Charles Kennel

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

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

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

Please contact [email protected] with any feedback.

ORAL HISTORIES
Charles Kennel
Interviewed by
Ryan Hearty
Interview date
Location
La Jolla, California, U.S.A.
Usage Information and Disclaimer
Disclaimer text

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

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

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

Preferred citation

In footnotes or endnotes please cite AIP interviews like this:

Interview of Charles Kennel by Ryan Hearty on December 18, 2019,
Niels Bohr Library & Archives, American Institute of Physics,
College Park, MD USA,
www.aip.org/history-programs/niels-bohr-library/oral-histories/XXXX

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

Abstract

Dr. Charles Kennel, director emeritus of Scripps Institution of Oceanography and (currently) Visiting Research Fellow at the Centre for Science and Policy, University of Cambridge, is interviewed at his home in La Jolla, California, by Ryan Hearty, oral history fellow at the American Institute of Physics. Kennel describes several milestones in his diverse career spanning industry, academia and government service. Subjects include: Kennel’s childhood in Boston; undergraduate studies at Harvard University; doctoral research at Princeton University, including work experience at Avco-Everett Research Laboratory; postdoc work at the Abdus Salam International Centre for Theoretical Physics (ICTP) in Trieste, Italy; his academic career as professor and later chair of physics, as well as vice chancellor, at UCLA; his service as associate administrator at NASA; directing Scripps; and recent work in climate change policy.

Transcript

Hearty:

We can begin at the beginning, I guess. This is Ryan Hearty. It’s December 18, 2019. I’m here in La Jolla, California, with Charlie Kennel, and we’re here to talk about Charlie’s experience as a scientist and later as a policy expert. Thank you so much.

Kennel:

It’s a pleasure to be here and a pleasure to have you here, and I’m flattered, in fact, that you’re here.

Hearty:

Great. Let’s start at the beginning. Give us a sense of your childhood growing up in the Boston area. Right? Was it Newton? You were born in Cambridge.

Kennel:

I was born in Cambridge at Cambridge City Hospital 10 days before the Second World War started; my family had quite a varied history until the war was over. (After the war) we ended up living in Newton, Massachusetts, where I went to Newton High School, then, as I think now, one of the best in the country. At that time (1955), they had a special program, called the Kenyon Plan, where they would put a few of the brighter students in the same class in order to qualify them for what we would now call advanced standing. After several years in the program, when I was a junior, I came home one evening and said to my parents — you know, “Bernie Lettau (who eventually became a colleague who ended up running the Antarctic program for the National Science Foundation) is leaving after our junior year to go to Yale, and I’m as smart as Bernie.” [laughs]. My father looked at me and said, “You can go to any school you want, so long as it’s Harvard.” [laughs]. My mother was a bit horrified, because I was young, and she felt, with some justice, that I needed social development. So they reached a compromise; I would apply to Harvard and Exeter Academy simultaneously. In the event, I got into Harvard but did not get into Exeter. I was very happy with the outcome. I recall my first interview at Harvard with my advanced standing supervisor. His name was — I don’t want to give his name, actually —he was a very ambitious gentleman. He started by asking “What do you want to do in life?” So I said, “My father is a lawyer, and maybe I’ll go into the law.” So he said (in effect), “Do I have a deal for you!” I was two weeks over 16 at that point. “I can arrange a five-year program for you; at the end of that time, you will have a bachelor’s degree and a law degree, after five years. You can qualify for the bar at age 21. I said, “I don’t know. I don’t really know what I want.” But he continued, “Well, the first thing you have to do is to get your science requirements out of the way, since you’re going to have to take a lot of social science, government, and so forth, to prepare for the law.” The solution was to take the general education course, called “Nat Sci 7” (Natural Sciences 7), that was taught by Bart Bok, the still famous Dutch astronomer, and the then famous geologist, L. Don Leet. (L. Don Leet was vehemently against continental drift, which is what I remember him for.) In the section of the course taught by Bart Bok, we had to write a paper, so I wrote this paper. After class one day in Allston Burr Hall (I was sitting in the very back row, and way down there was the professor): he said, “Will Charles Kennel come and see me after the lecture?. I walk down the stairs against the flow of students going upstairs; I come up to Bart Bok and say, “Sir, I’ll rewrite the paper if you want.!” But he asked, “Where did you learn all of that astronomy?” I replied, “I had a subscription to Scientific American and the only articles that I could understand were the astronomy articles. The rest were more technically complicated, but I used to read a lot of astronomy.” “Have you thought about becoming an astronomy concentrator?” (Concentrator was a Harvard word, they use the word major elsewhere) I replied, “No.” but he then said, “Why don’t you come to our next astronomy concentrators’ dinner at Leverett House?” So I went to Leverett House and contrived to sit in the very back of the dining hall as the great Dutch astronomer, the greatest one of the era, Jan Oort, was giving a talk. This time, the same thing happened. After the dinner, Bart Bok said, “Will Charles Kennel come up and meet us?” He introduces me to Professor Jan Oort, and he says to Oort, “I want you to meet our next astronomy concentrator.” [laughs] I think I had been intimidated by my advanced standing advisor, and was grasping at a straw. Somebody wanted me; It (Astronomy) seemed like a place to go, and I kind of figured that I was so young that if I made a mistake (by going into) astronomy, I could do something else, and I still wouldn’t be that far behind, so I signed up.

At that time there was this John Birch Society that was full of very right-wing people. They were saying then, as happens now, that Harvard was brainwashing people (with left-wing ideas). My parents got a letter from them, saying (in effect), “He’s majoring in science, and he’s about to be brainwashed by Wendell Furry, a known Communist, (a) lecturer in physics”. My parents, (were) somewhat horrified, (they were naturally concerned that someone would go to the trouble of writing about me). My mother (indirectly) conveyed her reluctance at seeing me go into science by saying, “You’ll never make any money in astronomy.” My father agreed. So I said to them what I just said to you: “I can always change. I’m young. I can do what I want. They (the Astronomy department) want me, and besides, astronomy’s much better to major in than physics, because you have to learn physics (anyhow) before you can get into the astronomy.” In some ways, astrophysics was a stronger education than just plain physics, so I went into it. The next year, Sputnik was launched, and my parents’ attitude changed dramatically; what I recall was a sort of an attitude change, maybe not explicit; they said (in effect): “How did you know where the future was?” [laughs] “Our son has latched onto something big in the future.” In the meantime, I had already decided to stay with it, probably because I noticed (what had happened to) my astronomy student colleague, John Gaustad, who actually had learned how to program a computer at Minnesota Mining in Minneapolis; he had been pressed into service by the Smithsonian Astrophysical Observatory as an undergraduate to program the codes that tracked the Russians’ Sputnik. Once again, I thought to myself, “You know, I could do that.” John Gaustad went ahead and had an important career at Berkeley, and ultimately, I think, at Swarthmore. But I continued along (as I had); I stumbled into theoretical physics, thinking to myself I’ll take advantage of doing — of learning--both physics and astronomy. I tried that, not with great success. My grades were not all that good, but I did take Applied Mathematics 202 (boundary value problems), and Quantum Mechanics, and so forth. Another thing happened by chance; it seemed to me that I fell into it. My junior year in college, Frances Wright, the (Astronomy) departmental secretary, a legendary woman, called me up one day and said, “Charles, do you think you’re going to graduate school?” And I replied, “I haven’t given it much thought. I don’t really know what I’m going to do.” And she said, “We like our best students to go to Princeton for graduate school; a Princeton professor is coming up for an interview, and we’d like you to go to that interview.” And I said, “Well, I’m not sure.” She said, “I’ve already made the appointment for you. Can you come at such-and-such a day?” I got applied to Princeton graduate school that way; I went and never regretted it.

