothers. In many respects, my brothers and I are very similar, but in that respect, I seemed to love mechanical things in a way that was certainly nurtured by my parents in that they said, okay, he wants to do these things. We'll buy toys like that for him. But my other two brothers didn't seem to like that. Now it turns out that working with your hands and building things gives you some sort of spatial intuition and things like that that turned out being valuable once I became a scientist. I could see things in my head very clearly and could rotate them around and this idea of picturing things geometrically has always been a part of my thinking even though one doesn't think of that, at least the lay person doesn't think of that in terms of physicists. They think in terms of mathematical equations. And I only discovered later that most physicists do that. - Now then you went on to collge and you were freed of your brother (laughs) so to speak, as a model to emulate. Where did you do your undergraduate work? - I went to the University of Rochester. I applied to the Ivys. They didn't accept me. And University of Rochester was wonderful because it's an excellent school. It was then, and as I said, my brother was an unknown. I could be my own person. And I guess I started working very hard in a regular way still, but in a very directed way. But what's wonderful about college is beyond a few required courses, you can take what you are interested in. And then there was another thing that happened in college that I wasn't really thinking of, and that is, as I studied more mathematics, it actually affected my writing. My writing became more linear in its thinking and you could definitely see some logic in the writing that I didn't have when I was in high school, grade school. So my humanist courses, the professors there were being, hey, this is a coherent paragraph, and I wasn't really thinking about that, but it was almost as a magical transition from the mathematics I was studying. - Drawing on that right now, why is it that the public understanding of science doesn't proceed at a higher pace? Is it because not enough scientists are doing the writing? - That's a difficult question. As I give more public talks, as I get exposed to these issues more and more, I think that there are two things. One is unfortunately there is a public fear of science. They think, especially physics, they think, oh my gosh, physics is hard. It has math in it. It's gonna be very difficult for me to understand what's really going on. That, I think, is something that might've happened in grade school or high school. (both laugh) - [Harry] They didn't have the right teacher, right? - Yes, and the trouble with learning physical science and mathematics is once you slip behind a little bit, and you just didn't get this concept, or it's not quite firm in your mind about this mathematical thing that you have to know, well, then next week you're on to something else, but they really expect you to know something about last week. So that's one of the issues. The other issue is that the ideas are complex. But if you step back and if you spend some serious time thinking about it, the kernel of the ideas are quite often not complex. And so the essence of an idea, which we try to work with-- As a professor with my graduate students, we would read a paper, and look at the paper and say, well, what's the essence of the idea? What's something new? Forget about the equations. Forget about the complicated arguments and try to identify the kernel of the idea and I think that can be communicated, but it takes effort. - Now you did your graduate work here at Berkeley, so in a way, you're coming back to deliver the Hitchcock Lectures. What contribution did your Berkeley education make overall? I'm sure you could go through a long list, and who was your mentor here? - Well, my mentor was Eugene Commins. He was a wonderful professor. He has a history of having many graduate students that have gone on and done wonderful things, and he's revered as a classroom teacher as well. And also in the way he does things, the way he goes about life. So in every respect that I can think of, he was a mentor, not only in terms of the science, but how you handle yourself in terms of situations and how you handle yourself in the world. Now he had one remarkable quality that I wish I could copy and that is he made all of his students feel special, and they felt that they were special. They could do something and he really got all of us to live up to the highest we could do without saying you must do this or without making us feel pressured or guilty or something like that, and he would work side by side with us, often late into the night, as a colleague more than as a professor. So that was a remarkable experience, to grow up in that environment. The other thing I learned here is to try to think of things to do that would be important in science. There are many things you can study in science, but focus on some big questions. Try to identify the correct questions and there it was not only my advisor, but the Berkeley professors around here at that time, there were six or seven Nobel Laureates in the physics department active, and you can watch the way they approach problems. And this would come out not in a formal lecture or something, but in casual conversation, when they're maybe giving a colloquium, how they approached it, how they thought about it, and this also enters in a very subconscious way. That is probably why there are so many distinguished graduates of Berkeley as well. - Now help us understand what the prerequisites are for doing science well. If a student or