Oral-History:Alexis P. Malozemoff: Difference between revisions

From ETHW
Line 31: Line 31:
<br>PLACE: Charlotte, NC
<br>PLACE: Charlotte, NC


===Introduction===
===Early Life and Education===


'''Hochheiser:'''
'''Hochheiser:'''


This is Sheldon Hochheiser of the IEEE History Center. It is the 13th of August 2014 and I'm here at the ASC Conference in Charlotte, North Carolina with Dr. Alex Malozemoff.  
This is Sheldon Hochheiser of the IEEE History Center. It is the 13th of August 2014 and I'm here at the ASC Conference in Charlotte, North Carolina with Dr. Alex Malozemoff.  
===Early Life and Education===


'''Hochheiser:'''
'''Hochheiser:'''

Revision as of 20:13, 16 October 2014

About Alexis P. Malozemoff

Alexis P. Malozemoff was born in Santa Rosa, California and raised in New York and Connecticut. He received a Bachelor of Arts in Physics and Chemistry from Harvard University and a Ph.D. in Material Science Engineering from Stanford University. Beginning in 1971, Dr. Malozemoff served nineteen years at IBM Research as a staff scientist, manager, and senior manager in the study of magnetism, condensed matter physics, and superconductivity before joining American Superconductor (AMSC) in 1991, where he served first as Vice President for Research, and then from 1993 until his retirement in 2009 as Chief Technical Officer.

Over his more than thirty years of experience at IBM Research and AMSC, Dr. Malozemoff is recognized for his work in the field of superconductivity in the co-discovery of the “giant flux creep” and the irreversibility line in high temperature superconductors (HTS), as well as his developments in magnetic bubble technologies and HTS wire and applications, and his numerous publications and patents. Following retirement he continues to be active in the magnetism and superconductivity community serving as chairman of the Conference of Magnetism and Magnetic Materials, and and serving on various national committees and centers of HTS research and development.

In this interview, Alexis P. Malozemoff discusses his early interest in engineering and applied sciences. He goes on to describe his introduction to the field of magnetism and superconductivity and his experiences in research and research management at IBM Research, and his technical leadership role at American Superconductor. He reflects on both the accomplishments he achieved and the problems he encountered over his lengthy career, which spans over both theoretical and applied research and applications. Despite the promise that the field of superconductivity shows, he comments on the difficulties of the commercialization of a product produced in a still developing field.

About the Interview

ALEXIS P. MALOZEMOFF: An Interview Conducted by Sheldon Hochheiser for the IEEE History Center, August 13, 2014.

Interview #657 for the IEEE History Center, The Institute of Electrical and Electronic Engineers Inc.

Copyright Statement

This manuscript is being made available for research purposes only. All literary rights in the manuscript, including the right to publish, are reserved to the IEEE History Center. No part of the manuscript may be quoted for publication without the written permission of the Director of IEEE History Center.

Request for permission to quote for publication should be addressed to Oral History Program, IEEE History Center at Stevens Institute of Technology, Samuel C. Williams Library, 3rd Floor, Castle Point on Hudson, Hoboken NJ 07030 USA or ieee-history@ieee.org. It should include identification of the specific passages to be quoted, anticipated use of the passages, and identification of the user.

Alexis P. Malozemoff, an oral history conducted in 2014 by Sheldon Hochheiser, IEEE History Center, Hoboken, NJ, USA.


Interview

INTERVIEWEE: Alexis P. Malozemoff
INTERVIEWER: Sheldon Hochheiser
DATE: 13 August 2014
PLACE: Charlotte, NC

Early Life and Education

Hochheiser:

This is Sheldon Hochheiser of the IEEE History Center. It is the 13th of August 2014 and I'm here at the ASC Conference in Charlotte, North Carolina with Dr. Alex Malozemoff.

Hochheiser:

Okay. If we could start with a little background, where were you born and raised?

Malozemoff:

I was born in Santa Rosa, California; grew up first in the New York area, then Connecticut, and then went on to high school at Phillips Academy Andover, undergraduate studies at Harvard and graduate studies at Stanford.

Hochheiser:

Okay. Were you interested in science and technical things as a youth?

Malozemoff:

Yes, I had many interests. I wasn't so clear exactly where I would go but as I got further into my education, science became more and more important, and already in high school it was a focus of interest for me. I majored in physics and chemistry at Harvard.

Hochheiser:

Was that a double major or an interdisciplinary one?

Malozemoff:

Yes, it was a special double major.

Hochheiser:

Was that what you had in mind when you went to Harvard?

Malozemoff:

By the time I entered Harvard I already was thinking about science as a main interest. I had many other interests but my fascination with science kept growing and I enjoyed it and was excited by it; so it was clearly the direction that I was headed.

Hochheiser:

What led you from there to graduate work at Stanford in general and in material science engineering in particular?

Malozemoff:

Stanford was another wonderful school.

Hochheiser:

Absolutely.

Malozemoff:

I did want to have a change of venue. Moving out to California, I had a lot of nice experiences there; it was a beautiful place. For my degree, I considered many different options. Of course, physics was a possibility, chemistry was a possibility, but I was already leaning a bit in the direction of having exposure to engineering and applied issues; and so the Material Science Engineering Department of Stanford appealed to me, and I signed up with them. I had access to all the other technical courses at the university and I took a wide variety of such courses, but my PhD was in material science engineering.

Hochheiser:

What was your dissertation?

Malozemoff:

It was on the optical properties of orthoferrites, a particular group of magnetic insulators. My thesis work focused on their magnetic properties. And magnetic properties of materials became a theme of my entire career, and it continues to be to this date. These were very interesting materials, and my thesis already involved the combination of both experimental and theoretical work; so I had a foot in both camps, which also became a theme in my entire career.

Hochheiser:

Who was your advisor?

Malozemoff:

Bob White, R. L. White. I guess an amusing detail is that when he went on a sabbatical during my last year of my PhD work, the substitute professor who helped guide me through the last year was called Bob White! This was R. M. White, also a very talented guy, and it worked out fine.

Research and Development at IBM

Hochheiser:

What led you to IBM Research on finishing your dissertation, as opposed to some other opportunity?

Malozemoff:

I was always interested in fundamental physics and material science but also in applications, and this biased me towards an industrial environment. I did a number of interviews before I left for my post doc. I did a post doc at the Clarendon Laboratory in Oxford, which was also a wonderful experience where I extended the kind of work I had done for my PhD in the area of oxides, studying the Jahn-Teller effect in some rare earth oxides. This was a very fascinating area, but by that time I had already accepted an offer from IBM Research at Yorktown Heights in New York State. They had been recruiting me at Stanford. I had a very good record there and I already had a good contact with an IBM recruiter. I did also have the possibility of going to Bell Labs. I remember I interviewed also at Oak Ridge National Laboratory and another couple of places, but somehow IBM smelled right to me. It was a great research lab already but obviously on the up and up. Bell Labs, at that time, was perhaps the premier lab, but IBM was rising to challenge it, maybe equaling it, and in some areas exceeding it. There were a lot of wonderful people there and I just felt I could make more of an impact there, and it worked out really well.

Hochheiser:

What group did you join when you started working at IBM?

Malozemoff:

It was a group under the manager Steve von Molnar, another wonderful individual. The group had a wide range of activities spanning a variety of magnetic and oxide materials. The branch of work that I coupled into was in the magnetic bubble area, and soon I established a link to a very great theorist who was working at IBM, John Slonczewski, and in collaboration with him and others we did very exciting work. It was an amazing period; we discovered so many novel properties of the magnetic bubble materials. IBM Research also had excellent materials synthesis capability and there was a wonderful thin film crystal grower, growing garnet epitaxial films, called Ed Giess. We immediately had a great advantage, looking at the best materials in the business. These bubbles are magnetic domains, cylindrical domains with a magnetization in one direction (up) in a sea of oppositely-oriented magnetization (down). They could be moved around, and they were of great interest for a magnetic memory.

I was in the Physical Sciences Department which is the more fundamental end of R&D at IBM Research, and there were applied groups with whom we had a lot of interaction. Our group was exploring the fundamental properties of magnetic bubbles, and John Slonczewski already was theorizing about some of their dynamic properties and recognizing that there could be a phenomenon, called Bloch lines, in the domain walls of these materials. The idea there is that when you have a domain wall separating up and down magnetization, the magnetization must rotate to get from one domain’s magnetic orientation to the other. It can rotate in a variety of ways. If you look at the orientation of the magnetic vector in the middle of the magnetic domain wall it could be pointing in any direction in the plane of the film. That means that you could have a structures inside the domain wall in which the magnetic vector in the middle of the wall rotates from one orientation parallel to the domain wall to the opposite orientation. This creates a linear structure inside the domain wall called a Bloch line, running vertically through the thickness of the film. John Slonczewski had some ideas about how such Bloch lines might affect the dynamic properties, how they might be nucleated in dynamic situations.

At that time a remarkable discovery really got this whole field going; we noticed in these materials some magnetic bubbles which were a little bit bigger than most of the other ones and they were also very ornery. We called them “hard” bubbles. They would hardly move. If you applied a pulse of magnetic field gradient, most magnetic bubbles would move straight down the field gradient, and that property underlay their ability to be used in a magnetic memory, but these hard bubbles hardly moved at all. Often they moved in the direction orthogonal to the field gradient. At first, we just couldn't figure it out. I remember we had discussion about these hard bubbles with the Bell Labs people and the Bell Labs guys said, “Oh, you IBM guys just don't have any good materials; that's just a defect.” We were convinced it wasn't a defect; it was something special that was occurring.

I'll never forget one time, over a weekend, when I woke up on a Sunday morning and I'd had a dream about a helix, almost like a double helix, and somehow my mind was associating it with these hard bubbles. I thought, oh my goodness, if the magnetic vector can rotate in the plane of the domain wall, just go around and around and around and around and around, winding up like a big spring, there would be an exchange energy in such a spring which would push the bubble’s domain wall out, making it bigger. That was the conceptual breakthrough, and I developed a model for the bubble’s increased size. John Slonczewski soon pointed out that when you have these kinds of structures inside the domain wall, their helicity causes a kind of gyromagnetic force: if you try to push the bubble forward, it would instead move sideways. This led to an explanation of the peculiar behavior of these bubbles, and from that point on the whole field broke open because we realized the bubbles could have one twist in their domain walls, or two twists, or three twists, and so on – different numbers of Bloch lines. We found ways to detect the different kinds of bubbles, to show how they got formed, and to explain their dynamic properties. The bubbles even showed ballistic properties as if they had mass. All sorts of fascinating things came out of it and we had a wonderful time researching this. It was very exciting coming up with all these new phenomena, and ultimately John Slonczewski and I wrote a book, which I think has become the standard in the field, Magnetic Domain Walls in Bubble Materials. That was a very, very exciting time in my career.

