Clifford Shull - Nobel Laureate Lecture in Physics at MIT 1994

Video Statistics and Information

Video
Captions Word Cloud
Reddit Comments
Captions
[MUSIC PLAYING] PRESENTER: Well, in case anyone here doesn't happen to know, Cliff Shull is the 1994 laureate in physics. [APPLAUSE] Cliff is sharing the honor with Bert Brockhouse from McMaster. And the Nobel citation, in fact, quoted Cliff as being the one who told us where the atoms are, and Brockhouse telling us what they do. Cliff was born in Pittsburgh in 1915. There is a-- perhaps apocryphal-- story that his great skill and intuition in the laboratory came from his family hardware and home repair business in which he was quite active, I understand, as a youth. Got his degree at Carnegie Tech in 1937, and went off to NYU for his graduate studies. There, he got into the beginnings of what you might call nuclear technology with an MIT connection, as during his thesis time, he built a Van de Graaff generator. And of course, Van de Graaff was a professor here at that time. He did not, however, work with neutrons at that time. And in fact, during the war, as well, did not get to work with developing neutron technology, as he was too valuable in his work with Texaco-- or what is now Texaco-- working on high-octane aviation fuel during the war. This time was productive, however, of course, in a different way. In 1941, he married Martha, with a very important partnership-- 50-plus years and going on. He did get to work with X-ray diffraction techniques, which presumably played a major part in seeding his subsequent activities with neutrons. After the war, Cliff went to Oak Ridge, and using the large graphite reactor-- which was developed there-- began his study of neutron techniques, working on that for many decades, and of course, resulting in the prize which he deserves today. I will not get into the work with neutrons. You'll hear that from Cliff. You'll hear some further remarks from other people. In fact, I'll just end with some departmental personal remarks. First of all, I think that Cliff's colleagues have a deep appreciation to Cliff for his many contributions to the department, his pleasant aspect, his dedication to high standards and to re-doing experiments. And everyone is extremely delighted with this recognition. Indeed, Cliff won The Buckley Prize-- a very prestigious prize in condensed matter physics-- in the mid-50s, and this is a 0-year overdue recognition by the Nobel Committee. Cliff, with his characteristic modesty, I think is genuinely surprised by the news he received a few days ago. And I would just say that many of us were far less surprised than he was, since we all feel very, very strongly this was a very timely and a very, very well-deserved prize. So we're going to have a few remarks now before Cliff gets his chance to tell you about neutrons and how he helped develop this important field. Chuck Vest, the President, will say a few things. The Dean, Bob Birgeneau, who is one of our most active, certainly, neutron-scatterers today will add something. And then finally, one of Cliff's students, Dave Moncton, who has flown in from Argonne, will give a brief perspective from the students' view. So, Chuck. VEST: Thank you very much, Ernie. I think that moments like this are extremely important in the life of a great institution like MIT, and certainly are extremely important in the life of science period. It's a great honor to be associated with people who stand for such excellence in their scientific work and in their being as humans. There are two things that particularly struck me as I listened and talked to people-- I must admit, from a distance, because I've been out of town the last two days-- and had to sense all this through the newspapers and the television. And running into a couple of people, including a former physics student here from the Department, in the Washington area. The two things that struck me the most were first-- every single person I have talked to about this award, within the first one or two sentences, has used the word "nice." Professor Shull is known as a nice person, which is another way of saying a warm human being and somebody who is viewed as being very supportive of his students and his colleagues, and to the Institute community. The second thing that really struck me was a quote in the newspaper. It may have been from Bob Birgeneau. I'm sorry if I'm attributing it to the wrong person-- that said that they had worried that the revolution that Professor Shull and his colleagues brought about in the use of neutron scattering to measure properties of materials and so forth had become so standard, so well-known, so much the stuff of textbooks that they were afraid that the person who originated it might be forgotten. That is a very strong and very important statement, and I think that one of the terrific benefits of the Nobel Prize recognition is that it reminds people-- and particularly young people-- that there are people behind science, and it is their passion, their enthusiasm, their commitment to excellence and belief in the importance of science that really drives it forward. So we thank you very much-- to use a rather hackneyed phrase, I suppose-- for being such a fantastic role model for those who follow behind you, Professor Shull. And it's just a very, very great honor to be a member of this MIT community and to share this moment with you. Thank you very much. [APPLAUSE] Bob. BIRGENEAU: As many of my friends out there are know, this prize is a really extraordinary pleasure for me, because actually, I began my research career as a summer student in Bert Brockhouse's group, Chalk River. And it was because of Bert Brockhouse-- who also is a nice, wonderful person-- so this is extraordinary that we had two such special human beings recognized with one prize. And in fact, my first publication was with Brockhouse's group, and embarrassingly, it's still my most referenced publication. [LAUGHTER] After having got launched by Brockhouse, I then meandered around the world, actually, and then finally ended up here at MIT. And while I was learning about this field from Brockhouse and Brockhouse's former cohorts, they had told me how there was this genius who had created the field in the first place by the name of Cliff Shull, who was then at MIT, so it was one of my goals to come to MIT and to meet the great person himself. And I remember first walking down the hallway the first time I saw Cliff, and Cliff was so nice and so self-effacing, I couldn't believe-- especially in the MIT environment-- that this modest person-- [LAUGHTER] --was this great hero that I had heard about. But indeed, it was. And actually, we had many years together. We served on a number of academy committees together, flying back and forth to Washington. And we shared one extreme vice, which is both Cliff and I like to smoke cigars. And Cliff had the wisdom to retire before it was banned. But I presume-- actually, there's still a residue in his office if you go into his office, which is now occupied by my and Marc Kastner's post-docs, who are hoping that some of this greatness rubs off on them. Besides being a wonderful colleague and an outstanding scientist, Cliff also was excellent as a undergraduate and graduate student teacher, and as a graduate mentor. And he educated many of the leaders in the field. One of the premier of those is Dave Moncton, who's Associate Director at Argonne National Laboratory, and who is Head of the Advanced Photon Source, which is this wonderful new source which we and many others hope are going to advance the kind of science that Cliff has done-- albeit using protons instead of neutrons, in that case, a level further. So maybe Dave can make a few comments from the perspective of someone who used to sit here as a graduate student. [APPLAUSE] MONCTON: I'm really overwhelmed by the privilege of being able to stand up here, have