students were to watch this, what should they know about this way of life and what they need to bring to it in training and so on? - Well, I think the first thing is they have to be interested in it. They have to be genuinely interested. They have to have some curiosity. Science is really about describing the way the universe works in one aspect or another, all branches of science. How a life form works, how this works, how that works. And so you're really trying to understand what's around you and you have to have a natural curiosity for that. In certain types of science, there might be some prerequisites. In physics, you should have some mathematical ability. Otherwise, I think it would be very hard. But beyond those prerequisites that a lot of people do have, you need to have first this curiosity, but a really driven curiosity. You want to know the answer. And with that curiosity comes with it a certain doggedness because there are gonna be setbacks. You're gonna discouraged. Things aren't gonna work. You're gonna have trouble understand things. Things are gonna be hard to understand, especially the first time. Science doesn't come naturally to people. I had the hardest time in my first few years as a freshman and sophomore understanding physics in a really deep sense, and also in high school, in the sense of, I could do well on the exams, but to really get it inside your stomach and to really say, okay, I have a real feel for it took awhile to develop that intuition. But there were other drivers for that. It seemed like a beautiful way of understanding the world. But I'll go back to the other thing, this doggedness, this real saying I'm not gonna quit. I really wanna find out. It enters in other walks of life. If you think of an athlete who wants to become a good athlete, well, there's gonna be a lot of training involved and sometimes you don't feel like getting up early in the morning or staying late in the afternoon and spending hours training. And so that turns out to be one of the most important things that separate in graduate school. At graduate school at Stanford, you have some of the best students in the world, and the ones who you can see are gonna go on and become world-class scientists, and the ones who are very smart and are gonna be good, but the thing that really differentiates is this passion to find out what the answer is and I'm not gonna quit. And so after those prerequisites, the thing that separates the people who are really gonna excel from people who are good and not is that, this internal drive. - In a joking aside yesterday in your lecture, you said that in science, once you announce something, that first everybody tells you you're wrong. Then they tell you it's trivial, and that you were not the first to discover it. Well, it really emphasized what you just said, that there has to be an inner drive, but also it suggests an element of courage, that basically you're gonna stand up to people and say, no, this is what I think, and then keep on going even if you're proven wrong and then try to adjust what your experiment has shown. - Yeah, that's right. And when you make something that's unexpected and it's a little bit out of people's expectations, they're first gonna reject it. And actually, that's one of the strengths of science. You really have to say, no, it's not because I said it is, but you're gonna have to convince them, and by convincing them, it's really through discussion and additional experiments, because in the end, the experiment is gonna be the final arbitrator. There's no high priest or priestess of science that says, no, you're right, you're wrong. You go back and you do some more experiments. So the reaction, if you're a little bit off center, or a lot off center, is no, that's preposterous. You've gotta be wrong. And the more outlandish you are, the more unexpected the finding, the more you're gonna get that reaction. Now in the end, after one really understands what's going on, and it goes back to really understanding the science, then you say no, no, no, it's all right. Yes, we could've foreseen that, and that's where it becomes trivial is the, well, sure. But it wasn't trivial at the beginning, but after you see it, then it becomes easy. (Harry laughs) But that's actually a mark of really understanding something. Then they say of course. And then the final one is you're not the first to discover this. That is also true. (laughs) There are always precursors. (both laugh) And there's always someone before you, had a glimmer of this and a glimmer of that, and science is based upon a lot of rediscovery. But going back to your point, namely you're gonna be rebuffed, and oftentimes rejected, and it's not a personal issue, and you just gotta stand up to it and go back. Now you could be wrong, but you're gonna go back and convince yourself you were right. The rule I tell myself and my students is once we convince ourselves, we have to be our worst critics. And once we convince ourselves that we're right, then we should have no problem convincing everybody else we're right. And so a good scientist is their worst critic. They're always trying to prove themselves wrong, which is hard, because sometimes you've got and idea and you think you're really right, and you have to force yourself. Well, where are the weak points of this argument or the weak points in my experiment? - One of the points that came out in your lecture for me, a non-scientist, was really the importance of collaboration, not only with your own students, but with other scientists, and even other scientists who are in subgroups within physics, but even beyond that to scientists in other fields of