Hochheiser:

Now did these lead to any commercial developments?

Malozemoff:

There was very active work both at Bell Labs and at IBM in developing magnetic bubble memory, but I have to say that ultimately the advances in magnetic recording swamped the initial advantages of bubbles, and I don't believe they are used anymore. This in some sense ended up being an academic effort in terms of its value, but it's beautiful physics with a lot of amazing phenomena. In addition to Bloch lines, you could have Bloch points, which are point-like singularities, the properties of which one could study. There's a lot of neat stuff there.

In 1979, I took a sabbatical year from IBM, which offered this option, and I spent a year in Germany at the Max Planck Institute of Metal Research, the Institut fuer Metallforschung in Stuttgart. That was also a wonderful year. What an opportunity to start looking at other areas and other research fields. There had been a lot of work in amorphous magnetic materials, and I became interested in these and did some work on these kinds of materials when I was in Germany. I did some light scattering work and a variety of other things. When I got back to IBM, I moved into a management position, but in the Physical Sciences Department there was a tradition of managers combining a certain amount of technical work with their management responsibilities.

Hochheiser:

So you were able to continue to do research.

Malozemoff:

Indeed.

Hochheiser:

I've spoken to quite a number of Bell Labs people for whom the movement to management pretty much meant they were managers.

Malozemoff:

There was certainly a lot of work to do to keep the funding going and all this kind of stuff, but I was able to continue, at least half time, sometimes more, on the technical work, and I had a lot of interesting projects during this period. The work on amorphous materials got me into a very fascinating area called spin glasses, which have a lot of interesting properties. We were studying the susceptibility, finding a way to scale the data and demonstrate a phase transition. We had a number of visitors; we always had a very strong visitor program from different countries. I worked with some of these visitors to address these issues.

One of the projects actually led to my second most cited paper, understanding the behavior of magnetic layers and in particular the combination of an anti-ferromagnetic with a ferromagnetic film. In such structures, the hysteresis loop of the ferromagnetic material was very often displaced by what was called an exchange field, and this phenomenon was often called exchange anisotropy. This is a very interesting phenomenon which is very relevant to the magnetic recording industry, but the nature of it was not understood. There were various theories about how it arose. When you have an anti-ferromagnet, equal numbers of spins are oriented in one and the opposite directions. If somehow more at the interface pointed in one direction rather than the other, you would get a net coupling which would be like an internal field acting on the ferromagnet. But no one could explain the size of this effect and a lot of its characteristics.

I'm proud of an idea that I had which led to a nice theoretical paper, in this case, exploiting ideas about randomness, which I was very familiar with from the work on spin glasses. The idea was to realize that surfaces are never perfect. They can be bumpy, rough, and if they were rough there might well be a region where there was an extra layer, and here you might have more spin of the anti-ferromagnet at the interface pointing one way rather than the other. If the interface had been uniform and perfect, there would have been an equal number of interface spins pointing in each direction and then no net exchange field acting on the ferromagnet. In a given area of the rough surface, a given domain size, just by random statistics you would have a net coupling, which would go as the square root of N by typical Gaussian statistics, where N is the number of interfacial sites in the domain. This recognition then allowed me to understand the exchange anisotropy phenomenon; I was able to explain its size and its many characteristics. That was a very exciting development.

Introduction to Superconductivity

Hochheiser:

Is this now in the early eighties?

Malozemoff:

Yes, mid-eighties.

Hochheiser:

Mid-eighties.

Malozemoff:

Mid-eighties. By that time I was already a second level manager. For a time I substituted as the head of the Physical Sciences Department but most of the time I was second level manager of groups in magnetism and superconductivity and then of condensed matter of physics.

Hochheiser:

Was this then your first exposure to superconductivity when you started managing a group that included it?

Malozemoff:

Yes. That's also leading into the whole superconductor story. What was going on during this time was a major project in the applied area of IBM Research at Yorktown Heights, IBM’s main research laboratory. This was the Josephson computer project.

Hochheiser:

And a very well-known one.

Malozemoff:

Yes. There was correspondingly an effort in Physical Sciences of exploring different aspects of superconductivity on a fundamental level, and this group was in my area at that time, doing very interesting work. To some extent we were exploiting these Josephson junctions for various devices or novel transistors or sensors, but also doing experiments on how to monitor the uniformity of superconductors through clever techniques like laser scanning microscopy. We were making nanostructures in those days long before nano became the hot word that it is today; so it was also a very exciting time. My own personal technical interest had always been magnetism and we had groups on magnetism in my area. John Slonczewski was continuing some of his efforts.

At that time we also entertained one distinguished visitor in our group from the IBM Research Laboratory in Zurich, the Rüschlikon laboratory in Switzerland, and this was none other than Karl Alex Müller, Alex Müller. He came at that time and worked with Mel Pomerantz on some microwave properties of aluminum films which became superconducting below 1 degree Kelvin. Alex was already very interested in superconductivity. His career had been on oxides and magnetic resonance, and things like that. He was already a quite distinguished researcher at the Zurich laboratory. With his interest in superconductivity, we decided to arrange a little self-teach course on superconductivity. We got the book by Michael Tinkham, a wonderful book on superconductivity, and we went through it chapter by chapter and getting together weekly. There was a group of five of us; Mel Pomerantz was one, and a couple of other folks at the lab. I could see that Alex's brain cells were churning there. There were a couple of oxide superconductors which had already been discovered and they didn't have very high transition temperatures, but they had extraordinarily low density of states. Obviously, there was something unusual going on with these materials. He was really fascinated with that and he was aware of the mechanisms of phonon coupling and he was thinking about what could be the way to enhance phonon coupling. One of the areas he had also touched on, just as I had touched on it in my career, was the Jahn-Teller effect, where you get ultimately a permanent distortion of a structure because of an instability, a possible degeneracy in particular states of the lattice, and he was toying with this idea of a Jahn-Teller effect in the oxides as a way of getting a giant phonon coupling, and this thinking is what led him a year or two later to the high temperature superconductor breakthrough. We had a wonderful two years with him during his stay in Yorktown.

Hochheiser:

What years were these?

Malozemoff:

This was '83, '84. He went back to the Zurich lab where Georg Bednorz joined his group.

Hochheiser:

Right.

Malozemoff:

He was, by this time, an IBM Fellow; so he could do whatever he wanted and no one quite knew what he was doing, and guess what? They came up with this incredible breakthrough of high temperature superconductivity in 1986, which of course changed everything.

Hochheiser:

Do you recall your reaction upon learning of the breakthrough?

Malozemoff:

Initially it was very exciting, very amazing, but also there was a certain amount of skepticism because the data were not overpowering. The transition temperature was higher than the niobium-germanium transition temperature, but this transition was very broad, and there was all sorts of talk about filamentary superconductivity. No one really knew what was going on, but it was clear that this deserved further work and the word spread. I think it was at the MRS meeting that December that we learned the Japanese —

Hochheiser:

MRS?

Malozemoff:

It was the Materials Research Society meeting in Boston, I believe, and the Japanese group under Tanaka at the University of Tokyo had succeeded to reproduce Bednorz and Müller’s result. We hadn't yet done that in Yorktown Heights. Then the floodgates blew open and people really started studying this intensively. I remember Praveen Chaudhari, who was the head of the Physical Sciences Department at that time, asked me to take on a new role as the Research Division Coordinator for High Temperature Superconductivity. It was clear that a lot of work was starting. There were three main IBM labs; Yorktown Heights was the main one; then there was Almaden in California, and there was the Rüschlikon laboratory near Zurich in Switzerland. Everyone was going off in all directions and my job was in large part to insure communication to make sure we weren't just duplicating too much; trying to work together so we could advance the whole field.

Hochheiser:

Basically you had efforts in high temperature superconductivity going on in all three laboratories?

Malozemoff:

Absolutely.

Hochheiser:

And your job was to make sure it all meshed.

Malozemoff:

Yes. That was, of course, hopeless, but we had regular joint meetings, people would fly in, we would share information and results, and I think it did a lot to stimulate work and new ideas. One of the aspects of my job was to lead a taskforce on applications: what were the opportunities to apply these materials? In those days it was very early, the properties were hardly known. When we started the taskforce, already YBCO had been discovered and we knew we had superconductors which could work at 77 Kelvin. This taskforce was a very interesting enterprise. We could draw on some significant resources within IBM. One of the people who had joined IBM, working in another area, but had significant experience in large-scale superconductivity, was Bob Schwall. He had worked on MRI, he had worked at Intermagnetics General, and so he was able to supply a lot of input on potential for magnet applications. And he later would join me at American Superconductor. Dick Garwin at IBM had done a famous paper with Juri Matisoo, speculating about the use of low temperature superconductors for 5-gigawatt DC transmission lines, which formed a very nice basis for thinking about how high temperature superconductors could be useful for power transmission cables.

We pulled together a report and there's a nice published paper we have on this. We looked at data processing opportunities and sensors, all these kinds of opportunities. This actually was very important for me personally because it opened my eyes to areas beyond the interests of IBM, which were mostly on the data processing side of things. One of the conclusions that emerged from this study, thanks to Bill Gallagher who was part of this and played an important role, was realizing that getting the high temperature superconductors into a computing environment was a long putt. People talked about potentially using high temperature superconductor lines to connect the transistors to speed up transit times, to reduce losses, and so forth, but this was going to be very challenging. The connections had to be extremely fine and it really was not a hot prospect.

But I became quite interested by prospects for power transmission lines. The early Garwin/Matisoo idea never got off the ground; no utility would want to put 5 gigawatts onto one cable, especially at helium temperature. While it didn't lead to anything, it was a brilliant analysis. Once we began thinking about high temperature superconductors, we realized that now we don't have to deal with liquid helium, we can use liquid nitrogen, and so we have a much more plausible environment for a power grid application, and that looked quite interesting. It opened my eyes to this other area. As time went on I also got involved technically in work on the magnetic properties of the superconductors. I'll talk a little bit more about that.