the opportunity to thank Cliff personally in front of this particular audience for all that he gave to me as a student. I'd like to speak mostly to the students in the audience. I guess the faculty members can listen in if they'd like. The first contact that I had with Cliff was when I was still a senior at Cornell. I had applied to only one school as an undergraduate, and realized that that might be risky as a graduate student. So I had applied to a number of places, and I was fortunate to have gotten accepted at all of them. And I was fortunate, also, to have gotten scholarships at all of them. So I was having a hard time making up my mind until a letter arrived in the mail, and it was a personal letter from Cliff, who had been busy in the graduate school office looking over the applications of graduate students and personally writing to those people he thought would fit into his laboratory. And there was no question in my mind, when I read that letter, that this was where I wanted to work. I don't think I took my cap and gown off before I got in the car and drove from Cornell to Cambridge and began working in the summer of 1970. I think the 1st of July, essentially. I was so excited about a real laboratory and real experiments that I couldn't think of anything better to do that summer. That's an example of Cliff's fastidiousness, I suppose, that has become legendary. And all of his students will tell you about that, and it's a gift that I think he's given all of his students. The other night, my wife and I were arguing why it was that when I go to sleep at night, I like to have all the three-way switches in my house all in the down position. You don't need to have them all in the down position. As long as they're up in pairs, then the lights will all still be off. [LAUGHTER] But it seemed to me that that was not fastidious enough. I mean, after all, if someone came in in the middle of the night and turned on a light and turned it off, I mean, you'd know if you remembered where you'd left them. [LAUGHTER] So throughout my career as an experimentalist, I always tried to emulate. I never successfully emulated that level of fastidiousness that Cliff brought to bear on experimental science, but I never regret it. Any time I did something, I lined a goniometer up to be a 0 of angle when it was perfectly vertical or perfectly horizontal. It didn't have to be defined as 0 there, but that just gave you some clue as to whether it had been moved inadvertently, or the motors had slipped or what have you, because you could line it up with various other vertical clues in the horizon. And those were the kinds of things that Cliff just did routinely. During my-- I think early in my second year, Cliff thought it would be a good exercise for a student to write a Nobel Prize recommendation. So he's told me that he would like to see Brian Josephson get the Nobel Prize, and would I go off to the library and take a crack at writing the case for Brian Josephson? Which I was very enthusiastic about, and I did that, and I gave it to Cliff. And he puckered his lips as he usually did and held his pipe out and said thank you. Of course, that was the last I saw of it until Brian Josephson, a few months later, got the Nobel Prize. And I thought, "That was easy!" [LAUGHTER] Little did I know what work goes on in getting Nobel prizes-- on either side of the equation. The one other gift-- and the most important gift that I think Cliff gave me as a teacher-- was I guess what I call the gift of self-confidence. As an experimental physicist, it's very important-- there are lots of bleak moments. There are long periods between good results, and even longer periods between excellent results, and longer between publishable results, and longer even between results that people want to read about when you publish them. [LAUGHTER] And in those long waits, you have to be confident that you have something to bring to bear on the field-- something to offer. And you need to have a lot of confidence in yourself. And what Cliff did was, to a large extent, he left me alone, because he had what I believe-- he can correct me when he gets his chance-- what I believe is a lot of confidence in me. I didn't work on a problem that Cliff gave me as a thesis advisor, and that was very unusual. And I didn't even work in his own lab, finally, on my own thesis. I went to Brookhaven and worked at their reactor. And throughout the entire period of my graduate school, he was-- especially in the later years-- almost a secondary figure in my day-to-day life. But this tremendous confidence that he had in me translated-- some magical fashion-- into my own self-confidence as an experimentalist. So it's not the-- I don't know-- the book learning, in a sense, that I took away from my graduate education with Cliff. I took it away from MIT, because goodness knows we had many, many courses. But what Cliff gave me-- and I think all of his students-- was that self-confidence that he himself-- and you probably won't get him to admit it, but he has it in great quantities. Quantities enough that he is able to share it with so many of us. So thank you, Cliff. [APPLAUSE] PRESENTER: OK, without further delay, Cliff will tell us about the development of neutron scattering. SHULL: Thank you, gentlemen. As Ernie has mentioned-- and Bob-- the field of research that I want to mention this afternoon goes back a long ways; back to the immediate years following World War II, back in the late 1940s. And this research deals with the development of techniques that can utilize slow neutron radiation. Now, neutrons were first discovered in 1932 by Chadwick in England. And Chadwick found that if you take an alpha particle source-- and essentially, the only alpha particle sources known at that time were from naturally-occurring radium sources. If you surround a radium sample with some material-- most usually beryllium-- you find that, emanating from the beryllium, is an intense radiation alpha. And that was later investigated by Rutherford and other people-- classic people-- and labeled alpha particles. Now, it was also found, shortly after Chadwick discovered-- when Chadwick was arranging such configurations-- alpha particles sources-- he found that if you further surrounded that source with other media, that there seemed to be something coming away from the other media which was certainly different than the alpha particles that were originally released from the inner piece of radium. And on investigating this, he found that it was a very penetrating radiation, and that was the surprising thing to him. It's radiation that could pass through copious thickness of the material, and there was no known radiation at the time that would do that. So with further studies of the production of this anomalous radiation with different materials surrounding the alpha particle source, he came to think of this as being some neutral particle type of radiation. The reason being that a neutral particle would be expected to have much greater penetration through media than any sort of charged particle radiation. It was ruled out that it was an electromagnetic type of radiation from certain observations. So if it had no charge-- if it were radiation without charge-- you could think of it as being somehow neutral in an electrical charge. And he gave it the name neutron-- the neutron radiation. Well, this was very exciting to nuclear physics-- what there was of nuclear physics at the time. Nuclear physics was very rudimentary then. And further studies-- mostly by Fermi in Italy-- showed that this radiation could be effectively thermalized by, again, passing it through and letting it bounce around in some other medium. And if you observe what comes out of this third piece of medium, he obtained what he thought-- what he classified-- as thermal neutron radiation. How could he determine that it had this sort of characteristic? By being