science, that those sets of communications working together are really important to push this process along. - Yes, that's another misconception that many people will have about scientists, or doing science and learning science. This misconception is you go to school, you take classes, you study. Years and years of study. You learn everything there is to know in a certain subfield, a very narrow certain subfield, and then you do work in that area. And that is a form, but it's rarely taken, and it's especially not true for the way I do it. Maybe it goes back to my high school days when I was not such a good student, but in actual fact, if one wants to go into a new area beyond your school days, you can do the same. You can pick up a classic textbook or something like that and begin to read the literature, but it's not as much fun. And when I was going into biology maybe a dozen years ago, I did try that. I picked up this big fat tome called Biochemistry, a classic textbook, and I started reading it. It was, I don't know, 1,500 pages. I got to page 150 and I was deciding, well, it's beginning to slip out of my head as fast as it's going in now. (laughs) I reached some steady state, and so I said, well, this isn't gonna work. So I would look around, and I had something for reading newspaper science, Science Times in the New York Times or Scientific American, things of that nature. So I had some interest in these biological problems and I would pick something that I was interested in. But, of course, since I wasn't an expert in biology, I didn't know, is this a stupid question? Is this a deep question? So I'd say, well, I think I can do something here. I have some interest. So I'd go over and trot over to the biology department or the medical school, say is this something worth studying? I think I wanna do this. And they would tell me, sometimes no, no, it's silly or it's been done before, or sometimes they'd say no, no, this is the central problem in biology. (both laugh) That rarely happens. But what happens is then I would start to collaborate with these people who spent their career in this specialty, and their students also, and they grew up in this culture, and they would say, you should read this article and that article and that article, and we would talk, and it was wonderful to learn that way. So you can leap frog over the years of school. Now, to be sure, I'm not pretending I have as broad or as deep knowledge of that, but then you start with a little thin sliver of a particular problem and you start to build knowledge around that thin sliver, and by the time you've done the experiment and you're starting to write the paper, you better have some knowledge of what's around because you won't even get to publishing the paper because you wouldn't have referenced the right people of the precursors before you. But it's learning in that way, and then you go back to the books, but now you use the index, and you say, I wanna learn about this. So now I've begun to teach my students. Many of my students are physicists wanting to go into biology, and I say, okay, we'll use the index. This is the kind of problem. Why don't you look here? Read these five pages in this book and these 10 pages here and these 15 pages here. By that time, you read this review article, and within a month, you're reading the primary literature. - In biology? - In biology, right, without the three years of courses. And then within a few months to a few years, then you're beginning to get a feel for it. It's very important that you get this feel because you have to ask the right questions. One of the most important things that a scientist does is ask a question that's important and that has a chance of being solved. You can ask important questions, like how does the brain work? But that's not sufficient. You have to pick a part of that question to where you can make a contribution, but a serious contribution, and something that others would be interested in. - Before we talk about some of your research in a way that I can understand, and maybe the public, too, in your career, there was a period when you went to the Bell Labs, and I wanna understand how that contributes to your research, and understand exactly what the difference is between being in a place like Bell Labs versus being in the physics department at Berkeley or Stanford. - Right, well, the reason I went to Bell Labs, I was actually here as a graduate student at Berkeley and I was a postdoctoral fellow here, and after two years of that, they actually made me a faculty member in the physics department, but this was a bit unusual because I had spent eight years at Berkeley, and I was essentially toilet trained here, and I had a very narrow vision of science, and what a department really wants is to bring in people from different sorts of cultures. But they decided they wanted me as a professor. It was a beautiful place, so I accepted. But then they did something very unusual. They said, well, you can start your group and go about your business, or, because you've spent so much time here, and this was your only real experience, you also have an opportunity to go somewhere else for a year or two. And so I thought, well, that's wonderful. I have a job at the best physics department in the world here, and so I'll go off and spend some time and broaden myself. And so I decided to go to Bell Telephone Laboratories, which, at the time, was one of the premier research industrial labs. When most people think of industrial labs, they think of, oh, you're making some better widget. You're making something that's gonna be good for the phone system. Now ultimately that's true, but at