Hochheiser:

You still managed to find time to continue to do research.

Malozemoff:

I did. I tried to keep the coordination going, but it was difficult to coordinate everybody, that's for sure. They were all doing their own thing and I had a fair amount of time to do my own research and that led to some of the most significant research of my career. While all that was happening, in Japan, in particular at Sumitomo Electric, there was a determined effort, as soon as the BSCCO materials were discovered, to develop high temperature superconductor wires. Initially with the polycrystalline YBCO materials, no one could get any meaningful current through wires, but with the BSCCO materials, lo and behold, the Japanese were producing quite high performance wires. I was watching these developments year after year; and the performance of these wires was getting so good I just couldn't believe it, and this led finally to my decision to leave IBM. Before I get there, let me just say a few words about my work on superconductivity at IBM, unless you have a question you want to interject.

Hochheiser:

No. When it looked like you were about to jump ahead to the next stage of your career I was about to say: wait, we're not done with IBM yet.

Malozemoff:

No, you're not. I had so many interesting areas of work, but this was one of the most exciting of course. The people measuring the magnetic properties of high temperature superconductors were very puzzled initially. There was a lot of relaxation – time-dependent decay – in those magnetic properties, and that was very surprising. Alex Müller had a very interesting idea to explain these phenomena and that was the idea of the Josephson glass, because it was already recognized that the grain boundaries in polycrystalline materials were like weak links; so it was no wonder that you couldn't get much current through these materials. There were all these grain boundaries in there which were acting like little Josephson couplings, but of course random because of the polycrystalline structure, and that led to the idea of a glassy-type of a Josephson configuration. Alex Müller had some collaborators who were doing simulations. Of course, since he had already won the Nobel Prize within the year, his thoughts had to be reckoned with.

We also had the good fortune at IBM at Yorktown Heights, of having an outstanding crystal grower called Fred Holtzberg, who has recently passed away. He grew some of the best YBCO crystals on the planet. Beautiful, single crystals, no grain boundaries, and lo and behold, the magnetic signal was relaxing away, just as in the polycrystalline materials. It was Tom Worthington at IBM who, I think, did the first measurement reporting this relaxation. Then I had a visitor join me from Israel, Yosef - Yosi - Yeshurun, and we decided to look into this more carefully, measuring this relaxation, and then we realized - oh wow! - it suddenly struck us: high temperatures coupled with a tiny coherence length meant that the pinning energies were very low compared to thermal energies. We knew that to get high currents in Type II superconductors you have to pin vortices and the quantized flux lines which they encircle; you have to sustain gradients in the vortex density. So what was happening was that with all this thermal excitation, the vortices were hopping out of their pinning wells, and as a result the magnetization currents of the superconductor were decaying. This phenomenon was known in low temperature superconductors, but there it was a tiny phenomenon, explained by the famous Anderson-Kim model of flux creep. If you did very careful measurements on low temperature superconductors - Mac Beasley was involved in such measurements - , one could detect a little bit of flux creep. But now, in the high temperature superconductors, this magnetic relaxation was a giant effect. In fact, we called it “giant flux creep” and that became a moniker that stuck.

Hochheiser:

The difference is between what's going on in these low temperature superconductors where it's a very small effect and now in the high temperature superconductors an enormous effect

Malozemoff:

Yes, it became a giant effect. Then Tom Worthington was doing susceptibility measurements and he was observing an unusual peak in the susceptibility and measuring it as a function of field and frequency. We started collaborating and putting his observations together with the flux creep measurements, and we realized that the susceptibility anomaly indicated something new. Up to this time, people were interpreting anomalies in the field-temperature plane in a superconductor as an indication of the upper critical field Hc2 where superconductivity disappeared. But this supposed upper critical field had a very unusual shape: It wasn't concave downwards, as in low temperature superconductors; instead it was concave upwards. We realized that, oh my goodness, the higher in temperature you go with flux creep, it gets faster and faster and faster, and eventually there's no pinning at all; you just have the flux lines floating around like in a liquid state. In fact, that idea later became the concept of the vortex liquid. In short, we realized that there was a line in the H-T plane where you lost the ability to pin flux lines at all, even for short times, and so you lost the irreversibility in magnetic properties, which is characteristic of pinning in superconductors. So we called this line in the H-T plane the irreversibility line. This was a term also known in other fields but we wrote a paper on the irreversibility line, which was quickly followed by many papers which were beginning to interpret it with different theoretical models.

One theory was that the irreversibility line was a melting line; the Bell Labs people wrote a paper about the melting line. It was really all the same phenomenon. The irreversibility phenomenon could be explained in different ways and ultimately it was understood as a glassy phase transition. Beautiful work on this was done by Fisher, Fisher and Huse. Matt Fisher, one of the two Fisher brothers and a brilliant theoretical physicist, was also at IBM at that time and we worked together. He developed his famous vortex glass model of superconductivity because pinning is random and disrupts the Abrikosov vortex lattice structure, giving rise to glassy behavior. And then because of the thermal excitation as temperature increased, you got to a transition which, he argued, could be actually a glass phase transition, not a first order, but a second order transition, and then above it you had this vortex liquid. At even higher temperatures and fields, you did have something like an upper critical field, but now there was no long-range order. This was just a thermal bath of these flux lines or vortices, which eventually disappeared as temperature or field increased further; the upper critical field Hc2 was just a crossover to the fully normal state.

This whole picture has revolutionized the understanding of Type II superconductivity. Up to this time, we had had the very beautiful picture of classic Type II superconductivity from Abrikosov’s work, with a phase transition at Hc2. People would draw a now-famous three-axis diagram of upper critical field and critical current density versus temperature, showing the limits of the superconducting region inside. But now we have this extraordinary modification of that vision, where instead of Hc2 we have an irreversibility line which separates a vortex glass from a vortex liquid and then, at higher fields and temperatures, we reach a crossover to the normal state. Now there is an interesting question: what do you call superconductivity? Is it still superconductivity in this flux flow vortex liquid state with no long range order? I think most people would say yes because persistent vortex currents are still circulating around the liquid flux lines, but obviously this is fundamentally different from the pinned superconductivity of the vortex glass; so it's a whole new picture.

One of the other projects that we worked on during this time was how to determine what is the effective Hc2 where the vortices and the local circulating currents finally disappear, and we were able to do that through magnetic measurements of the diamagnetic breakaway from the relatively flat and small diamagnetism in the high temperature state. This was work in conjunction with John Clem. My talk here at this Applied Superconductivity Conference was in memory of John Clem, who passed away a year ago. He was an outstanding theoretical physicist who contributed a lot to this whole area.

This was certainly a very exciting period, to revolutionize our understanding of superconductors and to bring in all these new concepts like giant flux creep and the irreversibility line. We also recognized that these phenomena were related to the superconductor VI (voltage vs current) curves. Instead of having the classic sharp transition at a critical current, the high temperature superconductors show a curvature; so if you plot on a log-log scale, the slope of log V versus log I is the index value n, and it has a value typically, for YBCO materials, around 30; so the V-I transition is quite sharp and still occurs at a reasonably defined finite value. We realized that there was a very close coupling between flux creep and this transport property, this curvature of the VI characteristic. Interestingly, the normalized magnetic flux creep rate, d log M / d log t has a value for many YBCO materials of around 0.03. Another amazing thing was that for many of these YBCO materials, including thin films and even some bulk crystals, the transport index value n was around 30, and we were able to derive that there was a relationship between the normalized flux creep rate, which we called S, and this n-value: S equals 1/(n-1). Hey, 30, 0.03; 1 over 30 is around 0.03. This was an astonishing connection to recognize.

This kind of curvature in the VI plot was known, of course, in some earlier superconductors. In fact, David Larbalestier was famous for his work to try to understand this kind of behavior in low temperature superconductors, and the obvious explanation in that case is inhomogeneity. If you have any homogeneity in a superconductor you have a spread where the transition occurs and Larbalestier had nice calculations in his paper with Warnes about this. Now in the high temperature superconductors everyone might say, ah, aren’t these materials also inhomogeneous? But why would it be inhomogeneous in crystals and in thin films deposited in different ways and yet always have the same index value around 30? Within a certain range, maybe from 25 to a little over 30, but it was just impossible to explain this similarity on an inhomogeneity model.

To top this all off, the vortex glass theory predicted a large region as a function of temperature where the normalized relaxation rate was constant and controlled by a glassy exponent times the logarithm of the measurement time divided by a hopping attempt time as the flux line attempted to hop out of its pinning well. This prediction gave S-values close to 0.03! So the result I think now is that we understand how a lot of these novel effects fit together. This is not to say that Larbalestier’s inhomogeneity model may not be appropriate in some cases. For instance, we did some experiments at AMSC as you bend superconductor wires and cause inhomogeneities to form; lo and behold, the n-value starts to drop. In other words, both effects can contribute. But actually, in many good YBCO materials, the n-value in the VI measurements appears to be controlled by flux creep; so that's an important and novel insight. All of this work was really very exciting and we wrote a number of review articles. In fact, I've had the opportunity, just in the last month or two, to write chapters for a book that Chris Rey is editing on superconductors in the power grid. We're putting in a chapter to review this whole area, describing the “old” superconductivity and the “new” superconductivity and how they fit together; it's an exciting story.

Anyway, as time was going on I could see the handwriting on the wall at IBM: Yes, there was this great surge of enthusiasm at IBM about high temperature superconductivity. Over a hundred researchers were working on it —

Hochheiser:

Among the three locations.

Malozemoff:

Yes, at Yorktown Heights, at Almaden in California, and in Zurich, all three locations were working on high temperature superconductivity and beautiful results were coming out, such as the famous work of Dimos, Chaudhari, and Mannhart, which established an understanding of the grain boundaries in these materials. But I understood that there wasn't a near term applications driver for —

Hochheiser:

For IBM’s business.