thermalized, it implied that the energy was down to thermal energy levels-- thermal energy levels that characterize atom emission of radiation-- thermal energy radiation. Well, one thing-- it was known that captured cross-sections-- absorption cross-sections of materials-- would be expected to follow a certain energy dependence. The lower the energy, the greater the absorption. And that, indeed, was what he found-- that if he put more and more media there, did more and more acting upon this neutron radiation, it lowered the energy of the radiation so that you could observe a larger absorption cross-section in something else. So this process could continue until he obtained what he thought was real thermal energy neutron radiation, where the neutrons had been degraded in kinetic energy down to a thermal energy level of some fraction of one electron volt-- some fraction of one electron volt. So this thermalization concept was a very important one. It was shortly recognized that if you had radiation of such a low energy, that this radiation would also exhibit a very anomalous wave character. And this goes back to the heart of quantum mechanics. A particle moving along is to be considered also as a moving wave. And there's a periodicity associated with this wave. This first was suggested in earlier years, most notably by de Broglie-- de Broglie-- who suggested that there was a wave length associated with moving electron radiation. Moving electrons should also exhibit a wave character. And it wasn't long before people did experiments that demonstrated sure, indeed, a moving electron does exhibit a periodicity and a wave length, and that was demonstrated by diffracting low-energy electron beams from crystals, just as people with X-rays had demonstrated many years earlier. So there was this weight periodicity associated with moving electron particles. Now we have neutron particles that are moving slowly, and they should have a characteristic periodicity also. The question comes, do these neutrons also show diffractive effects, the same as had been observed with electrons some years earlier? So there were two experiments-- two experimental groups-- that set out to establish-- to look for this periodicity; this de Broglie wave character for this neutron radiation. Von Halban and Preiswerk in France, and Mitchell and Powers at Columbia University in the United States. Both pursuing experimentation in 1936. And they both found, indeed, that this periodicity was there-- that there were diffractive effects-- excuse me. --from crystalline material when this neutron radiation-- thermal neutron radiation-- was applied to these diffracting objects. So now we have a wavelength. We have a kinetic energy. We have an energy-- a kinetic energy-- to be associated with these particles. Well, the experiments of demonstrating this diffractive effect were so complicated, and the intensities that were involved were so small that one could hardly think of exploiting it as a diffractive tool. X-ray diffraction was in a well-defined, well-developed state by that time, and many people were diffracting X-rays and learning lots of things about crystals and materials in general. One could think that maybe neutrons could also serve that purpose, but the intensities were frighteningly low, and nobody worried about doing anything seriously in that direction. That all changed a couple years later in 1939, when nuclear fission was uncovered by Hahn and Meitner in Germany-- 1939. Now, the exciting thing about nuclear fission is this. You bring neutrons onto certain fissile nuclei-- most notably uranium; normal uranium-- and you find that the neutron absorbs the neutron and breaks up into fragments, and more than two neutrons come off from that. One neutron in, more than two coming off. Now, this conclusion didn't come automatically. It came with a series of different types of experiments. But that was very, very interesting. I put in something, and I get over twice that same thing out. That immediately suggests that this process could be used in a cascade manner to produce more than you feed in, and the process could skyrocket, cascade, and become cataclysmic-- with more coming out than you're putting in. Well, it wasn't long after that that we became engaged in war years, and it was proposed-- thought very seriously by people-- that maybe this could be developed into some sort of weapon-- an atomic weapon. Maybe you could have a configurational system that could be triggered somehow or other, and it would just diverge. So many people became involved in that-- Fermi himself, and many of his colleagues, and very well-known people became involved in that through what was called the Manhattan Project, set up in the early war years. It was finally demonstrated, in experiments at the University of Chicago-- at the Stagg athletic field-- that, indeed, you could have a cascading system-- a reproducing system-- that you can think of. The idea was to collect enough uranium and enough graphite together. The graphite was to serve as a moderating medium for the extra neutrons that came from the uranium fissioning, and the graphite would cascade the energy down until those neutrons would then cause further fission events in other uranium. Giving, again, two for every one-- more than two for every one. And that process would continue and build up. Well, the chance of success of obtaining such a configuration, we recognize, would increase as the physical size of this configuration would get larger, because volume effects go up as the cube of the size, whereas surface area-- which determine leakage from a system-- go up only as a square of the area. So there's an advantage in increased size. So the test was going to come when one could get a large enough configuration of uranium and graphite collected together into a system. This was obtained, and it was detected. It was very carefully monitored with sensitive equipment in this first experiment at Stagg Field at the University of Chicago. That occurred in December 1942. December 1942. Once it had been demonstrated-- once Fermi had demonstrated that, indeed, you could have a self-reacting system this way-- the experiment was shut down. They didn't want it to get out of control. Shut down, and people immediately started to design facilities for using that. The thinking at that time was to construct a reacting system off somewhere else that would produce some isotopes which were badly needed during the war time years-- most notably isotopes of plutonium, which were considered to be useful in atomic weapons production. So they wanted to build this test reactor-- which would operate at a considerable power level now-- as a production test facility for even larger facilities which were to be built elsewhere. It was decided to position this test reactor at Oak Ridge, Tennessee-- when the army had absorbed a large area there for industrial use. And this reactor was put on paper. Construction was carried through, and was completed in an amazing period of 11 months-- December 1942-- until the first use of the reactor in November 1943. 