Princeton, as a university in science, was then at the top of its game. It had always been world leading in mathematics; it had a first-class theoretical physics group, and its astronomy department, headed by Lyman Spitzer and Martin Schwarzschild, was very small but extremely, extremely good. I started taking astrophysics (My tutor was George Field). I was enjoying it, but I got mononucleosis in the midst of my first year in graduate school; I was 19, I guess, at the time, maybe 20. I learned later that one of the side effects of mononucleosis was occasional depression; I got depressed and decided to myself, I’m not really sure I want to do this; I’d like to take a little time off and maybe earn some money — my father had died. My father died after my sophomore year, my mother had moved to Pittsburgh, and money was short. I thought maybe I can go work for industry for a while and just sort things out. I went to Schwarzschild and said, “I’d like to take some time off.” There followed a complicated discussion that I can’t trace any longer, but the upshot (after some weeks) was that Schwarzschild (who evidently had discussed it with Spitzer and others) finally said to me, “We were thinking about what we should do with you. We understand your reasons to want some time off; to think out what you want to do.” He talked about Henry Norris Russell, the great astronomer who also had taken a year off (and later came back for his degree and became a Princeton professor). Schwarzschild and Spitzer said, “We first thought that we’d send you to the Meudon Observatory in Paris. You’d have a very good time there. But then we didn’t like that idea because we thought you probably would never come back.” Then they also said, “We were thinking of sending you to some of the aerospace companies out in California, but you probably would never come back from that either. But we do have a place for you, and that’s called the Avco-Everett Research Laboratory. And they’re beginning the study of plasma physics,” which was Spitzer’s main field at the time and his physics competitor to his great achievements in astronomy. “We thought we should send you there.” I applied, and I got hired. I arrive on the doorstep of Avco-Everett Research Laboratory with my professors’ (Schwarzschild and Spitzer’s) requests (in hand) to look after this young kid, but Avco didn’t quite know what to do with me. As you probably know, when you get volunteer help, there’s always a cost. Avco were paying me, but basically I was volunteered. There’s always a cost to the mentor, and it wasn’t clear that anybody [laughs] really wanted to deal with me. But finally, one of my great mentors, Harry Petschek — who was at Avco-Everett and one of the best physicists I’ve ever worked with — calls me into his office and basically says, “We don’t know what to do with you, but I’ve got an idea, and the idea is this: we’ve been studying collisionless shockwaves in the laboratory.” They (Avco) had been studying collisional shockwaves for a number of years; they were in the missile re-entry program and ablation heat shields were just coming in as the way to make a missile re-enter without burning up. There would be a shockwave (standing ahead of a re-entering missile) with ablated materials from the nosecone participating in the shock; the radiation of those ablated atoms would carry the heat energy away from the missile. They’d been studying collisional shockwaves, but at the same time that they’d been doing that, controlled thermonuclear fusion research had been declassified. That required fully ionized plasmas, no atoms in it to radiate energy away (in fact, fully ionized plasma had been Lyman Spitzer’s main interest at the time — he wrote a famous book called The Physics of Fully Ionized Gases.) Somehow, Avco thought that what they ought to do was put me to work on studying collisionless shockwaves. They’d been trying to make them (collisionless shocks) in the laboratory, trying to rid their laboratory plasmas of atoms to create to the best possible vacuums, and so forth. They’d utterly failed to make them (collisionless shocks) in the laboratory, no matter how much they tried. But Petschek said to me, “This space physics has just started. There is a solar wind. (It is collisionless.) The solar wind will blow past the Earth, and the interaction of the solar wind with the Earth’s magnetosphere” — I’m now using modern terms — “would create a shockwave.” This will be the first collisionless shockwave; satellites will discover it and we’ll be able to diagnose it with satellites (data). I want you to read the space physics literature, and especially the space plasma physics literature, and get up to date on it. You will teach me what’s in that literature, and I will teach you plasma physics; when that discovery is made, we want to be amongst the first to interpret it (the data).” So I started out by reading the plasma physics literature and the space plasma physics literature, both of which were relatively thin at that time; I was relatively smart compared to now, and I was able to get past the main requirements that he (Harry) had on what I should know and (have time to) read rather more generally in the field. At that time, I started reading about the Earth’s radiation belts, the Van Allen belts. They had recently been discovered; remember, they were discovered in 1959 by James Van Allen, and by 1960-61, it was clear that they (the energetic electrons) would be trapped in the earth’s dipole magnetic field. There had been a theory developed by Hannes Alfvén that there would be an adiabatic invariant that the particles that orbited around the magnetic field would maintain as a constant of the motion. The area contained within the Larmor oscillation area (this adiabatic invariant) would be conserved as the particles moved back and forth along the field lines. Ultimately then, as they (the Van Allen Belt electrons) penetrate larger magnetic fields, the magnetic flux would be conserved through that Larmor radius, they’d be reflected, and they’d go back. That adiabatic invariant theory (Alfven’s) was used in magnetic mirror machines for fusion research; it turned out that the Earth (dipole magnetic field) was an analog for the fusion machines they (The US, Soviet Union and Britain) had been studying and building since mid-1952, ’53. The big concern in the thermonuclear fusion (community) was losses of particles from the mirror machines. You had to confine the plasma for a long (enough) time in order to allow time for the nuclear interactions to occur, and if there were losses, the plasma would lose energy before you could actually get to fusion reaction (energies). They (fusion scientists) had this concern, and of course, nobody had built a mirror machine that confined plasma for a long (enough) period of time, and it’s still hard. So there was a great deal of interest in the space — well, some interest in the space community- about the losses from the Earth’s mirror machine. I began to read about those losses; they turned up when the particles left the trapping region of the Earth’s magnetic field. They (the energetic particles) came into the Earth’s atmosphere; in colliding with the atoms of the atmosphere, they created X-rays; there was a flux of X-rays at the bottom of the field lines in the polar regions, (that) indicated that the electrons trapped in the radiation belt had a rather short lifetime. This had some significance as a kind of example: Here was another device, a natural device that was like a fusion device, losing particles copiously. I started to discuss all of this with Harry Petschek; it was clear that for Alfvén’s adiabatic invariant to be violated, (we inferred) there had to be fluctuations, electromagnetic fluctuations, in the radiation belts (with frequencies) comparable to the cyclotron frequency with which the electrons oscillate — rotate around the magnetic field; that there would be a kind of resonant violation of — when you had wave activities with the cyclotron, or near the cyclotron, frequency there would be resonant interactions, the adiabatic invariant would be violated, and you could scatter them out of the radiation belts. This was the problem that we had set out for ourselves; we got through all the conceptual formulation of that problem during my first year at Avco, and then of course, I went back to Princeton to finish up a degree on another topic. The idea was that the way the particles were trapped and lost from the Earth’s magnetic field led to a certain distribution of particle velocities in velocity space, if you will; that that distribution would be naturally unstable, and that that instability would occur at frequencies close to the cyclotron frequency. It was a natural process by which you developed the waves that would then subsequently scatter the particles from the radiation belts. Our prediction was that the particle intensities in the Earth’s radiation belts would never get far above the threshold for that instability, because once the instability occurs, the waves grow, the waves scatter the particles out to the radiation belts, they’re lost to the atmosphere, and you start all over again. Ultimately, that was the paper that we (Petschek and I) published, “Limit on stably trapped particle fluxes.”

Hearty:

That was in 1966?

Kennel:

That was ’66.

Hearty:

Right. Okay.

Kennel:

Then I go back to — I think another thing that’s important for (my) subsequent development: the theory of the whistler mode instability that Petschek and I used had originally been discovered in parallel by the United States and the Soviet Union, within their then-classified research programs; there (also) was a theory of the loss-cone instability that was developed by Dick Post for, I think, protons —(Basically the same theory had been developed by) Dick Post, from Lawrence Livermore in the United States, and on the Soviet side my subsequent friend and colleague, Roald Sagdeev and his sidekick, Shafranov. I knew about these theories. They were (developed earlier) but published around 1958 after the declassification of the fusion program. In any case, (in September 1961 about, I leave Avco with this theory half-finished, and I go back to Princeton (to find) a thesis advisor. By this time, I had decided that what I will do is leave astrophysics behind, .. i.e. the interstellar medium (that I had started to work on with George Field) and things like that, that I’d started on, and go into their (Princeton’s) controlled fusion program, (Project Matterhorn) which became the Princeton Plasma Physics Laboratory. I think I was the second student they had — I’m not sure — an early student. I decided I would go into that program and work with my ultimate thesis advisor and yet another mentor, Edward Frieman. (With him), I struggled along on a theory of drift waves in finite beta plasmas, but I couldn’t get the Van Allen Belt problem out of my head.

I might as well tell this story, too. Frieman was then — he seemed like an old man, but he was probably about 35, and he was already a major advisor to the government and one of the founders of the JASON program in La Jolla that advised the Defense Department on issues of physics. At one point, Frieman calls me and said, “I want to talk about your work with Petschek.” So we talked about the loss of particles (from) Earth’s radiation belts, and he said, “Someday, somebody’s going to artificially put particles in. If somebody injects particles into the radiation belts, how long will they last?” And I said, “There is this instability, the whistler mode instability, and probably it will grow,” and so forth; [I] explained the violation of the adiabatic invariant, which he (Frieman) understood and (undoubtedly) previously understood. So probably, they (the energetic electrons) wouldn’t last very long. (The conversation over) I go back to my work on a thicket of equations on drift waves, and he goes off someplace. He was always traveling, anyhow. But later on-he’d been away for a while- I had asked him where he’d been. He said, “I can’t tell you,” but he had a tan.(It was winter in Princeton.) Later on, in 1962, America exploded the Starfish (nuclear device) at about a few hundred kilometers altitude in the atmosphere, but on magnetic field lines that injected relativistic particles into the Earth’s radiation belts.