Entrepreneurial Environment and the Commercialization of Superconductivity

Malozemoff:

Yes, for IBM Business, and that therefore this focus on superconductivity would eventually die away; there would probably be some researchers continuing to do some work, but it wasn't going to maintain the same kind of tremendous excitement. Meanwhile my own interest in applications had begun to develop. Now I was bitten by the bug of high temperature superconductivity and I wanted to see it succeed, to have some commercial impact. I did believe that HTS offered really new properties which could have a tremendous impact on the power grid, and through efficiency and many other features, it could have a beneficial impact on humanity. And so I began to think about making a career switch. Another factor in my thinking, as I already mentioned, was the work of Sumitomo Electric and seeing real HTS wires with quite high performance coming out of Japan. Wow!

And I remember an article from Bell Labs when they proposed the idea of a melting line in the H-T plane; some journalist wrote a piece on the basis of the Bell Labs work entitled “The Party Is Over.” The implication was that there was going to be no application at all. So here was this famous article, the party-is-over article, and at the same time here are these incredible HTS wires coming out of Japan; so what kind of party is over? I could understand that while there was a limit on HTS wire performance set by the irreversibility line, you could still get pretty darn good performance in a useful range of field and temperature. No, the party was not over! I must say, a little bit of nationalism also crept in to my feelings. Here we are in the United States, where YBCO was discovered and where we've got the two great industrial laboratories, Bell and IBM, leading the charge to understand high temperature superconductivity. But where's the American wire industry that's coming out of our leadership?

Now there was a little start-up company, American Superconductor (AMSC), starting to develop wires, but what have they done? I heard talks by them, but oh, come on; there wasn't anything to compare with the Japanese effort at that time. I thought the United States was just going to end up behind the eight ball and the Japanese were going to run away with the whole applications area; so I felt we have to make a try in the United States to have a credible wire and applications effort. At that time it turned out American Superconductor was looking for a VP of R&D, and I was contacted by a recruiter. I went to visit and spoke with Greg Yurek; he's a pretty amazing guy. I could see the drive and energy and his ability to sell, his ability to make deals. This wasn't just your average contract house where you get a little government contract here or there and struggle to survive, but he already had a very interesting deal going with Inco on developing the metallic precursor method for making wires.

Hochheiser:

Inco?

Malozemoff:

Yes. Inco Alloys.

Hochheiser:

I thought I recognized the acronym but I wanted to make sure.

Malozemoff:

Then another deal with Pirelli, a leading power cable company, to apply this technology to develop high temperature superconductor underground cables. So this looked quite interesting, even though I was a little skeptical about the status of AMSC’s technology at that time. I thought, well, hey, maybe I could bring some knowledge base in and really get this show on the road as the VP of R&D, and soon to become Chief Technical Officer.

Hochheiser:

Was IBM still in good financial shape when you left? I know a little bit later into the nineties they fell on hard times.

Malozemoff:

Yes. After I left just before the trouble started at IBM, some people asked me, how did you know?

[Laughter]

Hochheiser:

Did you know or was it coincidence?

Malozemoff:

No, it was coincidence. I just told you what drove me to make the switch and I did it while IBM Research was in its prime, and the Physical Sciences Department was in its prime. But within a year people were hunting for jobs like crazy and it was a kind of rout. Of course, IBM Research is still a great research laboratory. There are still many great people there, but it's not at the level that it was in those days, and in the interim we've also seen the sad disappearance of Bell Labs: it's been completely transformed and the way it used to exist is mostly gone; it's a great loss for the country.

Hochheiser:

Yes, that I know very well because before I joined IEEE, I spent most of my career as the historian for AT&T.

Malozemoff:

Okay, good.

Hochheiser:

I know the Bell Labs story very well.

Malozemoff:

At that time, seeing this happen as I was working on building up the research effort at AMSC, I saw a lot of little companies trying to pick up the pieces. I remember John Rowell, who left Bell Labs to work at another start-up, Conductus, and he pointed out that it was now the little guys that were really becoming the leaders in the field and maintaining the research drive, although now more oriented towards some application issues, while the big research labs were dropping out of it.

Hochheiser:

How large a company was AMSC when you went there?

Malozemoff:

It had 30 people.

Hochheiser:

You're going from one of the largest companies in the world to a very small one.

Malozemoff:

That's right. That was part of the attraction. In terms of my overall career it was exciting to enter an entrepreneurial environment. I always wanted to get into this kind of environment. I'm not a founder of AMSC, but I came in early enough that I was really part of the whole entrepreneurial surge of this company. I started in January 1991. Shortly after I joined, Greg Yurek and I went on the road and we had our first IPO, a very successful initial public offering. This was another great experience. If it had not worked out, maybe I wouldn't have liked it so much, but it did work out. I joined and helped contribute to making AMSC one of the world leaders in the high Tc field.

Hochheiser:

I find this transition interesting. I've spoken to any number of people who worked at major industrial research labs and their lack of desire to do what you did in the second half of your career was part of the reason they went with places like IBM and Bell Labs and the like.

Malozemoff:

The number of people who told me, you're crazy - there were a lot of them. But a change like this is something that gives you new life; suddenly you're embarked on a new enterprise and your neurons get recharged. It wasn’t as if I was bored at IBM. I had just had gone through one of the most exciting periods of technical creativity of my career, but it was clear that I had to make another transition, perhaps to something new, and how about trying something on the entrepreneurial side and also as VP and then as Chief Technical Officer, beginning to get into some of the commercial issues? One of my main responsibilities at AMSC, in addition to leading and setting the strategy for our technical developments, was the IP and patent portfolio, which is a very big issue for a start-up company. This was a significant part of my activity there.

Hochheiser:

When you joined was there a research and development group already?

Malozemoff:

Yes. There were a number of researchers there. I forget exactly what the structure was, but I think they were lacking a leader for the effort. Fortunately these were really good guys and they accepted my role without much grumbling, and we also started to hire some additional people who have continued to be key players at AMSC. One of them is still probably AMSC’s leading researcher, that's Marty Rupich, an outstanding chemist who has been a key part of the success of AMSC’s HTS wire, including its second generation metal-organic-deposition process. Steve Fleshler was another key hire, a physicist who has now become the overall technical manager of AMSC’s wire effort. Qi Li, Bart Riley and Alex Otto were dynamic contributors, although they have since left the company. Cees Thieme is an outstanding materials scientist continuing to lead AMSC’s development of a non-magnetic substrate for HTS wire.

Hochheiser:

What was the state of their technology?

Malozemoff:

When I joined, frankly, they were all over the place in trying to develop an HTS wire. One of the challenges in this field is that there are so many different materials to play with. Already by that time we had BSCCO and we had mercury- and thallium-based HTS and we had YBCO. These materials families came in different forms; for instance, we had YBCO 123 and 124 and 247, and there were patents on all these varieties. AMSC had licenses from MIT, and soon we also took licenses from Industrial Research Ltd. (IRL) in New Zealand. Then consider how many ways had been proposed to make a wire: there was the metallic precursor process, and there was the oxide powder-in-tube process, among many others, both with active proponents at AMSC.

I remember another interesting development only a month or two after my joining the company: AMSC had a technical advisory group, and one of the members of that advisory group was David Larbalestier, professor at that time at the University of Wisconsin in Madison. He's a brilliant scientist whom, if you haven't already interviewed him, you should. He had just come back from Japan and he heard a very interesting report there from a company called Fujikura about a new kind of wire process, the so-called coated conductor process with a polycrystalline but highly aligned substrate. Oh my goodness, how many different processes are we going to have to deal with? But I immediately noticed this Fujikura work and I thought, this is very interesting. At that time, Fujikura had a rather pathetic critical current, but their idea was amazing. To get good Jc out of YBCO —

Hochheiser:

Jc?

Malozemoff:

I'm sorry, that's the critical current density.

Hochheiser:

Okay, thank you.

Malozemoff:

Having a high critical current is essential to be able to make high performance cables or any of the other large-scale applications for the power grid. To Greg Yurek's credit, he understood that to have a commercially successful company, we had to go after big markets and significant opportunities, and he realized that the power grid, even though there are a lot of challenges to get new products into the power grid, that was the biggie. Already we were focused on this direction, but how to get there? What kind of technology to use? I was very interested in this coated conductor technology. You see, with YBCO, there was no good way to align it directly and avoid the grain boundaries which limit current flow, except by growing a single crystal, which is obviously not a practical way to make a kilometer-length flexible wire. What Fujikura discovered was that through their clever IBAD (ion-beam-assisted deposition) process they could texture a substrate and then grow the superconductor epitaxially on it, thus aligning the grains of the superconductor and minimizing the effect of the grain boundaries. I remember at that time that I set the goal that if someone could demonstrate a level of one megamp per square centimeter at 77 Kelvin, AMSC should take off and pursue this wire technology approach. But in those early years, the coated conductor technology wasn’t there yet, and we already knew about the amazing work of Sumitomo Electric using the powder-in-tube technology with BSCCO-2223.

So my first job at AMSC was to select what our real focus was going to be, and I had to tamp down and turn off a variety of creative, but not very long-term-promising, efforts and try to refocus our work on one or at most two paths. I thought the Sumitomo Electric process with BSCCO-2223 was at that time the most promising one, and that became our so-called first generation wire. But there was another process, the metallic precursor process, which was based on the early work of Greg Yurek and John Vander Sande at MIT; so needless to say, there was a strong vote for pursuing this process at AMSC also. So our little company pursued both processes for many years. We had the collaboration with Inco Alloys on the metallic precursor process, which had many attractive features although it was never able to equal the performance of the oxide-powder-in-tube process that Sumitomo Electric had developed. We also made some very exciting hires and within a couple of years we had zoomed past Sumitomo Electric with our oxide-powder-in-tube process with BSCCO-2223, and soon we were making the best first generation wire in the world.

Hochheiser:

Did you have a license from Sumitomo Electric?

Malozemoff:

Well, there's a whole story on that.

Hochheiser:

I thought there would be.