11 months in total. And this was all 500 miles away from Chicago, where the first experiments were carried out. So a whole flock of physicists were conveyed and sent from Chicago down to the Oak Ridge area-- and engineers galore, and construction people. And this amazing development was pushed through. So this reactor started to operate at Oak Ridge in November 1943. And I'll show, shortly, a photograph of that. But let me first project this. This is a chronological listing of the starting periods of the very early piles-- they were called piles back "Reactor" as a term came into vogue later on. Piles. This Oak Ridge graphite pile-- which I was just describing its start-up date was in 1943-- at the same time as the construction of this reactor was going on-- designing and construction. Another one was under construction back in Chicago, in what was then called the Argonne Laboratory. And it differed from the Oak Ridge graphite pile. This should have actually been written Clinton graphite pile. Clinton is a town in Tennessee which is the nearest town to the Oak Ridge facility. So at that time, this was all Clinton, and Oak Ridge came in later. The Argonne reactor was to differ from the graphite pile in the Oak Ridge area by having heavy water instead of graphite as a moderating medium. The heavy water was to slow down the neutrons and cause, again, fissioning in uranium-- the same as in this first pile, which had graphite as the moderating medium. At later times, in '47, the Chalk River reactor came into operation. And this is the one where Dr. Brockhouse did all his research work-- up at this Canadian government facility. And then later ones in England. This is at England, and another one at Oak Ridge, and at Brookhaven on Long Island in this country. My next projection shows a photograph of this graphite reactor at Oak Ridge. The reactor is a very large assembly-- a cubic assembly, roughly 20 feet on an edge-- of stacked graphite-- pure graphite-- components, what were called graphite stringers. So these graphite stringers were square, configurational chunks of pure graphite, and that was a problem to prepare a graphite of sufficient purity. About 4 or 5 inches on a side, and 3 or 4 feet in length. Rectangular assemblies, just stacked together. On the corners of these graphite sections-- graphite pieces-- there was a diagonal cut. And so when these elements became stacked together, there would be channels running down along the length. In these channels, one was to place the uranium fissioning fuel. The uranium was pure metal uranium. Many problems associated with preparing a pure uranium. And the uranium had to be encased, because uranium will oxidize. And if it gets hot-- which it would in this operation-- then it would oxidize very rapidly and become uranium oxide and become a powder, and essentially collapse. So each of these uranium pieces, which are roughly an inch in diameter and 5 inches long-- an inch in diameter, 5 inches long-- had to be encased in an aluminum shielding can, and sealed up to prevent air from getting to the uranium and oxidizing the uranium. So we think of this configuration-- of all this graphite, tons of it-- stacked up. Not bolted together, or not clamped together, because the graphite would become hot and expand, and any banding or clamping might be broken. And so all of this graphite with these uranium rods running through the assembly. Much uranium. I don't have a figure for the mass of the uranium or the graphite involved, but you see the scale of it. Now, all of that assembly had to be surrounded by a shield to prevent radiation from escaping out to where people might be operating the assembly or doing something outside. So it had to be encased in a concrete-- essentially, a concrete house the whole way around, the concrete being about 5 feet in thickness. 5 feet in thickness. So we see sections of that around here. We're looking at what's called the loading face of this assembly. You can perhaps see vestiges of the channels that are traveling back away from us in the photo. That would represent positions where the uranium rods were placed, and this loading face is where those radium rods would have been originally pushed back into the graphite assembly. So there's a loading platform here that's on a hydraulic crane assembly. Can go up and down. Here are a couple fellas doing something at the front face, there. Perhaps pushing some uranium on back through the assembly. And the control room is over in here. And for no particular reason, they can look at the front face here. But there had to be shielding on this front face as well. Once you had completed an operation in an area, you had to shield that off. The main shielding was all around here. At the back side of the reactor-- the backside-- one could push these rods-- the uranium rods-- through, and they would fall down into a plenum chamber and could be collected together and transported with remote control facilities over to a chemical processing plant which is next door. The purpose of this reactor was to produce the first viable quantities of plutonium. One of the isotopes of uranium, when it's subjected to this neutron radiation, transforms itself into a plutonium-- the next element up in the periodic table from uranium-- into a plutonium isotope. And from previous measurements and studies of that isotope, that was potentially a very good atomic weapon component. Recall we're in the middle-- in the starting period of the war-- and everybody was desperate for new developments, here. So the primary purpose of the reactor was to produce sizable quantities that the chemists could get at and purify and study processes, and the physicists could get out and study processes with the plutonium for weapons potential use. So there was no experimentation as such with all this neutron radiation which was bouncing around inside this facility. There was one experiment that I know of that was performed, and that dealt with, again, producing information for further use on the project-- what was called a pile oscillator. A pile oscillator. Something moves back and forth. The idea is to have a detector back inside a section of this configuration-- a neutron detector. And you move a sample that you want to study back and forth in its vicinity. You can determine, with this device, the capture or absorption cross-section of this medium for neutron radiation. And that was of great interest in the technology at that time-- again, for the practical uses to which the project was devoted. So there was that one experimental facility, and that's the only use of the neutrons that I know of, other than to continue the fissioning process in the uranium inside the assembly. Well, this was used throughout the war for producing-- war years, until the war ended in the summer of 1945-- for producing quantities of plutonium. As soon as this reactor was operating, new reactors were under consideration for construction out in the state of Washington-- at Hanford. Much larger reactors. Much larger piles, I should say-- which worked to produce more abundant, more copious quantities of plutonium. So those were under construction out there when this was operating for in the first year. But this reactor continued to operate in that way. The total power production in this reactor was about 3,500 watts-- 3,500 kilowatts, excuse me. 3,500 kilowatts. 3 or 4 megawatts of power production. Present day power reactors operate roughly at 100 megawatts-- 25 or 30 times the operating power level of this assembly. I didn't mention-- how do you get this power out of here? Well, the power release was to be obtained by having air stream along these channels that surround the uranium plugs that were permeating the whole assembly-- having airstream along there. So you needed big fans in the exit side to draw air in. So all through this building, there was a sucking sound of air being pulled in here and cooling the reactor. Taking this 3 megawatts of power-- thermal power-- out and warming up the air, and then the air was filtered and sent up stacks. Well, at the end of the war years-- at the end of the war, in the summer of 1945-- one of the research scientists at the laboratory here, Ernest Wollan, who had come down with the original group from Chicago-- Ernest Wollan-- got to thinking seriously-- as other people did-- how can you use these facilities as research tools? And Ernie Wollan decided to see if he couldn't make use of some of these copious neutrons that were rattling around in here, trying to bring some out through a collimated tube, and trying to exploit this diffractive property of the neutron radiation. So he decided he wanted to try to set up a diffraction spectrometer in the