Initially, a lot of them (the relativistic particles) were dumped into the atmosphere. You couldn’t miss it. The sky lit up in an artificial aurora, and (passengers on) airplanes … saw the X-rays. Frieman said, “Now I can tell you.” And that was that — it was 1962, the radiation belts had been in existence for three years, and the British astronomer, A.C.B. Lovell, the radio astronomer, had protested very loudly that America’s plans to launch a nuclear weapon into orbit, and then explode it, would ruin the radiation belts forever. The particles would be there, and you would never know what they (the radiation belts) had been like before. Frieman said, “Now I can tell you (i.e how I got the tan). President Kennedy had called the JASONs and others together to give me (Kennedy_ advice on whether we should go ahead with this test; we finally agreed that we (the US) should — there’s no reason not to go ahead.” So they did have that test. I’ve never asked him (Frieman)— and now cannot — whether my idea that the particles would leave the radiation belts had any influence on their discussion. It turns out that there was a very rapid initial loss, but then they actually did stay for about 10 years; I put a graduate student, Richard Thorne, on that problem, and we figured out why these particles lasted 10 years, whereas the main particles in the radiation belts had lifetimes of minutes, and so we understood that. (This story is recounted in my memorial to Ed Frieman in the Proceedings of the American Philosophical Society)

So now, (let’s go back to) Princeton — go back (after) 1962, and the Starfish test and all of that; I’m working on my thesis, and somewhere along that time, Frieman announces he’s going on sabbatical. He won’t be there for a while. I had an assistant thesis advisor, John Greene, who was actually very good, and I appreciated him very much. But Ed wasn’t going to be there, and I said to myself: if he’s not going to be there, I’m not going to be there. [laughs] And so I said, “I want to go back and finish that radiation belt paper before somebody else gets it.” I leave Princeton without finishing the Ph.D., say that I’ll finish the thing at night, and go back and work with Harry Petschek on the radiation belt paper.

Hearty:

But you found their interests had changed. Right?

Kennel:

Yeah, then — no.

Hearty:

Not quite yet?

Kennel:

That wasn’t quite then. At that point, I came back — it was probably ’64, ’65 — and at that point, Petschek got even more interested in the radiation belts, and he decided that he would use the carrot of summer consultantships to convene various people who looked like they were experts in space physics and bring them to Avco for extended visits: a week, two weeks. We had summer seminars that he (Harry) had managed to get the company to finance. I don’t know how he did it, but he did. And this was, I think, where many of the various ideas that we had about — fundamental ideas about the magnetosphere were formulated, during those summer programs. There’s a famous paper on the — Axford, Petschek and Siscoe, George Siscoe, on the tail of the magnetosphere, the role of reconnection in shaping the Earth’s magnetic field and its variability. I think that was the main one. We did some work on propagation of whistler waves that was the precursor to understanding why the radiation — the Starfish electrons didn’t go away. It was a very seminal time. We invited quite an extraordinary group of people, all of whom became prominent in the field later, but were all of the same youthful age group. Ian Axford, Richard Thorne, Don Gurnett, Neil Brice, who died young;. I think I’ve compiled a longer list of names someplace else. But in any case, the point was the young people got together, and we kind of formulated a view of the Earth’s magnetosphere and its interaction with the solar wind, which is the one that we have today. And we didn’t — of course, it’s much more sophisticated now, and it’s been verified time and time again, and there’s fantastic data now that didn’t exist at that time. But the basic conceptual outlines of the interaction of the solar wind with the Earth’s magnetic field, and the reconnection model placed in the context were refined by work there and similar work (elsewhere) that was — you know, ideas were in the air, so it wasn’t unique what we (at Avco) did, but it was ahead of its time, or leading its time. Of course, the idea- later on, as I got into the history of science I, of course, learned that the idea that material was flowing outward from the Sun and interacting with the Earth was actually an old one, and it went back to the 1920s and 1930s with Sydney Chapman …. What we (at Avco) had done was insert into Chapman’s point of view a greater sophistication about the fluid flow around blunt bodies, which was a specialty that was much pursued at the Avco Research Laboratory because of the interest in missile re-entry. (We also added) Harry Petschek’s understanding of the reconnection process, where magnetic fields dissipate and reconnect — I think they were the two things — and so what we actually did take was the Chapman model and make it more complete and able to handle the time variability of the processes — conceptually able to handle the time variability better. This model, of course, did have a collisionless shockwave standing up ahead of the Earth, the Earth’s bow shock. I think that it’s not clear that we had all of that in mind in 1961, but by 1964, the picture was becoming clear. Of course, many other satellites had been launched in between, and so we knew much more than we did in ’61.

Hearty:

Let’s back up a bit and just review some things that you mentioned. If we go all the way back, I wanted to ask you more about why it is your dad wanted you to go to Harvard.

Kennel:

There was tension between my parents on that issue. My mother was afraid, I think, that I would become a nerd, or maybe that I was on the spectrum, and that I would be socially maladjusted; this was a great fear of hers. My father, on the other hand, was a frustrated lawyer, a very good technical lawyer who did not prosper in the aristocratic law firm in which he worked. And he had had a — there’s a seminal experience in his life which my mother did not share. He was a poor boy from Missouri. He graduated from Westminster College in Missouri. He was, I guess, a wizard as well as a good tennis player. He got himself a master’s degree at Vanderbilt, and he ended up doing — I guess participating in debates at Oxford. Then he went off to Harvard Law School; a he graduated in 1936 from the Harvard Law School, and went almost immediately to the Federal Reserve Bank of Boston, which was then in its initial formative stages. He worked at the Federal Reserve for a few years: this then got him a position as a lawyer, an associate, at the law firm of Choate Hall & Stewart, which was the oldest law firm in Boston. It’s not now the biggest, but the oldest, and a real white-shoe firm; so he worked along with them. Then when the war started, he got an assignment. He was pulled out of the law firm and got an assignment with the War Department, actually — the War Department asked Choate Hall if they would ask my father to work for Lockheed, and they (Choate Hall) did (say yes). The story was that Choate Hall had had a partner named Robert Gross, a famous man. This Robert Gross had bought a small, struggling aerospace firm from a gentleman named Allan Lockheed. When the war was starting, Lockheed’s chief counsel was drafted into the service; they (Lockheed) needed (a lawyer) to negotiate fighter plane contracts with the federal government. [laughs] So my father was pressed into service to work for Lockheed, and he did. He had, I think, a career that actually — you know, nobody liked the stress and so forth of the war, but I think he actually prospered; he spent a lot of time traveling between Burbank, where Lockheed was at that time, and Dayton, Ohio, where the Wright-Patterson Air Force field is. He went back and forth for quite a bit of time; he discovered the aerospace industry, and he wanted (to go into it), as I understand it — I was not privy to family arguments — at the end of that time (the war), he had a choice, and that was to go back to Lockheed, not in the original high-level position that he was assigned, because the original person was coming back, but he could have a job at Lockheed as a lawyer, or he could stick with Choate Hall & Stewart. My mother’s attitude, summarized briefly, is that nobody we know lives west of Dedham, and she prevailed.

Hearty:

So she was Boston-raised, your mom?

Kennel:

Yes, she was Boston Irish, and of the type, like the Kennedys, that were (sic) incredibly jealous of what the WASPs were doing in the North Shore; and always wanted — somehow felt one-down, that they were not accepted. I think that was an underlying motivation. In any case, they had this tension between them about whether I should be socially adept or intellectually ahead. Ultimately, both of them had an influence. [laughs] But I think that’s how it happened. (After the war) my father was (back to being) an unsuccessful lawyer in the sense that he was not a rainmaker in the firm. He was not socially well connected; he was from the Midwest; he didn’t have aristocratic Boston habits, accent, characteristics. He was a Midwesterner, so he never got far in the firm, and he never made partner — the equivalent of tenure. And he died young. He died in 1956 at age 49, of heart disease. On the day of his funeral, his boss, Mr. Pengra, who was the number two or three person in the firm, came up to me and he said, “You know, he was a very unhappy man in his profession, but I think I found the solution for him; had he lived, I would have recommended him to the Securities and Exchange Commission, and he would have made it, and this would have utilized all his technical skills as a lawyer.”

Hearty:

The idea came a little too late.

Kennel:

Yeah, and so I think that in this background, I had the background of a dissatisfied father and a dissatisfied mother but dissatisfied over different things.

Hearty:

Did you have other siblings?

Kennel:

Yes, my brother John. The other thing that I think is worth saying is that my mother was an alcoholic and had been for — I don’t know when she started, but it was certainly likely during the war, and had been an alcoholic through my young years. And my brother, who was four and a half years younger, she was completely unable to handle him, so he had a much more unhappy life than I did. And I escaped, basically, into work.