Bell Labs in that time, and this is 1978, they allowed a small fraction of us, 50, 60, 80, to do whatever we wanted, really to do whatever we wanted, and so I joined Bell Laboratories, and my department head said, well, Steve, you can do whatever you want. It doesn't even have to be physics. All we ask is that you don't go to a high energy accelerator and do high energy physics because that would be hard on the stockholders. (both laugh) And because my thesis project and what I worked on as a postdoc did have something to do with high energy physics, Not something to do, it was addressing a high energy physics problem, and he said, and by the way, don't do anything immediately. Spend six months, talk to the people around the labs, and just keep an open mind. This was a devastating experience for me (both laugh) because the freedom to do whatever you want and being told don't do what you think you want to do now, but explore. And so I'd spend some time exploring, and there I really felt pressure because he would say, we expect great things out of you. And I don't wanna hear that. (laughs) It's much nicer to have a little problem and you're working on it. It's very cozy. But it did have a real influence on me because it got me in that mode of going and talking to people outside of my field and when I finally started doing things at Bell Laboratories, I started first in some area that was in condensed matter physics that I knew nothing about, but using techniques in my own field, atomic physics and laser physics. But it got me into the mode of, I've got this crazy idea, and going to some colleague at Bell Laboratories and saying, well, how does this sound? And they would tell me, no, this is the stupidest thing I've heard, or yeah, maybe you have something there. And it really set the tone for what I've done for the rest of my life, and collaborating with people, especially outside my local expertise. And so it was a wonderful experience. I also should say in the years I was there, '78 to '87, there was an economic slump in the mid-'70s. Bell Labs just started hiring people and there were a group of us, maybe a few dozen, two or three dozen, and we all were young, energetic, bright-eyed, bushy-tailed. We were all being put in this position. Do something important. Here are the resources of American Telephone and Telegraph System. We expect you to do something wonderful. We were there at night. We were there in the weekends. We knew each other, what our cars looked like. We knew who was in there, let's say, on a Saturday or Sunday. We would party together and it was usually the salad days, and I think something on the order of five or six of us got Nobel Prizes. Over a dozen are in the National Academy of Sciences. It's like we all were growing up together and then we had these really wonderful senior scientists there as well, and it was a remarkable period of time, when everything was exciting and something would come along that was not in my field, and I would say, wow, this is really interesting. We'd go and we'd discuss it, and then people would jump fields or jump areas, and there was this feeling of the excitement of the science and even though we were doing this, it was all right to move and do that, and you wouldn't be considered a failure because you gave up this because something else even more exiciting came along, either from your own laboratory or from a colleague's lab or from the outside world. - So freedom in the best sense, but in an environment where it could lead to new levels of understanding. - Yeah, it was a positively electric atmosphere. You'd go in the lunchrooms and over lunch, everybody went there around noontime. You'd sit there in these big round tables and, okay, what's new? And people would leave, other people would come, but you'd be sitting there chatting, socializing, but talking a lot about science and a lot of ideas were invented on those lunchroom tables. So there again, it was something where there was this real community. It was pretty magical and I think the world, people who are close to science, especially in the areas that Bell Labs was touching, knew that there was something magical going on in that time. - And how can we distinguish that experience from, say, being at Berkeley in the physics department or being at Stanford? Was it just a question of there were enough resources to bring all these people together to create this magical moment? - No, there was some other things. For example, in a university like Berkeley or Stanford, you're a professor, you have students. And because you have graduate students, first you have students as undergraduate students, and part of your job is to teach undergraduates, part of your job is to teach graduate students, you teach graduate students by giving them and developing with them their own projects. And so a lot of energy and time is spent nurturing students and because of that, your first duty is to look towards your group. Well, at Bell Laboratories, we didn't have groups. The biggest you could have would be a technician and a postdoc, and usually not both. And so if you wanted to something that required more than one or two people, you would have to work with other people. And so that builds into the system let's collaborate, and because no one had an empire, or even a mini-empire, that in the basic science areas at Bell, you're one and two or three, you have a lot of time. You're not taking care of people. You have a technician or a postdoc. So it's a very different structure. We are trying to do something like that at Stanford, in which we, in a multidisciplinary way, are bringing people interested in biology, physics, and