way that X-ray people had been doing for 15 or 20 years earlier-- bringing a beam out and defracting it from samples, seeing what novel things might be available from such studies. So Ernie arranged to have sent down from the University of Chicago an old X-ray spectrometer unit that he had used back up there. He did his thesis work in an earlier time with Compton, studying X-ray scattering by gas samples. So in that, he had built an X-ray spectrometer, and he knew that was still up there, and he arranged to have that sent down to Oak Ridge in the fall of 1945. So he and a colleague of his, Robert Sawyer-- who was there during parts of the war years, from University Lehigh-- configured this spectrometer so that they could try to bring a neutron beam out from the reactor, now, off to a side, and have it fall upon this spectrometer. Which, a spectrometer, in general, is something that you can control angular positions. Have a detector on it, and you can move the detector around and determine a pattern of scattering, or a pattern of diffraction. So that was the idea in this first effort. And his spectrometer, I show in my next projection, here. This was the original spectrometer. The reactor is behind here, off to the back side. And there has not been a channel-- an open channel-- through this graphite assembly that lets neutrons stream out-- those that happen to be coming in the right direction-- and stream out through the shield and the face of the reactor. Well, they have to be contained, so there's another outside shield built up around here. So the idea was to place a monochromating crystal in that beam. Neutrons were coming out and falling on a large, single crystal of some sort, and being diffracted from that single crystal. Even if these were different energy neutrons, as they are-- thermal neutrons with a thermal-- one wants to select out a monochromatic section of that spectrum-- a monoenergetic section. And that can be done by diffraction from a crystal. One wavelength. One energy. So back inside this secondary shield was this monochromating crystal. Now, about the only large, sizable crystal that was available to Ernie and Sawyer at the time was sodium chloride. Sodium chloride seems everywhere. And people had to learn the technique of preparing the large crystals of sodium chloride. The infrared people were hot on alkali crystals-- very fine, infrared window material. And so the development people had developed the technique for preparing large crystals of alkali halides-- most notably, sodium chloride. So Ernie obtained a sizable single crystal of sodium chloride, roughly an inch and a half high and 2 inches long, and a 1/2 inch thick-- which is an ungodly size for a single crystal of substance back in those days-- and used that as a monochromating crystal in here. From that monochromating crystal, one would have, then, a pure diffractive beam from it coming out, and that came out through this secondary shield, crossed over the axis of this spectrometer, about which I've spoken. Here is this old, original, University of Chicago X-ray spectrometer bolted down to the platform on the floor, here. And this neutron beam-- purified neutron beam, now-- coming out and crossing the axis of this spectrometer. One would place, then, a further scattering sample or diffracting sample on this axis so that the beam would come in and strike the sample, and now be diffracted around from that sample. So here is a detector. This is the most massive thing on the assembly. The detector itself is not large. It's about 2 inches in diameter and about 15 inches long, and it's a sealed chamber sealed with an insulated wire along the center, and the two electrodes, the wire, and the case, and filled with a special gas-- boron trifluoride. Boron is the important thing. Neutrons are absorbed copiously in boron, and so the idea was that a neutron coming in along the axis of this chamber would be at least partially absorbed by boron nuclei in this gas, and that would trigger an electrical response between the wire and the piece, which were charged to a high potential difference between the two. So the chamber is inside here, and the surrounding here is secondary shielding. The shielding is the massive part of the assembly, here. All of this shielding to keep stray radiation away from that detector. But this whole assembly can be moved around the axis of the spectrometer. Now, this weighs-- this has a mass-- much larger than this spectrometer shafting can support. So to take off some of this loading, there were cables running up to a gimbal joint up on the ceiling, here, so that this off-axis loading-- mass loading-- was supported from this universal gimbal joint up on top, here. Well, this had to be aligned pretty well. And I remember, after I came into the picture, all the troubles that we had in keeping it aligned-- making sure that this operation of angular position here was a smooth operation. But here is shown on here, now, a typical sample-- a sample in this circular sample. And this is a powder sample-- random crystallite direction within the powder. Sealed up between aluminum windows in a plate. The sample thickness-- you don't get the impression here-- is perhaps 1/2 inch thick. The neutrons had to go through perhaps 1/2 inch thickness of the powder sample in the diffracting process. And the ceiling is just to keep the powder in there. They're aluminum windows, there. Aluminum is very transparent to neutron radiation. In operation, here, one had to turn handles. There are some handles, here, turning the detector around. Also, separately, controlling the angular position of the sample. It was all completely hand-operated. If you wanted to study the scattering around here, you positioned this to some point and counted neutrons that were being detected in this detector over some period. That period might be as short as a minute. It might be as long as 10 minutes or 15 minutes. And then after a conclusion of obtaining that information, go and turn to a new position and continue the process. Powder diffracting patterns were very time-consuming in their obtainment. Well, with this sort of assembly, Ernie and Sawyer indeed found that they could take powdered diffractions from specimens. And I show here a powder pattern obtained with a sample of, again, sodium chloride now. Sodium chloride powder. There's a sodium chloride monochromator way up in there. That's a single crystal. This is now the powder sample. This is intensity versus scattering angle position. This was presented in a monthly project report at the Clinton Laboratory in April 1946. The assembly was collected together in late 1945 and early 1946 by Wollan and Sawyer. And here are a couple of diffraction peaks. So these are nice, well-behaved looking peaks with measurable intensity. They take time, and it was a chore to obtain such things, but it was available. Notice the background intensity-- the baseline intensity-- that's underneath all this structure. There's more of it out here, but patience ran out at this time. But notice the level of background versus signal intensity, here. The background is as much or larger than the signal, and that's a troublesome feature, and you'd like to do something about it. Here's another of the very earliest patterns. Ernie decided to look at water-- H2O. And not only water, but heavy water-- D2O, and here are patterns he and Sawyer had obtained in those very early days. Now, the distinction between deuterium-- heavy water and light, normal water. The interest here was to see if one could possibly determine steady, hydrogenous materials-- hydrogen-containing materials-- because-- and this is very important-- such information-- establishing the location of hydrogen atoms in materials and material structures-- was completely not available with X-radiation. Nobody had determined what the real structure of normal water was, or frozen water or