Hearty:

Right. I was wondering about your other interests. You mentioned you were reading a lot of Scientific American and that you were considering law. You were okay with it.

Kennel:

Yeah. I didn’t know. I mean —

Hearty:

Usual kid.

Kennel:

When I was, I guess, between my freshman and sophomore years, my father got me a job at the other big firm in Boston, which is now huge, the firm of Ropes & Gray, also similarly ancient. I think Choate Hall went back to 1840. Ropes & Gray was younger, but it was bigger. I was put in the charge of one of my father’s friends, Allen Eaton, and asked to do various things. Initially, I started off in my summer job as a messenger boy, taking messages back and forth. We didn’t have internet in those days. [laughs] I did that for a while, but ultimately they put me to work doing library work, just looking up things. At the end of that summer, Mr. Eaton calls me into his office and he says, “Charles, are you thinking of going into the law?” and I said, “Yes, sir. No, sir. Yes, sir. I mean, I guess so.” So he says, “Well, if you’re considering going into the law and you actually do so, we hope that you’ll consider working for us when the time comes.” It was pretty nice. Then he said something that turned me off. He said, “You know, we only take the number one or two man, and from Harvard or Yale law schools. This is what you have to do to get to work for us.” [laughs] And I think that turned me off. It was like my encounter with my advanced standing (advisor). They were trying to rush me, I think. All of these things, I think — much as I loved my father and much as I respected his deep attachment to the technical parts of the law, these things made me feel very ambivalent about going into it, and I think that, historically, science became prominent right at the time I was at this liminal moment; I suspect if it had been the Civil Rights revolution of 10 years later, I might have done that, but history touched me at that moment.

Hearty:

And it sounds like Bart Bok was also an influence on you. He encouraged you.

Kennel:

Yes. He made clear that there was a future there.

Hearty:

Yeah, and how about the tools that Harvard had in the astronomy department — they had the observatory. I think the Smithsonian Observatory merged around the same time that you were there.

Kennel:

That’s right. Harvard Astronomy, it turns out, at that time was in very great trouble with the dean. The dean’s name was McGeorge Bundy, by the way, and Bundy was a tough son of a gun. But the problem was that they (the Astronomy Department) had had an undergraduate, and probably graduate, program, (in which) the students were doing very badly. They were doing very badly, and it became important — the reason that I was sent on to Princeton (and told that I should do that) was that they (the Harvard Astronomy department} wanted — they were under the gun-- to recruit academically successful young people into the astronomy program to repair its reputation. In my class (1959), there was a very small number of concentrators, as I recall: myself, not unsuccessful; John Gaustad; and Peter Stone, who became a world-famous meteorologist at MIT. So there were three of us; I considered myself the third amongst those three. I think I ended up first amongst those three in life but third amongst those three at Harvard. So they (Harvard Astronomy) were desperate to prove that they could attract first-class intellects, and so I think that’s why I was sped along, [laughs] is that they probably said (to Bok), “You need to find some bright undergraduates and get them into the program.”

Hearty:

Right. Let’s go to — right. I’m just checking if there is anything else I wanted to review from what you said. I guess that’s enough for Harvard. A couple more questions about Avco. What was the kind of mix that you noticed there of disciplines? It seems like there were physicists working closely with engineers.

[TAPE PAUSED]

Hearty:

I was going to ask you just a little bit more about Avco.

Kennel:

Yeah, and what were the mixtures of disciplines.

Hearty:

Yeah, because it’s a very different kind of — I think it was fairly new when you arrived there.

Kennel:

Well, I don’t quite know when it was founded, but my guess would be 1956, 1957.

Hearty:

Yeah.

Kennel:

It was founded as part of the defense buildup; its main purpose was to aid the country’s missile ICBM program, basically. They (Avco) had the job of working out the physics of missile re-entry, and this included fluid (mechanics) — oh…, its director was Arthur Kantrowitz, a very great gentleman who had been, I guess, chairman of aeronautics at Cornell; he moved some of his people, including Harry Petschek, into this (new) company that was on recovered land in basically a slum around Boston in Everett, Massachusetts. They (Avco-Everett) built this building out of cinder blocks, [laughs] and started research; disciplines they were interested in (in addition to aerodynamics) were atomic physics, radiation physics, and some plasma physics. They got into plasma physics as the temperature of their gases increased and they (the gases) became more and more ionized. There was fluid mechanics. There was the engineering of shock tubes. There was a considerable degree of understanding of the engineering required for the missiles and the space launch capability. They had to understand it, although they didn’t have to actually build rockets, so there was a whole variety of disciplines that were — another way of saying it (how pioneering Avco was) is that missile re-entry was taking an object that had been in space and bringing it back into the Earth’s atmosphere; they were on the fringes of space physics from the beginning.

Hearty:

And did their shock tubes — you must have seen many of them there. They had this experimental part of their work.

Kennel:

Yes, that’s right.

Hearty:

Yeah. How did that influence you? Were you part of that too, because I know you’ve done both theoretical and experimental work.

Kennel:

Very little. Very little at that time. I was influenced by whatever the shock tube results meant to Harry Petschek. He would communicate (the meaning of the results) to me, and so most of my understanding of the situation was filtered through him. Getting into experimental work was slightly different. When I was recruited to the UCLA Physics department, once again (after Ed Frieman), people noticed my Avco-Everett career; the person that recruited me to UCLA, Burt Fried, was also a consultant at Thompson Ramo Wooldridge, TRW Systems. TRW Systems had the main responsibility of being the system integrator for the entire national ICBM program, which meant they had to do a little of every aspect of the national program; they were also building very large spacecraft. Burt, as part of (his effort) to attract me to UCLA, got me a one-day a week at TRW Systems as a consultant. TRW Systems was a much bigger version of Avco-Everett Research Laboratory, with many more divisions and much more responsibility for the actual construction of hardware, of building things: building ICBMs, building rockets, building spacecraft. But they (TRW) also had a research wing in space plasma physics, as it turned out, headed by a gentleman named Frederick L. Scarf. Scarf died young, about 20, 25 years ago. Scarf’s main interest in life — research interest — oh, he was very prominently placed with NASA. He was working a lot with NASA programs and was highly valued in the space physics business. Scarf had invented, or caused to be invented, (the first) detectors for electrostatic plasma waves in space, and that was then a new thing. Since I understood plasma theory and understood how to apply it to the space environment, I was brought into Scarf’s team along with a younger colleague from UCLA, Ferd Coroniti, a close friend. Once the detectors (Scarf’s) collected the data, our job was to interpret the data in light of what was then the new plasma physics theory, the modern plasma physics — now, we were talking ’66, ’67. Plasma physics, in terms of open, unclassified research, was probably a subject that was 10 years old, at most, at that point. As a result, I got included on many of Scarf’s papers. In fact, much of the data that was presented went through our hands [laughs] before it was presented (at meetings) and interpreted. I was on the discovery paper for the first electrostatic waves in the Earth’s radiation belts, papers on electrostatic waves and collisionless shocks, papers on waves in the solar wind; all of those were part of this business of having been sped forward [laughs] by things that had happened in the past, sped into areas where you couldn’t fail to do something new, because you were sent in there. Nobody knew how to do these things. These companies were grabbing at whoever they thought might know something, and using it; as a result I ended up with some greater understanding of experimental (plasma physics) — I never really did it (myself), because if I walked into the room, there would be electrical discharges and lights flashing and loud bangs and so forth. I was not good with my hands, but I ended up with a native — by just working with experimentalists, both at Avco and TRW, and later (at NASA), I ended up with some sympathy and comprehension of what experimental research was like. That in those days was not common amongst physicists. Either you were a theoretician, or you were an experimentalist.

Hearty:

Now that we’ve moved a bit to UCLA, I wanted to ask about what connections you sustained from the East Coast when you moved. You were at Avco. You had the Princeton connections, Ed Frieman, and then when you became a professor — I know we’re skipping over your postdoc, which I want to ask you about in a second — but I’m interested in this move to the West Coast and what impression it had on you and what effect — did you sever ties? Did you keep some ties? Was this a big — this must have been a big move.

Kennel:

It was the move my father wished he had made (the move to California).

Hearty:

Away from the —

Kennel:

I never saw it as a big move (It was, of course.) Recently, while I was sick in the hospital, I saw a number of documentaries on PBS about the Los Angeles aerospace culture of that time (1950s and 1960s). I realized [laughs] how different it (LA) was from the rest of the country. I got into this; it seemed pretty natural to me, but it wasn’t (to the rest of the country), and I realize now that I was subject to a different kind of cultural influence: the can-do attitude, the big-science attitude, that you can do an awful lot by getting smart people together in big teams. All of that was not so common in the East, although Boston itself, and Harvard and MIT, were on the cutting edge, it still wasn’t part of the whole (aerospace) culture, the buildings, the people. It (Boston) wasn’t (part of) the space culture that we know about now. It (Boston) was a much more aristocratic, old intellectual culture, but not the go-go space culture that was in Los Angeles at that time.