chemistry, computer science, center around biological problems, but where people from very many disciplines will come and have their own agendas of what to do. One of the things is to limit the size of the groups. You could have, in a university, let's say you're doing synthetic chemists, you could have groups of 40, and with a group of 40 people, you're not gonna have much time to interact with colleagues and you're not gonna have much time to explore elsewhere. And so it's limited the size of the group to 15, which is still very large. You couldn't limit it to three, because there are very few professors who have two or three graduate students in something related to the biological sciences. They're typically more. And so the structure's slightly different and so it's not clear how much you can create the structure we had at the laboratories because you have these other responsibilities and duties. - Before we talk about this new link that you're working on between physics and biology, let's talk a little about your research that led to the Nobel Prize, and if you could briefly give us a sense of how you came upon that problem set and what you in fact did. It's clearly related to atoms, lasers, cooling, and so on. - Well, I was at Bell Laboratories. There are two main branches at Bell Laboratories. The main research branch was in Murray Hill, New Jersey, and I think it was in 1983, a director in Holmdel asked if I would become a department head in his division in Holmdel. The director, by the way, is Charles Shank, who's the director of LBL, and he said, well, Steve, why don't you consider coming down and starting a department that would be really a basic science department here. Holmdel had excellent science and laser engineering and a lot of the great things that have come out of optical communication were spawned at Bell Labs, in large part in Holmdel and Murray Hill. So I said, that sounds like a great job. I went down and started this department and started hiring some people, and also inherited some very talented people. Actually, two of them are here also, Daniel Chemla, for a brief moment, was in my department. He's now a division leader at LBL, and Geoff Voelker, who's a professor in electrical engineering, was also in this department. And there was another really wonderful scientist there named Arthur Ashkin, an older department head, and I started talking to him casually, kind of in the hallways, and he had this dream, wouldn't it be nice if you could hold onto an atom with light? And he had tried to pursue this dream in the early '70s and mid-'70s, but it wasn't really working. They did some very key experiments demonstrating the fundamental forces, but it wasn't looking like they were getting closer to really holding onto atoms with light. And so finally, the manager said, well, it doesn't look like it's gonna come and you gotta move onto other things, and he said all right. But then I came on board and I was this new young person who he could corrupt (both laugh) with his dream. - You go do this, yeah. - And I started thinking, okay, this sounds pretty interesting, and started having a look at it, and I started doing a lot of calculations. It was getting to be bad. I was thinking, I can see why you quit. (both laugh) I was thinking it. I didn't tell him that, but I was thinking to myself, no, it's not looking good. (both laugh) There's a few eureka moments you can have in science. Mostly they're gradual eurekas, which I can come back to later, but there was this time and it was not looking good at all. I would try it this way, that way. It's all on paper, not looking good, and finally, there was a big snowstorm in New Jersey. They said over the PA system, the forecast looks very bad. There's going to be nine inches or something. The lab is going to be closed. You should all go home. Now I live very, very close to the lab so I said oh heck, and everybody left, and it was very quiet, and it's one of those beautiful things where you could see the snow falling down and everything's turning fluffy white. And maybe it's appropriate because it's snowing, and then I found you can do an end run. The thing there was first you hold onto an atom, then you get it cold, and then you can really do what you want with it. And then I said, well, what if you reversed it? What if you cooled down the atom first? Don't hold onto it, but maybe in the process of cooling it down, it's gonna hang around for enough time that you could have a chance of grabbing onto it. And so a little calculation, and he says, holy smokes, this looks like it's gonna work. And then I said, well, I wanna refine the calculations. So then this was gradual eureka. I wanted to refine the calculation. So you're surrounding this atom with light and it looks like it's getting very cold, to the point where your feeble little trap that was gonna hold the atom could work, but it needed to hang around for awhile. So I said, tomorrow I'll come in and I'll start to write a program to predict how long it would hang around. I start to write the program. Luckily, I'm not that good at writing programs and get through about three lines. (Harry laughs) Because if I was really good at writing programs, I wouldn't have thought about it at all and just written a program. So I get the three lines of code and said, I've seen this problem before. Einstein solved it (both laugh) in 1905. - Good old Albert. - And what he did is he looked at a dust particle in fluid and he was studying Brownian motion, and here's this dust particle and it's being battered from both sides by atoms and molecules, and he said, okay, if I take this particle and I move