any other hydrogenous material, because X-rays were insensitive to hydrogen in hydrogen scattering. That was not expected, but it was hoped that maybe with neutron radiation, there would be a difference enough between deuterium and hydrogen to show a distinction. Well, these first patterns-- which, there was no attempt at analyzing these. But they did show differences, here, of this type that you see. This peak-- you're peaking. And these are liquid patterns, so you don't expect sharp diffraction lines. But this peak is considerably more prominent than for deuterium-- the heavy water-- than for the normal, light water. So this suggests that, again-- that, indeed, hydrogen and deuterium are scattering differently, and that one is sensing that in these diffraction patterns. Well, there was no analysis of this at the time, but that was the purpose of the experiments. Again, this was reported in this same project progress report-- that I had shown earlier-- in April 1946. I came to the laboratory in June of 1946, a month or two after this period, joined the group. And Wollan and Sawyer decided he was going to return to the university. So he came and went back to Lehigh. And from then on, I worked very closely with Ernie Wollan in all of the developments. Well, knowing that you could successfully take powder diffraction patterns, then we just decided to look at many materials, trying to make some sense out of what was controlling these intensities. What were the scattering properties of the different atoms for this neutron radiation? There was something known about that. There had been a compilation of what is called total scattering cross-sections for a series of different atomic nuclear species. But that's not the whole story. That's not the whole thing that's going to determine these powder peak intensities. It was recognized that there are different isotopes for any element-- for an element. Different isotopes might very well display different scattering properties. And eventually, one would have to unravel all of that. Another thing-- many nuclei have nuclear spin, and nuclear spin-- if you have a nuclear spin, then you have two nuclear spin state scattering amplitudes-- one where the neutrons come in this way, and one where the neutrons come in this way to that nuclear spin. So what that implied-- and in a general sample, one would have these nuclei arrayed-- and random, of course. There was nothing to say the spin should be up or down or whatever. Under those conditions, what that implied-- and the theory had predicted this from the very earliest-- that this, what we now call the coherent scattering cross-section, should, indeed-- or may, indeed-- differ from the total scattering cross-section. We had information on total scattering, but nothing at all on coherent scattering. And the coherent scattering arises and this difference comes in because of this spin state-- the possible difference in these two spin state scattering cross-sections, or scattering amplitudes. So in the long run, all of that had to be unraveled, and nothing was known about this distinction between what you were going to see in a diffraction pattern and what would be predicted from these total scattering cross-sections. You had to look specifically at this quantity called the coherent scattering cross-section, and that might be anything from a very low value, up to the total scattering cross-section. So that was the purpose of looking at many of these early samples-- to see if we could unravel and understand these intensities and how they were affected by cross-sections; and, indeed, what the basic cross-sections were. Well, this was all very choresome with this hand operation, and we were all very anxious to try to improve things so that we didn't have to sit there through the night and collect the data by hand and write things down. And we were very fortunate, in 1947, when a couple chaps came down to Oak Ridge as part of what was called the Oak Ridge Training School. After the war, the administration management decided to set up a training school where they would encourage people to come in there from universities or industrial laboratories and learn, become familiar with all the techniques and the new techniques which had been developed during the war years, and which were very secret. So these people were to come in and talk with people, work in the laboratories in many cases, but mostly to attend lectures that were to be given by experts who were brought in for that purpose. So this Oak Ridge Training School-- which had, perhaps, 150 students at some times-- these were senior people from universities or from industrial laboratories that were sent to Oak Ridge for this purpose. And the name Frederick Seitz may be familiar with you. Seitz, who was a very respected solid state physicist, was brought there as Director of this Oak Ridge Training School. Well, two of the students that came in-- one from the RCA Laboratories in New Jersey, and the other from Goodrich Tire Rubber Company over in Akron-- George Morton from RCA and Bill Davidson from Goodrich were two of the students, and they became interested in these neutron diffraction patterns. And in particular, Morton was a very skilled instrumentation man. He had worked with-- name's Zworykin, who was a prime developer of electron microscopes in former years, and RCA was in the business of selling electron microscopes-- first of the type. So Morton was a very skilled experimentalist and instrumentation man. And he took on the task-- between the two of them, they took on the task of automatizing this slow-operating spectrometer. And they made many improvements. We made many improvements, there, this first year-- in 1947. This is the same spectrometer. And now, about a year later than in the earlier photo, here's the basic spectrometer. Nothing changed there. But now here's a nice, automatic, control, automatic operating system, here, that drives a worm, here, with a clock motor and some cam wheels with some microswitches. So that on demand, this would now change its angular position. And at the same time, there was a cadmium shutter assembly that could flop in or out of the beam that was falling on the sample. And the electronics system was improved over here on the control cabinet. And the detector-- there was just general improvement things done to all of the components. Now we had a recorder to print out data. We didn't have to be there to see meters print out data recorder. This was what was called a traffic counter. You've perhaps seen these gimmicks along the highway, where they have a tube running across, and gives a signal, and this thing prints a new number over here on the side of the road. That was the only type of recorder-- digital recorder-- that we could find at the time. All the electronics in the system-- amplifiers, high-voltage power supplies-- those all had to be made homemade. There was no commercial instrumentation of that type available in this period. So there were improvements in these amplifiers. Vacuum tubes had gotten better. This was long before the day of solid state electronics. Tubes had become better and less noisy. But still, here's a preamplifier, and you'll notice it's sitting up here on top of a padded blanket, because it's microphonic. This preamplifier is microphonic. You have to separate it from components, and you don't want to bump it, or else you will get a false signal through your circuit. Well, with these sort of improvements, this changed things in a very great fashion. One can now put a sample on here and get things ready and go off and do something else, and come back the next morning or something-- some later time-- and look at the data, or change something and do things automatically. Hey, analysts, spend your time analyzing things, rather than collecting things. So I can't overestimate the importance of these instrumentation features that came