Hearty:

Right. You notice this in hindsight, but at the time, did you — I mean, what was your impression? Was it excitement? Did you have — when you went — I mean, you were a new — you were just hired as a professor. That must have been —

Kennel:

Yeah, I was mostly concerned with getting ahead. I stuck my head down. I wanted to make a name for myself. I realize now I needn’t have worried, [laughs] but I did (worry). I had brought my first wife (Debby Bochner) out; I had a young child (Matthew Kennel) soon thereafter, and it seemed to me… — oh, I was hired to tenure. I’ve never been an assistant professor. So I shouldn’t have worried at all, but in fact all I wanted to do was write papers and earn the respect of my colleagues. I think this was where my mother’s social anxiety came in, that I wanted to be not only intellectually successful, but I wanted to be socially acceptable; I wanted to be wanted by these people (scientists), and so I kept my head down. The unfortunate thing was that I missed understanding the significance of the Civil Rights revolution that occurred in the 1960s. All of that was whirling around me. I was the youngest tenured faculty member at UCLA, and maybe even UC; I was (even) pressed into service on a student faculty committee to deal with student unrest, but all of that kind of rolled off me. I mean, I did it and so forth, but there was some (emotional) distance between me and commitment to those issues that I now regret. I missed a lot by not participating, I think. But I didn’t.

Hearty:

Yeah, because you arrived at the height of the Vietnam War protest, too.

Kennel:

That’s right.

Hearty:

So the campus was — do you remember the campus being in a kind of tumult?

Kennel:

It was never as bad as it had been at Berkeley, but I do remember one faculty meeting at the UCLA Physics department. We were there meeting, and the meeting was interrupted by — this was the time when students were taking over offices and using the expensive long-distance telephone to call each other and do what we would now do on social media—but they were communicating on our dime and we thought that was outrageous. I mean, they were using our money. But nonetheless, we had this meeting interrupted by about four beautiful young ladies with long hair in sort of angel costumes, white costumes; they were carrying candles sort of like this. They (the young angels) wandered down the aisle in the midst of this faculty meeting and begged us not to build atomic bombs on campus. Of course, the reaction of everyone [was], “We don’t do that,” and we don’t. In fact, UCLA had had a nuclear reactor — an experimental nuclear reactor-- they’d shut down. But what we didn’t tell them was that the University of California is the only university in the world that manufactures nuclear weapons, but through Los Alamos and Livermore. I think the feelings were running high. At about the same time, there was a riot on campus; the helicopters came, and the police shot water cannons and chased students all across the Social Science end of campus. I stood on the Physics department stairs and watched this happening, but with a sense of distance, and then went back to work. I’m not proud of that.

Hearty:

Right. And a lot of funding was also coming from Air Force, but also NASA, right?

Kennel:

Yeah. Most of my funding was from NASA. The military funding actually was beginning to go away, but I think it had an important influence in the ’50s in particular at UCLA; things like the first computer network actually started at UCLA. It was a collaboration between UCLA and John Von Neumann’s group (and the Rand Corporation, I believe); they actually connected up two computers to get them to talk to one other, and that was the beginning of ARPANET, and so there was a lot of military funding that way. But that kind of funding was beginning to go away during the period of student unrest; we (also) had Governor Reagan at that point. And so the dominance of military funding for fundamental research was waning already; It never went to zero.

Hearty:

Right. And you’ve always thought of yourself as working in fundamental research. Is that right? How did you — how are you thinking of these terms?

Kennel:

I saw fundamental research at its best at Princeton in my graduate classes, and in particular, I remember a class in — you didn’t have to take this to be an astronomy graduate student, but I took a course in quantum electrodynamics from Eugene Wigner; there were some very famous people in there, people that became famous, in that class: Steve Adler, Oscar Lanford and a number of other people. I saw what they were doing, and I realized that I didn’t have the flash-bang mathematical skills that these people had. They were just incredible. They could do anything, I thought. So I thought to myself: you’ve got to go into some other field in which you can use your verbal abilities and your other conceptual skills, non-mathematical conceptual skills. But I thought that this (fundamental theoretical physics) was wonderful. You need to know it existed. You need to be able to talk the language, but I could never originate in that field; that’s probably a true judgement. So then the question was, “How do you get ahead?” And the thing is, how do you know what a new field is? And my answer was, they’re not beautiful like the established fields. They’re ugly. There’s all sorts of stuff going on, and heuristic things, and the physicists are scorning the work that’s happening there, and nobody knows how to deal with the data, and things are a mess. That’s a new field. And space physics was like that in the early days; there was no established canon. Nobody knew what to do. Nobody knew which data to look at, what was good data, what was bad data. It was a bit harder to figure out. So I think my starting off — my first serious reading of literature was in a nascent field and not an established field.

Hearty:

So you didn’t have — I was going to ask you about that. You didn’t really have textbooks when you were starting out in Princeton in the first year, so you were looking at articles, mostly.

Kennel:

All articles.

Hearty:

Yeah.

Kennel:

My guide was Lyman Spitzer’s book on the physics of fully ionized plasmas, but that was to provide a conceptual guide to a confused phenomenological literature.

Hearty:

I think one of your Princeton mentors — or teachers at least — Thomas Stix, I think he wrote a textbook.

Kennel:

Thomas Stix wrote a textbook on plasma waves that was the bible for many, many, many years, and that was what, ’67, ’68. I don’t know exactly when.

Hearty:

Earlier than that, I think, he made the first edition.

Kennel:

Yeah.

Hearty:

But yeah, you didn’t use that, because you were already at Avco by that point.

Kennel:

Yeah, and I knew it (i.e, what was in Stix’s book).

Hearty:

Working on it, and you knew it. Right. So you had to — and you had mentioned maybe getting to Trieste, which I know you’re working on a piece for AGU on Trieste.

Kennel:

Yeah.

Hearty:

But you mentioned that that was a moment for you that seemed like a discipline was forming, because you noticed these different international researchers who had similar — actually had some things in common. Right?

Kennel:

Yeah. What really happened, I’ll repeat very briefly for you. I was at Avco, and Avco — Harry Petschek and Kantrowitz were very famous in the very small world of controlled fusion research, or basic plasma research. And when I was at Avco in my first year there, which was ’60-’61, there was a visit to Avco by my colleague and subsequent friend Roald Sagdeev and Leonid Sedov. Sedov was the primary serious person, he was an expert on fluid flow, and was the main person that Avco, dealing with fluid flows around missiles, wanted to talk to. But Sagdeev came along, and I met him the first time, and then subsequently doing this long — after I left Princeton (when) I went to Avco, during that time, Sagdeev made another visit, on his own this time. Well, with other people. I forget who — Soviets never traveled completely alone. Sagdeev came to visit Petschek, and Petschek introduced me to him; they had somehow a (private) discussion, and then Sagdeev was supposed to fly to Princeton to visit my thesis advisor (Ed Frieman). And he (Sagdeev) was not allowed to travel alone, so I joined him. It was agreed that I should escort him to Princeton; it was during that time that he told me that he and Marshall Rosenbluth were going to have this school on fundamental fusion plasmas in Trieste, at the International Centre for Theoretical Physics. His colleague, Alec Galeev, was going to work, he hoped, with Rosenbluth, and he — Sagdeev — needed somebody, an American, to work with, and would I be his disciple? He (Sagdeev) was an extraordinarily wonderful, compelling physicist, and he had this agile mind that you just recognize — this agile mind, able to frame problems that the best physicists always had. So of course, I accepted, and we went. This (the Trieste meeting) was all on fusion research, but all the greats were going to be there, it seemed. Not that there were that many at that time, but the ones that were there were all going to be there, or visit (Boris Kadomtsev did not come to the first Trieste seminar) . It was during that period of time, ’64-’65, that my reading of the space physics literature and relative mastery of it came to the fore, because we would discuss plasma physics and magnetic mirror fusion machines, then I would immediately pipe up with the analogue in space (The Van Allen Belts). [laughs] Ultimately, despite the fact that Rosenbluth only wanted to talk fusion, we did do some space research there. That was the time in which it became respectable, amidst scientific culture on both sides — the space side and the plasma side — to do rigorous plasma theory for the space environment. The two were joined at that point. The other thing that I learned there (was) about the role of international politics, (and about) the third world — the International Centre was set up to bring physicists from the developing world to the Centre. They would get advanced training and postdocs there, and then go back to their home countries. And I learned from the great humanitarian physicist, Abdus Salam, about the difference between international science and global science. International science was taking your Harvard Ph.D. (in Cambridge, MA) and going to Cambridge (UK). And then there was (Salam’s) global science, which was to bring the benefits of science to the development of the world. That had an influence on my later life that is undeniable.