Malozemoff:

Let me just say that on the technical side of things we were very successful in developing that wire. One of my other responsibilities was developing and managing our external technical collaborations. One of these was a consortium I pulled together of various researchers at national labs and universities to complement the capabilities of our own research team. We called this consortium the Wire Development Group. The work at the national labs and universities was supported by DOE. The collaboration had great value to all parties and continued successfully over almost two decades, providing many insights into the materials science of our wires. I was supported in this effort by a key participant, Prof. David Larbalestier of the University of Wisconsin, Madison, who later moved to Florida State University and the National High Magnetic Field Lab. Another long-lasting collaboration was with the Industrial Research Ltd. (IRL) of New Zealand, which started because of our license of their patent on the BSCCO-2223 material but morphed into an effective technical program. And we also had seemingly countless technical programs with our industrial and utility collaborators – Inco, Pirelli, Hoechst, ABB, EDF, Reliance Electric, Nexans, Southwire/Ultera, Siemens, LS Cable…

Meanwhile, Greg Yurek was always pushing for manufacturing scale up as early as possible so that we would really move down the experience curve; and so we started investing in a manufacturing line. And of course bringing in the funds from the IPO gave us the financial means we needed. We also had money from the Pirelli deal, and from Inco; so we started building up a manufacturing line and started producing reasonable lengths of conductor. All of this was enabling the first tests of power cable at Pirelli. AMSC’s first generation wire effort was a great effort; we led the world in this field. I'm sure Sumitomo Electric would say we were equal; we were both up there in a neck-in-neck competition. And Sumitomo Electric was also developing cables in Japan. It was really very exciting to be in this international race.

Meanwhile, Fujikura’s coated conductor concept made the long-awaited breakthrough thanks to the far-sighted and daring commitment of Dean Peterson, leader of the HTS effort at Los Alamos National Laboratory. I remember that he fought with his superiors to fund work on the coated conductor concept, and lo and behold, those guys got to 1 megamp per square centimeter! Then I said okay, let’s go - and at AMSC we took off and started work on this second generation wire.

Development of Second Generation Wire

Hochheiser:

Roughly when are we talking about?

Malozemoff:

This was in the late 1990s.

Hochheiser:

By the late 1990s you're now working on the second generation.

Malozemoff:

We started to work on second generation wire. There was a long way to go, it turned out. But meanwhile we had the world's largest production of first generation wire, the best wire in the world, and so we could support applications efforts around the world.

Hochheiser:

Whom were you selling the wire to? What was it being used for?

Malozemoff:

Pirelli was one major customer. They were building trial power transmission cables, and eventually a number of DOE-supported projects got started. We had various demonstration projects with them, culminating in the infamous Detroit Edison Project, which I will come back to in just a moment. There were a lot of other customers: There was HTS motor activity at Reliance Electric; we had a cooperative effort with them developing high temperature superconductor motors, which was very successful.

Hochheiser:

Yes.

Malozemoff:

There was quite a bit of other activity as well. By that time the metallic precursor program finally died when Inco gave up. In a way, that was fortunate because we just couldn't maintain three different programs in parallel. So now we still had two: We had the first generation BSCCO oxide-powder-in-tube wire with an active manufacturing line; and we had the development program on second generation YBCO coated conductor wire. With time, our focus shifted more and more onto second generation wire. Those years were very critical in the company.

I can come back to that, but let me first answer your earlier question about the IP and the license from Sumitomo Electric for the first generation wire. I heard that during the first year of high temperature superconductivity, Sumitomo Electric filed a thousand patents and their patent attorney in this field had a heart attack. This is hearsay, of course, but it’s a good, if sad, story. They were so aggressive in their filings that we realized, oh my goodness, we really have to do something to build our own countervailing position. We started filing very aggressively also, hired a lawyer, and got some really good patents, although the patents were mostly in the United States. Then we had to face the challenge from Sumitomo Electric about getting a license. Their proposed license terms were quite aggressive, enough to more or less finish off our company. We had a quite an adventure in negotiating this license. It was a very exciting experience for me because I led the negotiations, supported by an excellent patent attorney Jim Lampert from the firm Hale and Dorr. Fortunately we had our patents in the United States. The irony was that Sumitomo Electric wasn't selling HTS wire in the United States while we were trying to sell internationally including into Japan; so we were subject to their patents and they weren't subject to our patents at that time.

Hochheiser:

Alright, we're ready to break now.

[End of tape 1; begin tape 2]

Hochheiser:

You were talking about patent positions.

Malozemoff:

Yes. Concerning Sumitomo Electric, we had a long and exciting negotiation that finally converged on an agreement which was more or less reasonable because ultimately Sumitomo Electric did want to penetrate the U.S. market and they needed some of our patents as well. This was a very interesting experience to learn about how to conduct these negotiations and to get to reasonable terms and to fight for what we believed in. In fact licensing became a significant element of my work at AMSC. We had many negotiations. We also had a long and tough negotiation with Oak Ridge National Laboratory because we decided to adopt, for second generation wire, their so-called RABiTSTM (rolling assisted biaxially textured substrate) patent. I must say, in some ways the negotiations with academic institutions or national labs are more challenging than with competitors. You might think competitors would be tougher, but if you have a patent you can hold over them you can get a reasonable deal.

Hochheiser:

Correct.

Malozemoff:

But no patent that you have makes any difference to a national lab or a university, and so we had a terrible time to get terms from them that would allow the company to survive. There was somehow the feeling that this field is so phenomenal, with such vast and lucrative markets, that they could ask us for anything they wanted; millions and millions and royalties through the roof. But the wire business is not like software, where your profit margin can be 90% or whatever. In the wire business, we're going to need to fight for every dime. After all, if you give them a royalty of 5% or 10% and you have to get 10 different patents, before you know it there will be no money left. So we fought long and hard in a lot of these negotiations. At times, we had to knuckle under and agree to some pretty terrible terms, and later we went back and somehow had more reasonable people to negotiate with. We pointed out that if we continued with the onerous terms, they would not have any return because there would be no company left to give them anything. So we renegotiated a couple of these license deals and finally got packages which, while still financially painful, are at least allowing the company to go forward in the field.

Hochheiser:

In terms of the size of the company, there were 30 people when you joined at the beginning of the nineties and how large was the company in terms of any number of employees, sales, whatever, by the end of the decade.

Malozemoff:

I guess I could've reviewed that number before coming to this interview, but I'm sure we were in the range of a hundred people, maybe more. There's the whole story of how we got into some other business areas that I'll come to in just a minute, because that’s what led to a dramatic growth spurt in the company. In terms of the superconductor portion we were probably in the range of a hundred. It was a big effort and we had had a very successful further public offering. Through our history, there were five public offerings and that was the magic of Greg Yurek; he was able to pull this off time and time again. He was just brilliant in these financing efforts and kept us with strong funding to keep us going in our efforts. As we got into second generation HTS wire, this was very tough sledding in the first years because —

Hochheiser:

Are we talking about now the late nineties now?

Malozemoff:

Right around Y2K. At that time, in our second generation coated conductor program, we could occasionally get terrific performance but more often than not, we couldn't, and when we tried to make longer lengths, the results were somehow just so un-reproducible. I remember we hired a fellow who had been a postdoc at Oak Ridge National Lab, a Chinese guy who was so hard working; he produced so many samples, but it was just disheartening: bad result, good result, bad, bad, good, bad, good, bad, bad; it just went on like this. We were at the point of just killing the whole effort. But it's amazing how these things happen: During this time we were negotiating the license deal on RABiTS with Oak Ridge; this was a very tough negotiation and it went on and on. It was very complicated because at that time Pirelli was still a partner and AMSC’s agreement with Pirelli required us to have terms which would allow Pirelli to have access to everything that we got from Oak Ridge. But the national labs had rules that licensed products had to be manufactured by U. S. companies in America, and Pirelli said no, we want to be able to manufacture wherever we want. So, the negotiation just went on and on and on for two years.

The irony was that the key technical clue to what was holding us up in our second generation wire process and making it so unreproducible was back there at Oak Ridge and they wouldn't tell us about it until we signed the agreement. So we lost two years in developing second generation wire. When we finally got the agreement, we started talking to some of their researchers, and – well, I don't think I should say what this particular phenomenon was, but it was an amazing trick in the epitaxy of the superconductor stack, and with this one trick, all of a sudden our struggling Chinese researcher started getting consistent results. I remember that when our team finally got a meter length with reasonably uniform current density, I sent around a note to all the management team about “breakthrough results” and Greg looked at it and said “what kind of breakthrough? Come on, give me a break.” After such a long period of failure, he was naturally very skeptical and was ready to kill the program. But soon he had in his office a copy of these data framed, and all the members of AMSC’s Board of Directors signed it. And so AMSC’s second generation wire program was saved, and that's the way things go in these kinds of entrepreneurial companies. I mean it's hair-raising at times, but of course so exciting. And then, once we had the breakthrough, the technology progressed very steadily.

Committing to this coated conductor technology was another very difficult decision from a strategic point of view because there are so many types of coated conductor processes out there to choose from. We chose the one, called MOD/RABiTS, that we thought would be the lowest cost, and we still think it's a low-cost process, but that's not to say that it's the only one, and in some other companies, the other processes, I have to say, have done better than I expected them to, including the original IBAD process which Fujikura had used. So there are a lot of viable competitors out there, but American Superconductor, I think, is known as one of the leaders in the field with probably the largest capacity production worldwide. So we made the painful decision to close down our first generation wire factory and decided to march forward with the second generation coated conductor technology, developing a new manufacturing line while continuing R&D to improve the wire performance.

Another important part of the story of American Superconductor has to do with the other “legs” that American Superconductor developed. How these things happen is absolutely amazing, a great story for the history books. It all started with one of the applications we always thought was very interesting for high temperature superconductivity: this was SMES, superconducting magnetic energy storage. The idea was that since superconductor currents can flow forever without losses, one can store a lot of energy in the magnetic field created when the superconductor wires are wound into a coil. Now there is a significant and growing power quality problem in the electric power grid. Short power interruptions caused by short circuits can knock out microprocessors and bring down a semiconductor line and other microprocessor-controlled manufacturing lines. You'd like to be able to control these so-called dips and sags in voltage, and so someone had this great idea that SMES might be a way to do it because with superconducting coils you can eject electric power very rapidly; so it's a good way potentially to inject real power into the grid and somehow compensate for these dips and sags. That seemed so exciting that AMSC purchased a company that was developing this technology, a company initially called Superconductivity Inc. It was located in Madison, Wisconsin.

Hochheiser:

About when was that?