in at that time, and they were due to our having these two gentlemen as part of the training school-- students as part of the training school, there-- that affected them. Well, with this, then, you could think of taking patterns regularly, and many more over a given time span. We were always interested in these cross-sections, or amplitudes-- cross-sections, or amplitudes-- defining just what different nuclear species-- what the real scattering properties were. Because that's data that has to be obtained empirically, and it can't be predicted from theory-- or at least, theory was far behind the stage where it could predict how these things should go. So we were very interested in studying as many samples and looking for consistency from one material to another. Does sodium scatter the same way in sodium chloride as in sodium hydroxide? That wasn't really an established thing. So we set out to study these consistencies, looking at chlorine in different species. And to our dismay, there was much that was discrepant. And this gave us a great deal of worry. We were trying, at that time, to look at the total diffraction pattern-- and you still do that. You look at the diffuse scattering, as well as the coherence scattering which is showing up here. There should be some-- if you combine these two things together to represent the total scattering information that you're getting-- there should be consistency. We weren't finding that. And to illustrate that, let me show this. Here are some observations made on three different forms of pure carbon. Diamond, which is pure carbon. Graphite-- pure carbon. Charcoal-- pure carbon. Now, what you'd first think is that if I put a diamond sample on here, graphite and charcoal, I ought to get the same total scattering if I look at everything that I can see in the scattering pattern. Turns out that that isn't the case. It seemed not the case to us. We were not, when we added everything up in the whole pattern, including a diffuse scattering and coherent scattering and peaks. There was not consistency from one to another. And we looked at other things, as well as this series. And we were having extreme trouble in obtaining what we consider to be consistent information from sample to sample. And I remember a conversation that we had with Eugene Wigner at the time. Wigner was the Research Director at the laboratory, and Ernie Wollan and I bent his ear. And we were telling him our tale of woe, here, about these inconsistencies that we were obtaining from sample to sample. And we were very puzzled about this, and could the great man think of some help or relief for us? Well, he surprised me very much by saying, no, he thought that maybe this is the way it really is. Maybe there's some other processes, here, going on that we just don't understand. And he gave us great encouragement to keep on going, and he wasn't taken aback by the fact that we were getting these seemingly inconsistent observations. Keep on going-- back to the laboratory. Well, it wasn't too long after that that actually, we found the solution. And the solution turned out to be this, and we were both embarrassed by knowing it. Multiple scattering in samples. Multiple scattering in samples. Both Ernie and I had had previous X-ray pattern experience where you scatter X-rays from samples and obtain these patterns. And we never encountered a problem of this type. But here, things are very different. The neutron penetration through samples is so much larger than X-ray penetration through samples there, because the cross-sections are different. In a neutron case we're using samples which are orders of magnitude thicker, larger samples than in the X-ray case. And now you're susceptible to this problem of multiple scattering in the sample block. A neutron comes down to this glob at one position, gets scattered up through the glob-- partly-- and gets scattered again. And the result is that you don't have consistency in the total scattering in your pattern. So once we recognize that feature and that fact, and we're kicking ourselves that it took us so long to recognize it. Then it all became clear to us. We had to place controls on our sample-- physical, geometrical conditions-- which we, of course, did. And then things got clarified. Now we were able to get satisfactory consistency from one sample to another sample with the recognition of this troublesome multiple scattering problem. Well, we had this great interest in knowing whether you could determine hydrogen atoms in crystals. There was this earlier observation that I showed on water, but that was too complicated to interpret. So we searched for crystals that contained hydrogen, and which might have a nice, ordered, crystal structure. But it wasn't known whether the hydrogens were, indeed, ordered in these materials. Well, one of the simplest of the hydrogen-containing crystals is this-- sodium hydride. Sodium hydride, one-to-one. And if you take an X-ray pattern of this, you see only the sodium and you know nothing about where the position or what the hydrogen atoms are doing. But the X-ray pattern shows a well-developed pattern characteristic of a sodium lattice. That's known very nicely from the X-ray pattern. But the question is, what are the hydrogen atoms doing? And that's what the neutrons can hope to show. So here are two patterns of the same position in the diffraction pattern from two isotopically different samples-- sodium deuteride, and sodium hydride. And you see the patterns are completely different. So this says immediately that you're seeing the deuterium or the hydrogen in these patterns. If these were just looking at sodium atoms, they ought to be the same. Why the difference? Well, the difference turns out to be that the algebraic sign of the deuterium scattering is opposite from the algebraic sign of the hydrogen scattering. And that's why the two patterns are seemingly reversed, here. Big peak here, little peak. Little peak, big peak. And that persists through more of the diffraction pattern. So sure enough, here what we're seeing is the real determination of the hydrogen atom positions in these two crystals. This was the first and real indication that one could, and this is a very important feature. Nowadays, there's a lot of biological structure work that goes on with neutron diffraction in neutron diffraction studies, looking at determining where hydrogen atom positions are in organic polymeric, biological materials. That's an important field of application. There are other applications that I could go on here, but I'm afraid I've gone too long here. But let me close by showing one other pattern, here. This is a different type of neutron pattern. This is what's called a Laue pattern. In this Laue pattern, one uses a direct beam of radiation from the reactor-- not being crystal monochromated-- so that you have all neutron energies, all neutron wavelengths falling on a crystal-- a single crystal. Under such conditions, the different wavelengths and the different atomic planes in the crystal diffract the radiation to specific directions, and one obtains this spot pattern, rather than a peak pattern as in a powder peak assembly. So we have these Laue spots here. Well, this was not the first picture that was obtained. There were quite a few attempts before getting one that looked even as good as this. And this has several interesting features about it. One is that first of all, it's a photograph. There's photographic film. And now, photographic film is insensitive to neutron radiation. So you use a trick, here. You put a piece of scintillating material next to the film. Neutrons come in and are absorbed in the scintillator. In this case, it was a sheet of indium metal. The indium then releases-- after capturing neutrons-- releases a charged particle, and that charged particle reacts nicely with the photographic emulsion. So the sensitivity comes from this combination