Hearty:

So this idea had an influence on you, but then when you arrived, as you mentioned, there were so many different people that it almost — I think you said — you wrote that it was like, all you had to do was listen, and you could write a paper.

Kennel:

Oh yeah, that’s right. It was a young field. You remember, a primitive field is ugly, and nobody knows what they’re doing, but the other side of it is that the people that think about it, every thought that they have is new, so you could listen to this discussion, and things would pop up in the seminars, [laughs] and then we’d just go and write them down. It wasn’t quite that easy, but it was pretty close, and I can remember my roommate was Alec Galeev, and he was the fastest plasma computational person who could do the fastest plasma computations of anybody I knew, including my thesis advisor, Ed Frieman. So if Alec caught an idea, it got written down overnight. [laughs]

Hearty:

Did you keep contact with Sagdeev?

Kennel:

Of course, throughout the rest of my life, actually. Years later, I’m in Pacific Palisades at UCLA with my present wife, Ellen Lehman. And we’re sitting at dinner, about 9:00 at night, and I get this telephone call, and I answered the phone call. “Charlie, this is Roald. What are you doing? I’m here in Los Angeles.” “Oh, you didn’t tell me you were coming.” “I’m here at the Rand Corporation, and we’re discussing arms control, and we’ve just finished our discussion, and I wondered, did you have anything to eat? Can I come over for dinner?” Later on, he comes over for dinner, and he does not have an escort. He’s traveling on his own. This is 1987. This is perestroika, and he was on a fishing expedition, I think, for the Soviets to explore. Remember that Gorbachev took a profound decision, and that was to renounce — to settle, if you will, the nuclear conflict before changing communism. And so he wanted — he thought it was too — so in any case, Roald was there to discuss arms control on his own. That was when I discovered that actually, things were really changing. Then the funny thing was, we get to my house, and we’re having dinner, and he says, “Do you know how to reach my friend John Denver?” [laughs] We didn’t, but we tried calling Aspen, and we couldn’t reach Mr. Denver. They said he was unavailable, or something like that. But anyhow, Sagdeev was there (in California). We stayed in contact for the rest of our lives. I went to the Space Research Institute for a sabbatical in Moscow; just before that (in 1977), I had met all the great Russian physicists, who have now dispersed. But at that time, it was again — Moscow was then the center of plasma and nuclear research, so I just circulated around amongst those folks and had an extraordinary time. I wrote several papers with Sagdeev, of course, both at Trieste and elsewhere. But no, we stayed together.

Hearty:

Some of your other collaborators seem like — the ones that weren’t from UCLA, it seemed like you met them also either at Trieste or when you were going abroad. I’m thinking of Rene Pellat or —

Kennel:

Yeah. Rene Pellat was met first in Trieste, and he was part of the — you would consider him a Trieste alumnus. He (and Guy Laval) then went on to found a new group at École Polytechnique called the Centre de Physique Théorique — Centre for Theoretical Physics, and he invited me for a sabbatical in 1975. So I went. We wrote a couple of good papers together. Rene (had) also started off in fusion research but ended up in space research, just as Sagdeev did. Remember, Sagdeev started off in fusion, but by the time I knew him in later life, he was head of the Soviet space research institute. There had been this migration of people from fundamental — from the fusion program into the space program. Galeev was also a director (of the Soviet Space Research Institute). Pellat also headed the French national space agency for a while; he also headed the (French Atomic Energy Agency) and also the French NSF, CNRS. He was a master administrator, and so the French had him in each of its three big agencies,

Hearty:

Right. You also seem pretty well connected here in the States, too. I talked with Paul Kellogg, University of Minnesota.

Kennel:

Yeah.

Hearty:

Rene Pellat featured — he was pretty close to that program as well.

Kennel:

Yeah, Rene had a standing invitation to visit UCLA, and would just come and stay. Sometimes he stayed at our house. Other times they found an apartment for him. But he would come and retreat from the pressures of administrative life and spend a month or two just [laughs] hacking around at UCLA. He worked very closely with my good friend and colleague Ferd Coroniti from UCLA.

Hearty:

Any other lasting contacts from Trieste? I guess the entire community — most of them you stayed in touch with, right, because after that point —

Kennel:

Yeah. In 2014, we had a 75th birthday party for me at Scripps; it was a three-day party; we had 44 speakers, and they were — the different sessions were people from fields that I had worked in, so obviously, oceanography, atmospheric science, fusion plasma physics, space plasma physics, astrophysics. The president of the Royal Society came, Martin Rees. So we went through all of these different subjects that I had wandered into. But the fusion people were primarily people who were Trieste alumni that spoke. And the two that I remember particularly (were) Bruno Coppi ,from MIT, and Herbie Berk from the University of Texas. At that point, they (the fusion group at the party) started a debate — which, by the way, the whole thing is somehow recorded and televised, and you can find out what actually was said as opposed to what I remember)--But I do remember that there was a whole debate about the European (sic) ITER fusion project and the fact that its costs were running out of control. And they (the fusion scientists at the party) were informally estimating costs much larger than the official estimates — tens of billions of dollars. And a program so big and so clumsy and so hobbled by international — by the complexities of cooperative international work--- that it just wasn’t going to be a model for the fusion economy of the future. I still believe that, although it’s going to — it’s (ITER) designed like a high-energy physics machine, basically to turn out knowledge, but in fact what is probably more promising — some of the things now that, say, MIT are doing, where they use the knowledge, but are looking for small, compact devices, with big magnetic fields, much more automated control of the plasma than was ever possible in the early days, and things of that sort. So the goal of making a usable fusion machine is in that direction; ITER may, in fact, turn up some basic physics, which it probably will, especially if they’re burning plasma, but that the technology that goes into it — and all that you had to do to get that physics---, not much of it would be useful for power generation.

Hearty:

Right. Let’s see. Okay, so it’s been about two hours. Do you want to take another break, or you’re —

Kennel:

I’m happy if you are.

Hearty:

Okay. Yeah, let’s keep going. We’ve got through a lot of these. Maybe it’s time to return to UCLA, because that was over 30 years, right, that you were there.

Kennel:

Yes.

Hearty:

Right. I wanted to ask you about your research there. I know that’s a large question, but — and your teaching, so let’s start by asking you about — because I actually — not too much — I couldn’t find much, actually, on the UCLA Physics department during this time, so could you give us an impression of what the department was like when you arrived there?

Kennel:

It was a good, but not great, department. It aspired to the same kind of work that the very best people were doing. It didn’t want to do peripheral things. It didn’t want to pick some minor specialty and get well known in it. It wanted to be at the center of modern physics. Unfortunately, it couldn’t be, and UCLA has had — and might still have — an identity problem within the University of California, because there’s always Berkeley, and Lawrence Berkeley Laboratory, Lawrence Livermore Laboratory, and an overwhelming presence of Berkeley in the world of physics. This induced the kind of status anxiety that was palpable, and we considered our primary competitor [to be] Berkeley. Within the pecking order of physics, plasma physics was well down the line at that time. You could argue what people’s prejudices were, but in fact, the fact that we had a good physics department program in plasma physics, the fact that UCLA had a very good program in space plasma physics in the Department of Earth Sciences, and even some good stuff in the Department of Atmospheric Sciences, so that across the university, UCLA was pretty good, (despite all that) It (UCLA plasma research) wasn’t valued by the world of physics. What mattered was the experiments that you had at CERN. We brought (Nobel Laureate) Julian Schwinger to UCLA. He was a good friend of UCLA’s great chairman, Dave Saxon, who ultimately became president of the university. Dave had worked with Julian when they were young, and Julian Schwinger came to UCLA; but Julian came past the top of his career. So we weren’t the place where exciting fundamental things were happening in physics. And today — and you would not have believed, in my day, that things like atomic and molecular physics would get Nobel prizes. It just didn’t occur to me. It was particle physics. Nuclear. So we (UCLA plasma physics) had status (within our field), but plasma physics was well down the pecking order of the (UCLA Physics) department in terms of status as an academic enterprise. I was never completely satisfied there, even with my colleagues. I think the reason that I wasn’t (satisfied) was that they didn’t have a sense of what made absolute excellence in science. There’s a difference. I think that UC San Diego, now, has more than UCLA had then in that dimension. I certainly noticed it when I got to Scripps, which was the number one institution in ocean science. (While at UCLA) I was following in my mother’s footsteps — advice, internal advice. I was always serving on committees and trying to become socially involved with the field, so I was already serving on — I served on the Space Studies Board, and you could see my National Academy records (of participation in various NAS Committees) So I was doing all kinds of stuff, and I would come home from Washington and tell my colleagues, “You know, there’s this program here, and there’s that program (there), and there’s things that are beginning to happen. We ought to take advantage of that.” And the answer was, “Well, you know, we’re busy enough. We have enough grants. I don’t see why we should (always) rise to the newest thing.” And so I couldn’t — so they learned about new programs, government programs, after they were announced.