Malozemoff:

The acquisition was made already in 1997. But then we realized that while this application was technically interesting, a major problem was its cost. Superconductivity Inc. had developed a beautiful low temperature superconductor magnet based on niobium-titanium (not the high temperature superconductors), and they had an excellent cryostat and refrigeration; everything was beautifully worked out and they had working power electronics to convert the DC current from the superconducting coil into an AC current of the right phase and frequency to match the power grid. This power electronics was purchased from ABB, unfortunately at great expense. In fact, the cost of the power electronics portion of the SMES system far out-shadowed the superconductor portion; so we realized, if this is ever going to be a commercially successful project we've got to lower the power electronics cost.

There was a little company in Milwaukee that was developing advanced power electronics. It was led by a very creative guy called Jeff Reichard, and AMSC purchased his company. With him, we developed a very attractive power electronics package for the SMES system. We were purchasing the power transistors from various places and assembling the power electronic system in what became AMSC’s power electronics division in Wisconsin. Now we had very nice and commercially competitive power electronics, we had a beautiful superconductor magnet, and so we'd go to the utilities and say look what we've got; this is perfect for stabilizing your grid, or maybe we might go to a semiconductor fab and say, look at this; you can protect your fab from all these voltage dips and sags.

Some of the utilities would come back to us and say, hey, you've got a very interesting product there; we'll buy the power electronics but just leave off the magnet and give us a nice reduction in price. We were stunned. How could this be? How could they want a SMES system without the superconducting magnet? Come on; give me a break! This was incredible! But it turned out that they were onto something, because when there is a short circuit somewhere in the grid creating a voltage dip, what's happening is that you're shorting out the resistive load; so what's left is mostly inductance or capacitance, because most of the grid components consist of inductances or capacitances. It means that you've changed the power factor and that you can compensate for that dip or sag with reactive power, which means out-of-phase power, so-called reactive power, either inductive or capacitive power. In short, you don't need real power to compensate for the dips and sags in voltage and address the power quality problems in the power grid. Reactive power can be manufactured out of thin air, as it were, by power electronics and injected into the grid. It's as if the power electronics looks like an inductor or it looks like a capacitor and compensates for the out-of-phase behavior. Lo and behold, the power electronics gave an elegant solution to the power quality problem. No superconductor coil needed!

Superconductor and Power Electronics Applications

Malozemoff:

It was quite a painful decision to throw overboard our wonderful group of scientists and technologists who had developed this superconductor magnet and all the technology that went into it. But we had the perfect power electronics package to address the market need. So AMSC renamed this system DVAR, distributed volt amp reactive, and optimized it for power quality applications. This continues to be a successful product today. The general concept of using power electronics to control and upgrade the power grid was initially developed at the Electric Power Research Institute (EPRI). They called it FACTS or flexible AC transmission systems. DVAR is one of the most successful FACTS devices on the market today. AMSC has installed systems all around the world, especially in areas where there are specific requirements for grid quality, such as in many of the English-speaking countries; so we're in United States, Canada, U.K, Australia, New Zealand, Singapore. This business area is expanding; so it's a very exciting area and it led to a significant growth of American Superconductor. This is quite a story, how an idea coming initially from superconductivity branched off into a major new non-superconductor business for AMSC.

Hochheiser:

I guess that’s part of being entrepreneurial.

Malozemoff:

That's right.

Hochheiser:

Recognizing something like this.

Malozemoff:

Exactly. Once we had this optimized power electronics package, we also offered it for sale to people who might want it for other applications. Now there was a small company in Austria, Windtec, an engineering company in the wind field, that started buying a couple of AMSC’s power electronic packages, and soon after then they bought ten, then they bought a hundred. By this point, Greg became very interested: what's going here? He visited them, learned about their interesting application in wind turbines, and he decided to buy the company! They were providing an engineering package, particularly in developing world environments where companies were trying to get into the wind business. Windtec sold a package consisting of a license, a package of technology covering the entire design of a wind turbine, in some cases a joint development program, and finally AMSC’s power electronics, which acted like the brains of the wind turbine. This business grew like wildfire. AMSC was able to sign contracts with a whole bunch of companies around the world, including China, Korea, and India. All of a sudden American Superconductor was growing like crazy and reached 800 employees. Revenues grew to $300 million annually. The superconductor portion, in terms of revenue, had become a mere 10% of the whole business; the rest was power electronics and wind.

Hochheiser:

Right

Malozemoff:

The role of superconductivity in AMSC became a question, but there was still a conviction that superconductors were important, that they had a potentially significant role to play, and the commitment was there to continue the development in HTS wire and its applications. From the early years Greg Yurek had set that vision. We always wanted to do more than just wire. We wanted to have the value-add of applications. And so we were looking at a variety of superconductor applications. One of those was rotating machinery – motors and generators. We had a very active program, initially with Reliance Electric, and this developed in various ways. Then with the Navy we had a major contract to develop what remains the most significant rotating machinery project with high temperature superconductors to date: this was the 35 megawatt Navy motor, designed to power the next generation of electrically powered destroyers. It was a very successful program; the motor was built and successfully tested. But unfortunately, budget cuts being what they are today, this program has stalled and has not yet gone forward to on-board testing of the motor at sea. We hope it will someday, but right now the HTS motor program is not on the first burner, as it were. That's a bit of a frustration,

Another area we looked at very seriously was fault current limiters. For me personally this was a great opportunity also to broaden my scope of interest to learn about these applications, and we were actively working on them. I wasn't a hands-on guy in AMSC, but I was certainly in there in terms of a lot of the contracts and many of the technical concepts. I did a certain amount of work on, for instance, AC losses in power cables. There were a lot of questions about how losses arise in AC cables, also in fault current limiters. We started a project with Siemens and produced the first high voltage fault current limiter, learning a lot in the process. Early on, we also had a program with ABB on transformers. That program was not successful, it has to be said, because of the challenges of getting a sufficiently low-AC-loss wire. At that time, in a program launched in 1996, we were trying to do this with a first generation wire that proved to be just too difficult and too costly; so the transformer work went by the wayside.

If we look back at the history of the various collaborations that AMSC had, it’s an amazing series of collaborations. We collaborated with many of the leading corporations around the world in the electric power field, and a few of those have continued, but a lot have been discontinued. Pirelli was one of the most important collaborations for the first decade that I was at the company, and it was all culminating in the project to produce a cable demonstration at Detroit Edison. Unfortunately, the cable failed. It had nothing to do with the superconductor. The superconductor was fine, but the cryostat leaked; you couldn't pump it down.

Hochheiser:

Yes, you can’t keep it cold.

Malozemoff:

We couldn't keep it cold. The team tried everything and they just couldn’t do it; apparently the welding was not good. So that blow, I think, convinced Pirelli to drop out of HTS development. That was a critical moment for the whole field because Pirelli was the leading cable player in the world. Sumitomo Electric in Japan was another major player, and they kept churning along, but when Pirelli dropped out, it looked like the bottom was going to fall out of the whole high temperature superconductivity applications field. But we worked very hard and were successful in inducing another very significant company to come in and pick up the flag. This was Nexans, based in France, though a lot of our collaboration was with their division in Germany. They have been wonderful partners with us in a series of projects. The biggest one so far that everyone knows about is the LIPA (Long Island Power Authority) project, the first transmission-level 600-meter long, 3-phase cable, a very successful project. It's located in Holbrook, Long Island.

Hochheiser:

This is a current project?

Malozemoff:

Yes. It was energized in 2008 and ran for five years very successfully. The initial version of the cable was made with the first generation HTS wire. Now we are working to upgrade it with one phase of second generation wire.

This whole effort to develop a variety of superconductor applications, including cables, motors, fault current limiters, transformers, and so forth, was a tremendous effort. We had a very powerful group on a lot of these applications but they didn't all work out right away. AMSC still has the potential to go forward on many of these applications. I was particularly disappointed that we somehow never got a major contract to apply AMSC’s rotating machinery technology to 10-megawatt wind turbines for off-shore wind applications. Everyone knows that you can lower the cost of offshore wind if you can get to a very high power, 10 MW, in a single turbine. You build one tower and get a lot of power out it, instead of many towers with smaller output. But the problem with conventional technology is that 10 MW generators get so huge that their weight on these high towers makes the tower mechanically unstable. One of the benefits of superconductor equipment is compact size arising from the high power density of the HTS wire. That may be one of the main drivers for superconductor equipment in the grid - that it can be made more compact, more energy dense — -

Hochheiser:

And presumably lighter as well.

Malozemoff:

Lighter, exactly.

Hochheiser:

If you're dealing with a wind turbine, I imagine one of the problems is an issue of weight on these tall structures.

Malozemoff:

Exactly. That's it. And that's why it's not easy to go to a 10-megawatt system with conventional technology. AMSC’s HTS Navy motor came in at 75 tons, still a lot, but a conventional motor with that capability would have weighed 300 tons. It was a factor of four improvement in weight and size, and that's including all the cryogenics, everything. With that kind of advantage, you can now do a 10-megawatt wind turbine and this is widely recognized. At American Superconductor we have the HTS technology to do it. We're the first ones who built a machine of this scale, but demonstrating such a wind turbine is not something that we have enough funds to finance ourselves, and so it hasn't happened yet. Meanwhile there are a number of efforts internationally to take advantage of this concept; so I don't know what the future will bring. I would love to have seen American Superconductor lead the world in opening up this tremendous resource of clean energy because the best wind for wind power is offshore. On land the wind is much more variable, and half the time the land-based wind turbines aren't working, but offshore you could have much steadier wind, and much more powerful.

So that's the story of our move towards commercial applications. It culminates in the announcement which was made just two weeks ago of the new project to install an inherently fault-current-limiting cable in Chicago in the downtown Loop, to introduce what we call a more resilient electric grid, a grid that is capable of surviving potentially significant damage to the inflow of power from the transmission grid to a given substation. Now the general configuration in many cities is that transmission lines come into a substation, the voltage is converted by transformers down to the distribution voltage and then power is fed to a given sector of the city. In case of problems in a given substation, Con Edison always keeps two extra transformers at each substation in New York City; this is an N plus two redundancy scheme. They want to be able to survive potentially a loss of two transformers; so they have two extra transformers on hand. These power transformers are humongous, as you probably know; so that's a lot of space to allocate to transformers that are just standing around in New York. Of course there is always the danger that you lose power from the transmission lines coming in to a given substation, and then the whole area fed by that substation goes black. What if that area were the financial center of New York City for example?