of scintillator and photographic film. And there was an indium sheet, here, covering the film. And there was a section of that cut out in a whole circular region to let the direct beam weight it so it wouldn't scintillate and over-cloud the whole film. And we've obviously gotten it off-center. The whole assembly is off-center in the exposure. But there are two other features in here. One-- if you look closely at these spots, you see-- notice that they all have a double structure. It's particularly pronounced here. There's a double. And that puzzles us very much. Why should there be this doubling of neutron spots? Is that perhaps connected with the fact that there are two spin states in the nuclear scattering from this crystal? Well, we were anxious to explore that and see if we could explain the origin of this doubling. And as it turned out, after this exposure was made-- and this exposure took 20 hours, a long-time exposure; 20-hour exposure-- just after this exposure was completed, the reactor had to shut down for breakage of one of these fuel element cans. One of the cans containing the uranium had developed a pinhole, and the uranium inside it oxidized, and it had essentially exploded inside the reactor. So there was no way of operating the reactor under those conditions. They had to get it out of there. And that took some three weeks for them to cleanse that and get back to operation. So it took us a long time to ponder over this before we could hope to try to do something else to try to explain this doubling of these Laue spots. Well, it turned out, after we got back in business, that the spot structure was something associated with a particular crystal we used. The crystal had a mosaic layer on its two faces-- entrance and exit. And a mosaic layer means that the crystal scatters more strongly. So what these spots represent are, effectively, scattering from the front side of the crystal and the back side of the crystal. And that's, somehow or other, showing up all around here. So that was a trivial explanation. A good crystal without this mosaic layer on the surface doesn't show this. But another interesting feature-- you notice that there's a band structure, here. Well, there's no way you would expect that. That turns out to be the world's first neutron radium graph feature. The band structure represents the collection between pieces of indium. We didn't have enough indium, or in large enough area to cover this thing, so we had strips that we kept together by strips of scotch tape. Scotch tape line. Scotch tape line. Another one here. [LAUGHTER] So this is showing the radiographing of the Scotch tape that was used to hold these indium strips together. Well, that set us off to thinking, what else could we do? If a piece of Scotch tape shows up here like this, there must be better uses that one can exploit in this. Indeed, that's the case. People now radiograph large assemblies, metal castings, or other things with neutrons, in the way that normal radiographing is done, and exploiting the particular penetration properties of the neutron radiation. Well, these were very exciting times back then, and I often think really how exciting it was. Almost any corner you turned, there'd be new vistas showing up. And we explored many, many corners at the time. But on reflection, I'm sure that there's no difference today than back then. Almost anything that one can look at, there is new information and new things to uncover. So I think the challenges in present-day physics, present-day science loom just as important, as prominently as they did back at this time. So I say that to our younger friends who are with us here. Thank you very much. PRESENTER: Neutron scattering. Cliff's stories are rich. He actually only began to scratch the surface of all his many contributions-- magnetism, polarized neutron technology, basic neutron properties. I guess remain for the next installment of that. [LAUGHTER] Cliff has agreed to answer some questions, so we have time for perhaps a few, at least. Questions or comments? Bob? BIRGENEAU: Cliff, when and how did you first get the idea that you could look at antiparallel magents with this technique? SHULL: That came a year or two after these things that I've chatted about. And we were set off on this route by our desire to look for what was called paramagnetic scattering. Paramagnetic scattering was something that theorists had predicted would occur from a material-- a paramagnetic material. And at that time, a paramagnetic material-- as it is today-- represent a collection of magnetic centers, where there was randomness in the direction or the orientation of these magnetic centers-- paramagnetic material. In contrast to a ferromagnetic material, where they were considered to be all lined up. Now, in paramagnetics, with a paramagnetic sample, if one scatters neutrons from it, you expect to find a diffuse scattering in the pattern associated with this random collection of magnetic centers. This was shown in a very powerful theoretical paper by Halpern and Johnson back in-- years earlier than this period I was talking about. And so this was a prediction, and one that was capable of observation with the technology that was available here. So we were interested in looking for this paramagnetic scattering in a paramagnetic media. And the first thing that we looked at was in a material-- a typical paramagnetic substance-- hematite. Fe2O3-- hematite. There are strong magnetic ions-- iron ions-- in this substance. And it was known that you could make or you could measure paramagnetic susceptibility and obtain certain values of the susceptibility, and could interpret that-- its temperature dependence-- upon the strength of these magnetic ions, which were at random. So we looked at that first. And that was present. We could see that. We also looked at another one-- manganous oxide-- where, again, you have a strongly-magnetic manganese ion present in the material. Manganous oxide-- monovalent, strong, magnetic ion. And that, again, has a nice paramagnetic susceptibility. And we looked at that, and what we found there was there was, indeed, this diffuse scattering, but it showed some structure in the diffuse scattering. Instead of showing a uniformly varying paramagnetic scattering, it showed some structure. And that suggested-- I had had experience with short and long-range order in crystals prior to that time. That suggested that if we were to cool this medium down-- the sample down-- that this short-range order might develop into coherent long-range order. And sure enough, when we finally did that experiment, by cooling the manganous oxide down to a low temperature, the long-range order had developed. So the origin of the magnetic scattering all came in the interest of searching for the paramagnetic effects, and then the ordered magnetic lattice scattering followed from that. But that came in 1948, '49, 1950-- a couple years after these experiments that I've just been describing. PRESENTER: Other questions? Looks like we're coming to a certain symmetry since you started with the Clinton pile. And of course, now one hopes to build the new reactor in Oak Ridge, and if it succeeds, it will certainly be the Gore-- the Gore reactor. So it will be the Clinton-Gore reactor in history. [LAUGHTER] The only difference is the 11 months it took for the Clinton reactor. It'll be 11 years at least for the Gore reactor. With that, let's thank Cliff once again. It was wonderful. [APPLAUSE]
Info
Channel: MIT Video Productions
Views: 453
Rating: 5 out of 5
Keywords: MIT, video, education, science, math, business, robotics, massachusetts, institute, of, technology, school, college, university
Id: 92t9XLMexCs
Channel Id: undefined
Length: 96min 36sec (5796 seconds)
Published: Fri Jul 27 2018
Related Videos
Note
Please note that this website is currently a work in progress! Lots of interesting data and statistics to come.