I’m a director at Scripps for about a week, and I get this phone call from the White House: “We’re thinking of developing a research initiative in this field ( I forget which), and we wondered if you’d host a little workshop while we figure out what to do.” So then I realized that the really number ones formulated the programs. They were there at the beginning, they had a sense for what was important in nascent fields, and they were willing to jump in. That was more attractive than one more grant (for the same field). So I felt at that time that my (UCLA) colleagues didn’t have that instinct. That’s also the thing I like about Cambridge, is they have that instinct. Remember, Cambridge usually finishes 1 to 4 in the global rankings of universities, and so they want to be there before it happens.

Hearty:

It’s different than following the research. Right? It’s more of an institutional innovation, in terms of —

Kennel:

One never knows exactly what is responsible for creativity, especially institutional creativity, but you know it when you see it, and it’s — different universities have different secrets. I mean, (for) Berkeley in physics (it) was just the overwhelming power of its tradition, its pulling power, together with the vast technical capacity that they built, just in the neighborhood: Lawrence Berkeley Laboratory. Cambridge, it’s the knowledge that Darwin — there’s that cup there — well, that’s Newton. But Darwin, Maxwell, Newton — that the people who invented subjects invented them there, including Fred Sanger with molecular biology — and Francis Crick (with DNA). There are various things that contribute to the organizational atmosphere, and they’re all different at different places. If there were a single way of doing it, nobody would become a leader. So I think that — but it’s something there about the spirit and the willingness to absolutely go into new fields and take a risk and be celebrated for the failures. It’s sort of the same way as in the Silicon Valley culture. You’re pretty well respected, even if you [laughs] had three companies that failed, provided you tried the new ideas. I think it was something like that, and UCLA was just burdened by feeling second in physics and having to always play catch-up. The other part about it — remember, I became vice-chancellor at UCLA, so I knew about its (institutional) role, its role is to be a public university in the largest, most polyglot — second largest and most polyglot city in the United States of America, with a tremendous — with representation of every social problem that the U.S. has, you'll find someplace in L.A., whether it is immigrants, social and economic inequality, pollution, all the urban problems. So it (UCLA) was burdened by its tremendous (diversity and size)— and it’s not private. It can’t be grasping, like USC. It has to be publicly oriented. This pull (of public problems) made it (UCLA) into one of the great comprehensive universities in the world, with stunning diversity of programs. You know, everything from library science and nursing to engineering, they have to have a little bit of everything to satisfy the needs of this giant city. And that ultimately put a weight on things, that they had to be comprehensively excellent, and they couldn’t really cater to singular points of excellence, and it didn’t happen; whereas at Berkeley, somehow they managed it somewhat better; certainly at Cambridge, that’s what they think about all the time.

Hearty:

Right.

Kennel:

And UCSD is sort of halfway between. It’s becoming UCLA. No university has probably grown faster, I think. Well, they probably have; we’re now at 40,000 students, and when I first arrived, we were at about 17, 18,000. A vast amount of building, billions of dollars’ worth of construction. The money is flowing. UCSD is able to use its reputation for creating multidisciplinary programs out of diverse interests, so that with the money, they’re actually prospering at the present time, but they will be weighted down by their enormous size, just as UCLA is. It’s hard to know what to do. We had a debate in the (UCSD) vice chancellor’s council that I kept my peace on. I didn’t enter it. But the question was to form a business school; UCSD’s Rady School of Management is a very good one, and they decided on a unique profile (in) that they would try to create business skills for scientific entrepreneurs, basically. To a large extent, I think they do do that, but the danger of going into business schools is that the whole community would then come down on you and want you to do all the conventional things excellently, and you would fall into the UCLA trap.

Hearty:

Right. That’s fascinating. Yeah. Taking a step back on that topic of UCLA, I’m interested in this Institute of Geophysics and Planetary Physics, which seems a bit different from what you’re describing — the problems of the Department of Physics.

Kennel:

Yeah.

Hearty:

Was this institute different in any way from what you’re saying?

Kennel:

Oh, yes. It was founded after the war; at that point, remember that the academic world was alive to the notion that science had made a tremendous change and difference to the national fate, and that basic science…the Vannevar Bush hypothesis that basic science was the goose that laid generations of golden eggs. So at the end of the war, Louis Slichter and several other people convinced the state that it had two things that were unique: California had an oil industry, which even now, believe it or not, is the fourth largest in the country — it had an oil industry, and it had earthquakes; it ought to be a leader in that (earthquakes). So they formed a statewide institute with a small number of FTEs, faculty members, were dedicated to advancing basic understanding of seismology, basically, and Earth Science more generally. And I remember the UCLA Institute, when it was first formed, before I joined it in the 1960s, had something like eight faculty members associated with it. The whole idea was they would be given full time off and just do pure research, no social responsibility — what we would consider today social responsibility, the social responsibility that the great researchers had was to do great research, and the rest of society would translate that into value, that you personally were not responsible for the translation into value, which I think is — well, it’s not the way we think today. I’m not sure [laughs] which generation is right, by the way. But in any case, we (that is, IGPP) had something like six or seven academy members out of the eight or nine faculty members, and ultimately the contrast between the “privileges” of the IGPP and the ordinary faculty were great, so finally people had fractional appointments in IGPP (which reduced individual privilege and passed out the largesse). And I arrived first as a full-time faculty member (in Physics), but then I was recruited to join IGPP half-time, which meant that I was relieved of half of my teaching responsibilities at the Physics department. In terms of teaching, I realized the privileges that I had been given by the IGPP, but also there’s this whole question of social responsibility. I spent a lot of time teaching Physics for Poets, Physics 10. (to compensate somewhat for my lightened teaching load in Physics)

Hearty:

How was the enrollment in that?

Kennel:

Pretty good.

Hearty:

A good interest?

Kennel:

Yeah. And at that point, that’s when I got into the history of science, because Harold Ticho , my dean, (who) subsequently came to UCSD, pointed out (to me) that the university had this great need to be relevant to the younger generation of students, and he said to me, “There’s all sorts of other inputs to your salary consideration, and I’m sure they will be important, but mine is going to be proportional to the enrollment in that course.” So with that, I got together with a Physics colleague, Ernie Abers, and we decided to try to teach physics through its history, through how our ideas about the world evolved in some of the great people, and to the extent that we could, some of the personal — historians now would say--- “anecdotes” about the creators of the physics of the time. We ultimately designed a course called “Mechanics: From Aristotle to Einstein,”; as part of that, we (Abers and I) discovered that there were some pretty well-known historians of science on the campus: Robert Westman, who came ultimately to UCSD; a gentleman named Norton Wise, who went (from UCLA) to Princeton; and probably the deepest one of them all, Amos Funkenstein. Amos died young; he had a joint appointment, I think, with a University in Israel as well. We had lunch-table conversations throughout two to three years (devoted to) the history of science; (we discussed) how much you could trust of professional (scientists’) anecdotes, and so on. Ernie and I wrote a bad book, (“Matter in Motion”) but we tried. So that was my teaching, and that was again, for its time, something new. Our approach (in our book could have) attracted a great deal of suspicion (from professional social scientists), because it could easily be interpreted as the whiggish interpretation of the history of science, that science advances through its great individuals. This (suspicion) I think, was the (historians’) ethos of the time. The ethos (i.e., that great individuals matter in science) founded the IGPP in its original form, and probably sped me along, and a whole bunch of other people, but that time has changed. And I was (becoming) aware of the changes at the time, that things were different. But I got to be — also convinced that it’s not that individuals don’t matter. They are critical. You see them all the time — the impact that CEOs or presidents can have on (the) psychological surroundings of all sorts of people. Of course, it is (people are) critical, but individuals develop their individuality growing up in a social matrix, and that also was important, so they (the historians of science) had to work all that out. That was basically the teaching that I enjoyed the most.