it, there's a viscous drag in the fluid, and that slows it down. And the reason it's being battered around is because of this random imbalances between pressure from the left, pressure from the right. And what I wanted to calculate was how this particle would wander around. Because the previous day I'd shown, it has this viscous drag on it and that you still have these fluctuations. I was gonna write this computer program to say, okay, stepping to the right, stepping to the left, and balance all the forces. And I said, no, I can use his solution. I know where it is. It's in an elementary textbook, an undergraduate textbook. It's the random walk in a Brownian medium, and you just plug in those numbers, and voila, you get, wow, it's hanging around for a long time because it's a random walk. And so I got very excited and went to my boss Chuck Chang and said, look, Chuck, I know you're not keen on this after years of research, but this is so simple, it has a shot at working. (both laugh) And then you can get it cold, you can hold onto it, and we can go from there. And then I remember he thought about that and said, well, okay, you earned the right to do something crazy. But don't try to recruit someone else. (both laugh) So I said, okay, okay, just my postdoc and my technician and me. Again, because if you're onto something really big, you wanna bring in your friends and say, okay, we wanna go fast and we wanna do this. So we puttered along for a few months, going like the blazes, and then I talked to Art about it. He went hmm, hmm, okay. It's not the way I envisioned it, okay. Sounds promising. Then after a few months, it began to look like it was gonna work, really gonna work, and I said, come on, join in. (laughs) And it's gonna be a lot of fun. And, again, as I indicated in my lecture, it worked much better than anybody expected. - And what are the implications of what you discovered in a layman's way of understanding? - Well, what you can do once you get an atom very cold, and cold is really the average speed that an atom moves. The atoms in this room are moving at speeds of supersonic jet planes. In fact, that's why the speed of sound is what it is. It's determined by the speed of the molecules. Okay, once you get an atom really cold so it's moving as fast as an ant walks, that a fraction of an inch per second, then very, very weak forces can push them around, and you essentially can do what you want with them, for example, using electric or magnetic fields or light, and you can hold them. You can push them around. You can do things that you simply could not do when they're whizzing around like supersonic jet airplanes. And the ability to hold onto and control and manipulate these atoms means, for example, you can toss them up. They can turn around due to gravity in a vacuum can where there are no other atoms around, and you can make better atomic clocks. You can make what are called atom interferometer. You quantum mechanically split the atom apart so one part is the quantum wave going in one region of space. The other part of the atom is the quantum wave going in the other region of space. That atom interferometer can be used to measure the acceleration due to gravity or rotations with very high accuracy. In fact, in terms of acceleration due to gravity, better than any other way of doing it, and in terms of rotations, certainly better than any commercial or even laboratory grade laser gyroscope. So all of a sudden, you can measure changes in gravity so accurately that it's gonna become competitive with the current ways of measuring changes in gravity, which is used for oil exploration. You can probably put it on an airplane or helicopter and with global positioning satellite to tell you the height and changes in distance, and inertial sensing systems, and something that measures change of gravity over distances on a scale of a meter, it opens up the opportunity to map gravity gradients and pockets of oil, diamonds, things of that nature, minerals, on a very fast moving platform like a slow moving plane or a helicopter. So they have some real practical implications. Already the world is on the atomic clock standard defined by so-called atomic fountains of atoms. The atom interferometer was totally unexpected. It just popped out after we begun to realize, let me say that, people, even the researchers in the field, it's hard for them to think about what you can do with it, even if you force yourself, until you have it in hand, and you can then begin to see the abilities of this new method or technique, and then it's only after we had it, and not only me, my group, but the world in general, no one was talking about many of the applications that came out until we actually had it and we saw how powerful it was, and then began to appreciate it. And you can try to force yourself to think of what might come of it, and you can write down a few things, but you're gonna get a small fraction of them, and that's the wonderful thing about science. - And actually it hearkens back for me to what you had said about your high school experience in a funny kind of way, that sort of learning to look at something and think logically about it, and, wow, you're sort of taking it to, not that you were doing Nobel Laureate work in high school, but some of the elements are there in this work. - Yes, I think that's true. But it's always also just letting something happen. This is one of the things I did learn at Berkeley, and I watched great scientists here, and many of them were doing something that in hindsight looked very natural. They would say, here's an emerging technology. With this emerging technology, can I ride