That concern was what prompted the Department of Homeland Security to fund the Hydra program, which is another one of the important programs that American Superconductor has been involved in, to develop an inherently fault-current-limiting cable. That program is still underway. Certain shorter length cables have been now demonstrated successfully and the full length cable has been actually manufactured and is ready to install, awaiting some civil works from ConEd to bring it in. This application is a very powerful one. The idea is to link the distribution level busses of neighboring substations; so if the power for some reason gets cut coming into one substation, you can transfer the power from another substation. You can also avoid storing so many transformers; you have extra grid resiliency as a result. Having a superconducting cable which would connect these different substations also solves a lot of the problems of installing new grid infrastructure in a city like New York or Chicago because these power cables, also benefiting by the higher power density, can carry much more power in a very small size, and if you have ducts you can simply pull your superconductor cable through. Then you have three or four times as much power capacity as you had before; you haven't had to dig up city streets, which is one of the main expenses of this kind of an upgrade project; so all kinds of advantages emerge from this application.

But there is a critical requirement for such a link between substations, which comes from fact that the link increases the fault current. If there is a short circuit in one of the feeders, for example, power can flow from both substations into this short circuit and you can get a huge fault current. In many grids in major cities around the United States and around the world, fault currents are already near their limits. All power equipment is rated for handling fault currents because circuit breakers don't open that fast. At the transmission voltage level, it’s a couple of cycles, and longer in the distribution grid. The grid has to survive those kinds of events; so all equipment is rated for its fault-withstand capability. But if you have this kind of extra high fault current arising from linked substations, you can exceed the equipment’s fault rating and start burning it up. It's a catastrophe. So there has to be a fault-current-limiting feature to any such link between substations.

We realized that with our understanding of the fault current limiters generally, one could do this with the superconducting wire that’s already in the cable. That’s what it means to have an inherently fault-current-limiting cable. I was active in developing this concept with our team. We have a wonderful patent on this approach, and now, beyond the Hydra project, we have this new ComEd project in Chicago. This will be the largest high temperature superconductor project anywhere in the world to date. It's a tremendous next step for the ongoing commercialization of the technology, although this project is still not fully commercial in the sense that some support, especially for the earlier phases of the project, still comes from the Department of Homeland Security. It's very exciting to see this whole field growing step by step in terms of its ability to solve critical problems of the power grid and using high temperature superconductor wire. This has all been a wonderful adventure during my career at AMSC to see this kind of development.

AMSC's Business Relations

Malozemoff: American Superconductor had grown, by 2011, to 800 people with divisions around the world. We had our Windtec division in Austria, field offices in many countries such as Korea and Australia, and a power electronics assembly plant in China. Then something very unfortunate happened. Perhaps you know about this.

Hochheiser:

I read about it while doing my research for this interview.

Malozemoff:

The company had grown so rapidly; suddenly its revenues surged from maybe 50 million a year to 300 million a year. We were a profitable company, we really had it nailed. But seventy-five percent of that revenue was from a single customer, one of the most aggressive customers getting into the wind turbine field; this was Sinovel in China. We developed a factory in Suzhou in China to deliver all of the power electronics for Sinovel, and all was going very successfully. Sinovel was establishing wind farms all over China, and they went from nowhere to being among the top ten wind turbine companies of the world. They had excellent support from American Superconductor's field service team, and then all of a sudden they just cut us off. They informed us that all contracts were terminated. They rejected all our work-in-progress and final product ready to be delivered. Our contracts with them were violated right and left, and our IP, as we've learned, had been stolen. It turned out that one of our employees had been bribed in China to deliver our IP to Sinovel. This has become a famous case, and it was, as I recall, a feature story in Business Week and other business press.

Hochheiser:

The one I came across was in Forbes.

Malozemoff:

Forbes, and it's all over.

Hochheiser:

It probably was in every such publication.

Malozemoff:

Yes. Now there are legal proceedings, unfortunately in China, because the contracts were written in China. They are at the Beijing Supreme Court; AMSC seeks total damages of $2 billion; so this is a big, big deal. It's a high priority item for our Secretaries of State; it was on Hillary's list, and it's on John Kerry's list. And this development suddenly cut the revenues of American Superconductor by a huge fraction. The company had to fire half its people and contract drastically. The other business areas were still there, and fortunately the company - that was one of Greg's principles - had no debt. We had a good amount of money in the bank, and so the company has been able to survive. I had retired two years before; so I was not on the front lines during this crisis. But I continue to consult at AMSC; so obviously I follow developments as closely as I can.

There is skepticism still in the analyst community on Wall Street about AMSC’s future. They keep charting the quarterly loss and looking at the amount of money we have in the bank and predicting how many more months we have left. I sometimes wonder if they're surprised that quarter after quarter goes by and AMSC is still there. They don’t seem to factor in that revenues can actually grow or that the company can take actions to reduce expenses. Yes, the company's still here and announces a project like the Chicago Commonwealth Edison superconductor cable, with tremendous prospects for growth in the superconductor segment of the company’s business. Meanwhile, the DVARs are continuing to sell, and power electronics is continuing to go to wind turbine customers. After Greg Yurek retired, Dan McGahn took over as CEO; he is another capable leader, guiding the company out of this crisis and towards a very promising future. So all in all, I'm very excited for the future of the company.

Post-Retirement Work

Malozemoff:

I have to say that from a personal point of view, I guess I had always felt that it's very intense to be an officer in one of these kinds of entrepreneurial companies and I thought maybe I would not continue at this level for the rest of my life, but that at some point I'd retire; I have many other interests and would like to pursue some of those. So I decided to retire at 65. Since then, I have kept my finger in AMSC to some degree by continuing as a consultant.

Hochheiser:

Right and that was in 2009 right?

Malozemoff:

Yes. I decided to retire then. There were many exciting things that I could do; this was before the Sinovel disaster; everything looked wonderful for the company. I developed a contract to be a consultant for AMSC, running a technical advisory board and consulting on technology and IP issues. This was fun to do at a low level. I also now had time to get involved in some other activities, personal interests, as well as some other professional activities. This was a very, you might say, serendipitous move because as an officer you are legally liable for the company’s behavior, and when this Sinovel crisis broke, I was spared all the trouble of dealing with it. Of course in addition to the problems with Sinovel, a class action suit was filed against the company. In short, it's been a couple of tough years for the company, but it's coming through, and I believe it will be very successful. I continue to believe that AMSC’s vision for its various clean energy technologies, including not just the superconductors, but also power electronics and the wind energy, will prove very successful going forward and can lead AMSC to become a major corporation.

Hochheiser:

What other professional activities have you gotten engaged in since 2009?

Malozemoff:

Very soon after I retired I had a very interesting opportunity, which became almost a year-long and full-time enterprise for me, in addition to the work I was doing at AMSC. It started with my being asked to come to a meeting of the Basic Energy Sciences Advisory Committee. Basic Energy Sciences (BES) is one of the divisions of DOE, a very significant division that is one of the largest funders of fundamentally oriented work in the United States. It also funds a lot of the major analytical laboratories and facilities around the United States.

BES had a tradition of sticking to very fundamental work. Their view is that there's an applied section of DOE which does applied work, and they don't want to get sucked in to doing applied work as well; their mission was to do fundamental scientific work. At that time, however, Bill Brinkman, head of the Office of Science, asked whether BES, with its particular skills, could do something useful for energy; after all BES is part of the Department of Energy. So they brought in me and a couple of other people to give talks to BESAC, their Advisory Committee, about whether there was something useful that BES could do to help clean energy technology. This question is very much down my line. I've done fundamental work during my career, even during my time at AMSC with that Wire Development Group for example. We made a specialty of figuring out the kind of fundamental science that needed to be done and that would build a material science foundation for the HTS wire industry. So I gave a talk to BESAC on this. During my years at AMSC, we had often approached Basic Energy Sciences and they refused to support the Wire Development Group. Jim Daley in the applied area of DOE supported it, but always grumbled, saying something like: “this is not really my responsibility because you guys are doing more fundamental material science.” So I would go to BES and I'd say, “So why can't we get support from you?” “Oh, it's too applied.” So I groused about this story in my BESAC talk and thought that's the last I would ever hear from these people. I don't know how it happened, but to my surprise BESAC asked me later to co-chair a subcommittee with George Crabtree which would address the scientific opportunities for support of clean energy technology, opportunities which BES could potentially contribute to.

Developing the report of this subcommittee was a big project, which obviously was far beyond the scope of superconductivity, expanding now to focus on all clean energy technologies like solar, and batteries and carbon capture and so forth. We pulled together a very exciting workshop, at which we made a special point of bringing in leading scientists in industry who could really contribute to these kinds of ideas. Our mission was not only to find particular scientific topics of interest but to identify issues preventing effective interaction between industrial scientists and basic researchers. This is not to say that such interaction hasn't happened before – it certainly has - , but there are so many areas where such interaction is needed and isn't happening. Often national labs say, we're doing all this work related to energy, and then you ask the industry people and they say, what a waste of money at the national labs; why are they doing what they are doing? Often one finds industrial and academic scientists just talking past each other like this.

Anyway, at the BESAC workshop, we got the academic and national lab and industry groups into one room together, field by field. We addressed ten different clean energy technology areas. I think a lot of attendees were very excited by this workshop because this kind of interaction isn't done that often. Academic researchers were saying, wow, there were some really interesting and relevant topics to explore, and the industrial people said, oh, these academic labs really have good capabilities we need. So it seemed like things were really moving forward. We wrote a major report, published in August 2010 and available on the DOE website. We called it the SciTech Report, for short, which means Science for Energy Technology.

In addition to the particular fundamental technical topics of interest, we went into some of the ways of encouraging effective interaction between the fundamentally oriented academic or basic research community and the industrial community. For example, I wrote into the report the story of the Wire Development Group or WDG, which was, I think, a tremendous success over many years, linking several national labs and universities with AMSC. The WDG really laid out the materials science foundation for our HTS wire technology, discovered amazing things that we would never have done on our own. It was beautiful scientific work, yet at the same time directly relevant to practical HTS wire development. We always made sure that about 80 to 90% of this work could be published because if the industrial partner announces that no one's going to hear a word of this, that all results will be held confidential, then the academic community will totally lose interest in collaborating. They don't want to be industry’s slave; they want to have publications and an external reputation; so we found an excellent compromise on how to do this. There were indeed so many things that we needed to know and that didn't have to be company top secret.