piggyback on it? And can I use this technology to turn it backwards and do some new science? Normally you think, oh, basic new science discovery. It turns into a technology. You make a better widget. But what I appreciated when I was a graduate student here was that's all true, but you can also take that technology, turn it around, and you can use it, and a good example is radar. During World War II, the US and Great Britain especially developed microwave engineering methods to have microwave transmitters that allowed radar so we could measure and see things far away. And the scientists who helped develop that radar and other scientists who could see the power of that technology really seized on that shortly after World War II, and a string of Nobel Prizes came out of people who use this new technology to do great science. Charlie Townes here is a prime example of that. His knowledge of microwave science, he was, during the war, working on microwaves. - [Harry] At Bell Labs, too, I've heard. - At Bell Labs, that's right. After the war, he said, I want to do microwave spectroscopy because here's a new tool. We now have control of shortwave radiation, and he became one of the real leaders in microwave spectroscopy, wrote a classic book with his brother-in-law Arthur Schawlow and then invented an idea of stimulated emission of microwaves called a maser that then led to the extrapolation of those ideas from microwaves to optical wavelengths led to the laser. And so that's one example of first building on technology during the war, saying this is a wonderful way, a new scientific tool. Use it to do science. Then wanting to improve the tool to get to shorter wavelengths and, voila, you have the laser. And I saw this over and over again. When Charlie came to Berkeley, he wanted to use his knowledge in lasers and microwaves to do astronomy. So again he was gonna ride this technology and I was looking around and saying yes. Now, he's a brilliant scientist, but the lesson I learned was you don't even have to be brilliant (both laugh) if you're the first to look at something with a new tool. So I began to say, okay, what are the new tools? When I was a graduate student, there was something called a tunable di-laser, and I told my advisor, this is a wonderful thing. It's only a few years old and it's sort of like this is a tool that we should be using. Now let's go figure out some science to do with it. And luckily, it turned out fortuitously that there turned out to be some very fundamental physics questions you can address using this tool. And it's the fundamental physics that drove him, but from my side, yes, I was drawn to the fundamental physics, but also let's use a new thing to do it. If you use an old tool to tackle a problem, you've gotta be really smarter than the rest of the folks. Because everybody has this tool. If you're the first to look at something new, it's like discovering a new world. You just look around and everything you see is going to be new. - So bring this now to where you're moving now, because you, in a way, are going into biology. You already touched on that earlier, but you are bringing physics to the table of biology, so to speak, and tell us a little about what you're seeing, and you quoted Yogi Berra yesterday about that, that it's amazing what you can see. - It's amazing (laughs). Well, Yogi is one of my heroes. As I've mentioned, he's really the great American philosopher for the 20th century. - (laughs) That's right. - And one of the things is he said you could see a lot by watching. He said other wonderful things like, "If you come to a fork in the road, take it." - That's right. - Or, "We may be lost, but we're making great time," and many, many things he said. (both laugh) But anyway, so my entree into biology was exactly what I was telling you about. I was working on atoms, cooling atoms, holding onto atoms with light. I said, well, the same technology can be tweaked a little bit, and we can maybe hold onto individual molecules with light if you play some tricks. And what could you do with these individual molecules? And then naturally I thought of biology. So naturally I thought, well, let's first try to hold onto a piece of DNA. So strictly a technological thing. And so that worked, and then as I had some ideas of looking at enzymes, proteins walking up and down the DNA, and seeing what can learn in biology. But the first thing I did is I got this thing and we glued little plastic spheres to the ends of this big DNA molecule, so big that its length was something like 15, 20 microns, which could easily be seen in an optical microscope. Now you can't really see a piece of DNA because sideways, it's only 20 angstroms. It's not resolved by an optical microscope. But you do a trick. You put little dye molecules, fluorescent molecules. So think of it as a string of Christmas tree lights. You can't really see the string. It's too thin, but you see the light coming from that, and so it shines very nicely in an optical microscope and you can move it around. So the first thing we did is, okay, let's stretch it out. Wow, it stretched out. Well, let's see if we can break it. So we stretched it, stretched it, harder and harder, and we couldn't break it. It's very strong along that dimension, which is good because it holds your family jewels. You don't really wanna break it that way. And so in the end, what happened, it pulled out of the optical tweezers, these things we were using to hold onto these plastic handles we glued onto the ends of the DNA, and it sprang back like a rubber band. It just went boink and it just crumpled back up and I said, holy smokes, it looks like a rubber band. Why does it look like a rubber band? And the reason it looks like a rubber band is because when a molecule's straight, it's in very unlikely state. If left to its own devices, it's being battered around by water molecules. It wants to sort of do some random coil geometry. That's a much more likely state. And so the reason it springs back has nothing to do with chemical bonds and forces pulling it back. It has to do with whether it's more likely to be found in some random coil or straight. It's the same reason if you push all the molecules of the air into a corner of the room and let go of them, they don't even have to bounce on each other. They would say, where would I likely be? Well, equally likely anywhere in this room, and so the pressure evens out very quickly. And that's why it was springing back like a rubber band, and I said, well, now I can do this on a single molecule, and what things can you do there? Well, you can understand polymers. Polymers are long, skinny molecules, but you can look at it one at a time. And so it was a backdoor entree into polymer physics, and we did that for a half dozen years, and finally I started getting back into biology because the things we were finding out in polymers, well, this is unexpected. They all are behaving differently, even though they're put in identical situations. They should do the identical thing, but they don't. Well, maybe biology's not as simple as we thought. And so I said, okay, let's go back and look in biology. But it's the same thing. It was a little technical trick or two, but usually when you're doing a technical trick or two in the past, you say, well, I don't really know biology. There's a lot of fancy names and it's a different culture, and so I'll develop the technique and hand it over to the biologists. But at this time, I wanted none of that. (both laugh) Why should they have all the fun? - And you're seeing even in these early stages of your work that you were seeing a kind of, what did you call it? A molecular individuation. - Right. - As opposed to looking at these things and reaching a conclusion on averages. - Right, right, and so, that's right. While doing polymer experiments, we found that molecules act as individuals. In fact, they acted as individuals with moods, meaning you think you start with the molecules, it is precisely the same situation. You ask it to stretch. It would stretch in one way with a particular geometry. You put the same molecule back, same conditions, ask it to stretch again, and it would do something differently. And we realized finally with doing computer simulations that the reason it was doing that is because it wasn't exactly the same because it's bouncing around. It's in water, Brownian motion, and so when you ask it to do something, it's sort of trapped in its particular state, and if you ask it to do something fairly rapidly, it doesn't have time to look around and find the best path. It just does what it has to do. Imagine you're going down, you're in the Newark or Japanese subway. It's very crowded and there's two subway trains, okay? And one's the right way, one's the wrong way, but all you know is the doors are gonna close very quickly. There's a mob pushing you from behind. You're not gonna make the right decision. (both laugh) You're gonna go with the flow. The fork in the road comes, you're gonna take it. And depending on whether you're a little to the left, a little to the right of this, you're gonna be pushed forward into one of the cars. And so the same thing was happening in molecules. Depending on the initial starting condition, you're gonna take a certain path that gets magnified. You start with a slightly different random starting condition, you're gonna take another path. And the profound thing that was affecting me, it's not that. That, once you think about it, is trivial, but that many things in biology, this is an out of equilibrium process, and many things in biology at the molecular level might be out of equilibrium also. And then the way to look at it is to look with these methods and then to think of the non-equilibrium parts. And because we were trained to think of an equilibrium because equilibrium things are things we could measure easily with the techniques that we had before. But now if you can follow a single molecule and say, okay, non-equilibrium is a major part of this. So now we get to look at it, and we can look at how molecules change their shape in real time. Well, so again, it's going back to what I learned here at Berkeley. Use some new technology and have a first peek. - One final question requiring a brief answer, it's all been a random walk for you, then, right? - Oh, absolutely. Right. - And so what recommendation would you give to students looking on your life, reflecting back on it, if they wanna prepare for science? - You've gotta be very interested in what you're doing. You have to attack it with a passion. You can't give up. You have a plan, but if during the execution of whatever plan you had, something comes along, keep an open mind. - On that note, Steven, I wanna thank you very much for coming back to Berkeley to be the Hitchcock Lecturer, but also thank you for coming to our program and sharing your intellectual journey with us. Thank you. - Thank you. - Thank you very much for joining us for this Conversation with History. (electronic music)
Info
Channel: University of California Television (UCTV)
Views: 53,218
Rating: 4.9078341 out of 5
Keywords: nobel, Steven, Chu, research, science
Id: y-7gWsoXtUw
Channel Id: undefined
Length: 57min 14sec (3434 seconds)
Published: Thu Apr 24 2008
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.