All in all, the BESAC SciTech report was a really exciting project. My collaborator, George Crabtree, is a wonderful scientist and manager in the national lab environment; we worked very well together and produced the SciTech report, with wise guidance from Professor John Hemminger from UC Irvine, chair of BESAC. So the report is out there and I'm very proud of it. I'm just disappointed that the reaction from BES management was: Okay, we'll ask for more money for our budget because of course no BES-supported researchers wanted to give up the programs that they were already working on. They thought addressing issues in the SciTech report was a path to yet more funding, and needless to say, in the present tight budget environment, very little additional funding was made available.

It's, in a way, very frustrating because I think there's such an opportunity for major progress through close academic-industry collaboration that we saw in field after field. The opportunities are often with small companies like American Superconductor, that don’t have the vast analytical resources of the national labs and some universities. For example, there are many small solar companies with some very nice technology, but they don't have the analytical capabilities to track down the defects that are limiting their performance. The national labs are full of electron microscopes and all sorts of other powerful analytical techniques. A small solar cell may give you very high efficiency, but when you make a full-size solar panel, you find you may lose a factor of two in efficiency. It is critical to find out where and how that factor of two was lost. You need some good analytical work to find out what the problem is, but this kind of critically important work is not going to win some scientist at a national lab a Nobel Prize. However, potentially it could revolutionize solar power and make it truly economically competitive! After all, solar still not competitive today vis-a-vis carbon-based energy. Solar energy is expanding, but it's not expanding at the rate it could be expanding. If only you could just get these academic and industrial groups together and somehow convince the academic people that the work is really important and that you might find out some interesting science in the process by putting their tools and capabilities to work!

Meanwhile, I've been continuing my consulting work with AMSC; I run the Technical Advisory Board; I've been doing some analysis of IP; I’ve completed a paper with John Clem on ac loss in cables. Then, just recently, in the last couple of months I've been working on another fun project - a book on high temperature superconductors for the power grid edited by Chris Rey. He initially asked me to write an introduction to the whole book. I did that about a year ago and then I didn't hear anything for a year; so I finally I contacted him and asked, what's happening? Is the book still going to happen? He said he still hoped it would, but there was a problem: a lot of the authors whom he had recruited to write different chapters couldn't complete their chapters.

Hochheiser:

I've heard that story before.

Malozemoff:

I'm sure almost all books are like this, right? He was in a panic because some of the chapters that were missing were the most important chapters for a book like this, such as a chapter on HTS AC cables. Cables are the most promising application for high temperature superconductors in the power grid right now. How could you have a book like this without a chapter on high temperature superconductor cables? He asked this person and that person, but no one could do it, and so finally I said to Chris, I can get a lot of words down on paper really fast and I know something about the field. So I can help write the cable chapter. There was also a guy out of AMSC who might be able to help me on the areas that I'm weak in, Jie Yuan, a very competent technical contributor. Then Chris Rey could probably contribute a little bit himself from his knowledge base. Let's do it, let's just sit down and write it. Since I was retired, I had enough flexibility in my schedule. I pushed other things out of the way and sat down and worked for two months straight, and we produced a chapter of 60 pages plus figures, almost a book in itself.

The process was so exciting because I realized that what was in the literature was really not laying out some of the key technical ideas. We had had a guy at AMSC, another wonderful technologist, Swarn Kalsi, who left the company but wrote a book on high temperature superconductors for the power grid, and he covered a lot of relevant material, but we found we could complement his book in important ways. For example, I had been working on the whole area of AC losses in HTS cables and on what determines the AC losses in these very complicated helical structures. So I introduced that story and also the whole story of inherently fault-current-limiting cable which has recently become one of the most promising HTS power grid applications. Now, in addition to the Hydra project in New York City, we have a major project to introduce fault current limiting cable into the Chicago downtown Loop. Many people don't understand the technology for fault-current-limiting cable; they think the cable is going to burn up at any defect. So I introduced a whole section describing the underlying theory of fault-current-limiting cables. The relevant information is actually already out there in the public domain; there are a number of very beautiful papers from the Siemens' group in particular which explain the principles of fault current limiting using the superconductor-to-normal transition. But now I showed how you can understand how this phenomenon would work in a cable. This perspective just isn't out there yet; so it was an opportunity to get this story into the public domain.

On top of all this, I already mentioned the chapter which Chris Rey and I wrote for the book on an introduction to superconductivity. I was very happy to have the opportunity to contribute to such a chapter because I don't think a good accessible overview exists yet in the literature bringing the original low-temperature superconductivity and the new high-temperature superconductivity into juxtaposition, melding them together. Both perspectives are essential elements in our understanding of the whole field. So in this chapter, we have now, I think, this broader picture. This was a wonderful project for me because it was almost like reviewing my career in superconductivity, going back to those days at IBM when I worked on flux creep and all those related phenomena. Now I had the opportunity to pull this all together in this review, which will become, I hope, a valuable book in the field. It’s a nice way of tying together many strands in my career.

Recognitions and Reflection

Hochheiser:

How would you summarize your career as a whole if you had to do so?

Malozemoff:

First of all I can say that it's been wonderful. I have to be very grateful for whatever choices I made to get into the fields I did, to associate with the people that I have associated with, in the exciting and encouraging environments I've worked in, on the variety of topics I've had the opportunity to address. People have different kinds of productive careers, some where they remain very focused, others where they try many different things. I've been in that latter group, very fortunate to have such a variety of stimulating experiences and to be able to contribute in so many different ways. There were certain themes in my career, obviously. In particular, my specialty has been in the magnetic properties of materials of different kinds, superconducting materials as well as ferromagnetic and antiferromagnetic materials and spin glasses, but within this field, I was able to span both fundamental issues and applied issues, to work on commercial applications as well as to understand the fundamentals that were underlying them. It’s been fun. These fields are so dynamic; they're always changing, with new ideas and new opportunities appearing all the time. At times it has felt as if the world was collapsing, and then somehow you find a way through it. So many amazing coincidences have occurred in my career. I've told some of these stories, for instance the way AMSC got into power electronics that grew the company so dramatically, from an application, SMES, that we initially thought was so great, and that then just faded away. By the way, I should say that SMES does have a role for certain applications where you really do need the real power; so it shouldn't be completely waved away, but for this particular kind of power quality application, dealing with voltage dips and sags in the utility grid, the all-power-electronic system DVAR is the answer.

But back to my career, it has been really wonderful. High temperature superconductivity has been one of the most amazing developments in modern science and we hope also in modern technology; so it has been great to be able to participate in this, to work with people like Alex Müller, for example, - he's just such a fascinating person - and many others in the international scientific and technical community. I've had the benefit of knowing these people, of traveling to so many countries, working with, competing with, negotiating with, the Europeans, the Japanese, the New Zealanders… I hardly mentioned the very stimulating collaboration that we had with New Zealanders for almost two decades at AMSC. We started the collaboration based on the fact that the BSCCO-2223 material was actually first identified and discovered in New Zealand. Although the active technical collaboration has died down now after 20 years, it was a really exciting collaboration while it lasted. Going down to New Zealand was always fun; it is such a beautiful country.

And I am of course thankful for the recognition I have received, particularly with the IEEE Award for Continuing and Significant Contributions in the Field of Superconductivity, and being named Fellow of both IEEE and the American Physical Society. The financial benefits of my leadership position in AMSC were also a nice plus.

And now in retirement, I've also had some time for myself. I have a wonderful family – a lovely wife and three fine sons, all now pursuing their own interests. And I have a lot of interests of my own. So I am very grateful for so much in my life; I could hardly ask for more.

Hochheiser:

Is there anything else you'd like to add that we neglected to cover?

Malozemoff:

I guess you had this last item here in your list of questions: thoughts on how superconductivity and its applications have evolved over the years. This is, of course, a story that is not over. In some sense there's still a real question: will high temperature superconductors make it into truly successful and widespread commercial products? We know that low temperature superconductors have their role; they're very successful in MRI and in high energy physics, and so forth. High temperature superconductors look very promising, but they aren't there yet. There isn't real commercialization of the power equipment technology. Yes, we could say that HTS wires are commercial, since people can come and buy wire with good properties. But the power cable projects are still mostly government supported, and in this, there is perhaps one last theme I can mention.

We are, of course, very grateful in particular to DOE in this country for the program that they established to develop high temperature superconductors for energy applications. A very farsighted individual who deserves a lot of credit, Jim Daley, formulated the program which had various names over the years; Superconductivity Partnership Initiative was one of them. He encouraged vertical collaborations in applying for funding support - they would involve a utility, an original equipment maker (such as the cable maker, the motor maker, or whatever), and the HTS wire producer, all working together to produce something like the LIPA cable. These projects were tremendously significant in furthering the whole field, making the U.S. effort really the leading effort around the world in developing superconductor technology. The Japanese were always very strong in this field; there were some beautiful efforts in Europe too, but I think the American effort as a whole was the most significant.

When the DOE had decided enough was enough and they essentially cancelled their program, it was pretty traumatic. There was a fear that the whole field would collapse. It hasn't, of course, and a number of other agencies have stepped in, such as ARPA-E and the Department of Homeland Security supporting the fault-current-limiting cable Hydra project. Now, the irony is that when new commercial projects are proposed to utilities, two decades of funding from the government has taught them a certain lesson, that there's potentially government money out there to co-fund the project. Why would a utility fully fund a commercial project when they could do it more cheaply by getting government funds? And then, you're into the whole struggle of obtaining government funding, which takes years to establish. As a result, government funding has become a kind of obstacle to true commercialization because everyone has become addicted to government support.

End Tape 2, Begin Tape 3.

Hochheiser:

Okay you can continue.

Malozemoff:

Just to wrap that up, it's evident from this, how important the government support is in the early stages, but the danger, as you're trying to make that transition to commercialization, is that even the customer community gets addicted to government support, and then you have a real problem for commercialization to really take off because everything gets slowed down by governmental budgetary considerations. That's just a comment about the development of the superconductivity field as a whole. I trust that we will get past this problem eventually, but we're still at that stage where this issue is an important one.

Hochheiser:

Anything else you'd like to add?

Malozemoff:

I think that's it. I've covered quite a bit.

Hochheiser:

Yes we have and I thank you for your time and it's been a pleasure.

Malozemoff:

Great. Okay.