Later on (the dependence of science on its great figures changed); I will tell you an anecdote that I invented about how science has changed, going from the Vannevar Bush era to our present one. In 1994, I believe, there was the 50th anniversary of the Los Alamos National Laboratory, and we had a big celebration of — I guess Los Alamos was probably founded earlier than that, but there was something that happened. Oh, it was perhaps the first nuclear explosion. 1994. Anyhow, Sagdeev and I both went to the symposium, and there was a big, long discussion about the early days; that discussion, in many ways, was dominated not by the intellectual aspects of it, but by the emotional and psychological vision of Richard Feynman, and how they (Los Alamos) did the first nuclear implosions calculations. They had a large room with many women, who were then called “computers,” [laughs] working Marchant calculators. Richard Feynman was the ringmaster; the calculation was broken up in little pieces, and each person would do a part. If a machine broke down, or if a calculation went awry, Feynman knew not only the calculation, but how to repair the machines. He was ringmaster of this collection of people passing a complex calculation along; that was the first (nuclear implosion) code, or the first approaches (to one). There was a big discussion about Nick Metropolis and his contribution to the computational program. Later on (after the symposium), Sagdeev and I were invited by Scottish BBC to an interview; this woman (interviewer) asked, “You’ve seen so many things. How has science changed since that time?” I said, “Well, you know, when I first got into physics, the whole physics community, there were hundreds and thousands of physicists, but actually the whole progress of the field depended on the thoughts of eight or nine people. And one of them was Feynman. The rest of us, the rest of the physics community, were there to generate material for their thought processes. But now, things are very different.” She said, “Why?” And I said, “Well, actually this whole computational program — we’ve learned how to do all of that on a computer, but more than that, we’ve learned how to connect people together though the internet, and their computations together. So now, it is (at last) possible for a thousand people to work together to do more than Feynman could do alone. The government actually prefers that, because they can control those thousand, whereas they never could control Feynman.”

I think that science has changed, and that part of it has been — part of the move towards interdisciplinary work has been, first of all, made possible through the interconnectivity provided by the ’net, and all our new computational tools. But that is also, in some sense, awakened us. Let me put it this way. The internet and all these computation things bring forward the social implications of what we’re doing in a way that we were insulated from at earlier times. That insulation meant that all but the top eight physicists were shut out from the real action. (Now) They all can participate. You can work out the implications. If you have a new idea, it’s worked out and the implications are worked out almost instantly now, so it’s changed, and I think the social responsibility is a benefit, not a disbenefit. But I do think that what we lost was the idea that singular individuals, and the willingness to jump into things, has left academic physics. It’s still in Steve Jobs. Right? It’s still there. Steve Jobs had it. Elon Musk, God bless him, has it. There are a few people like that, but they’re not — they’re in the Silicon Valley culture now, and not in the culture of basic science. There’s a person that I worked with closely at Scripps (to lure him to Scripps), and that was Craig Venter; Venter is thoroughly hated by half of the biology community. But his entrepreneurial skills and his willingness to enter into new things and to see that biology was a big data subject — Craig Venter is one of those, and he would be amongst my pantheon of great people, with whom I’ve not worked, but have interacted with. I mean, Venter’s built a big laboratory on the Scripps campus. [laughs]

Hearty:

Right. This was the answer that you gave this interviewer, the one who asked you the question, the journalist at this 1994 —

Kennel:

That was — look, the answer’s been improved. [laughs]

Hearty:

Yeah, I was about to — this is Charlie talking in 2019.

Kennel:

Yeah, you’ve got to be very careful about that, and I know about all of that, and I just can’t help it. As you get older, you ruminate, and also knowledge is a compaction of things that you already knew.

Hearty:

Right.

Kennel:

And you get them down into algorithms, and then you remember the algorithm and not all the details. I understand the risks, and that any historian of science should go look at that damn interview. But this is what I was thinking, at least after the fact, that I gave a fairly good answer, I thought.

Hearty:

I was just also trying to bring it back to 1994, because I know this was a time of a lot of change for you as well. This was around when you were doing more NASA work. I think the mission to Earth started a little after that. We’re skipping over a lot of UCLA, I know, but we’re also running out of time for the morning, and I want to respect that.

Kennel:

I have a compacted anecdote about why I went to NASA.

Hearty:

Sure. Okay.

Kennel:

And it went something like this: in ’91, I was elected to the academy, and that’s almost everybody’s dream, except for Richard Feynman. [laughs] The department gave me a party, and a funny thing happened. After the party, I said to myself: well, now what? Then I said, “Well, maybe I’ll write a book. I don’t have to write papers anymore. I guess I’ve already got what (they could give) me. I don’t know what I said, I thought to myself — I started writing these books, and I started writing a book on basic plasma physics in magnetohydrodynamics. Dropped that in the end — dropped that in the middle and started writing another book which I finished, on space physics phenomenology. The whole experience of writing those books was actually rather disappointing, and the reason was that I was trying to clean up all the things that were left behind. I wanted everything to be perfect and to look good. [laughs] And at the end of the day, I realized, you know, you’re not doing anything. It was into this slough of despond that I get a phone call from Dan Goldin, the NASA administrator. This must be 1992, ’93. And he says, “Charlie, I want you to come to Washington.” And I said, “Well, you mean to run your astrophysics program,” because we — he said, “No, no, no, no. Earth science.” I’ve compacted this, you know. This happened over a period of weeks, but the compacted story is: “Why do you want me to do Earth science?” He said, “You’re going to have to come to Washington and talk to me about it.” I went to Washington and talked to him. He had an operation for a detached retina, and he had to live in the dark, minimum physical activity, for about three or four days, so we sat in the dark in his apartment in the Watergate complex, and we talked space physics and science and so forth. The basic message was: you’re a reputable scientist. You’re not an Earth scientist, but when you go before Congress, then as now, half the people won’t believe a word that you’re saying about climate change. And you have an independent reputation. I want you to make your decisions basically only on scientific excellence. I want you to work with the National Academy and make all the decisions about the mission to Planet Earth based on what is the best for climate — well, now we would say “climate research.” And so, for some reason, I did it. I did it. I tossed over all the grants that I had, gave them to other people, and went to Washington. Part of it was my social anxiety again. All my friends were institute directors, and I was still not an institute director. But the other reason had to do with my thesis advisor, and that is that I’ve been a long believer in small satellites for space research in particular, allowing of flexibility and so forth, and that I had been concerned that the gigantic programs would encase things in an unflexible mode. Sometimes it’s necessary, but we were looking at 10 and 15 years to construct accelerators and things of that sort. I’d been an advocate of small satellites, and it turns out that my thesis advisor, Ed Frieman, had chaired a committee for NASA that looked at the then-Earth Observing System. When it was first conceived, it was the largest scientific project ever conceived, I believe.

Hearty:

Bigger than Hubble?

Kennel:

Oh, yes. It was going to be $18 billion run-out, 1990 dollars, and it was conceived as part of a panic in the first Bush Administration about the potential dangers of climate change, and the idea was to launch two giant spacecraft every five years, so six of them — three of them — and $18 billion would be the run-out. And spacecraft were picked. The experimentalists were selected. And then reality set in. This was much too much money, basically. Other people wanted it. It was competing — it was eating other people’s lunches, and also the Congress were concerned about a political issue that they were sensitive to, and also some others. And that was that originally when NASA designs a program, it consults the whole community through the Space Studies Board, the national community, and it basically trolls for opinion from a wide range of people before it formulates its program objectives. And they can manipulate that. There’s all sorts of stories about that. But in the event they did that for designing the Earth observing system — but then they did a selection of the research teams, and then at that point, the research teams, which are a small subset of the whole community, have all the say about what will happen to the project in the future. And it was there that NASA — it’s a great organization — it was there that NASA could exert its influence over the programs, because many of the spacecraft that are built do not resemble the ones that were proposed. Things happen, designs evolve, and so forth, and all of that works through the existing teams that they’ve (NASA) assembled, which they have influence over. Some people would say undue; I’m not sure. But in any case, the problem was that about 100 scientists had control over an $18 billion budget, basically — technical control. And so a lot of people (other scientists, the Congress) were very upset about that, and this is why NASA — why Goldin wanted me to work with the National Academy of Sciences. But I had (also) to do it with an existing project team, and (what’s more) my thesis advisor (had) chaired the committee that undercut the primary rationale for why you had to have 24 instruments on giant spacecraft. You could put them out, spread the rest, and low the cost on smaller spacecraft. So my job was to take this existing program (with) the existing stated broad scientific objectives and rejigger it into a mode where it can be executed on smaller spacecraft. Much more diverse. So we finally ended up with a program, all (still) called the Earth Observing System, but [that] was implemented on something like 26 smaller spacecraft. And it was done as best we could by consultation with the broad community at each change in the requirements — any major change in the requirements and (that hopefully kept the essence of the) objectives of the mission. And so we finally — the one that we designed was the one that was finally built, roughly speaking, and it came in probably at $6 billion (1n 1996). [laughs] It’s still a lot of money, but not 18, and not the largest project ever conceived, and not one that was hobbled by all the bureaucratic — well, the other big problem was having 24 instruments in the spacecraft. You see, they’d (the instruments) never been flown together, so there were bound to be requirements conflicts between them as you tried to integrate them on the spacecraft, and it would cost a hell of a lot of money to rejigger the rest of the spacecraft to take care of the need of one small, sub-part of the program. The net result was that the potential for overruns in this already huge program was very large; we reduced the possible scope of requirements conflicts per spacecraft by putting them on smaller spacecraft that were conceptually easier to integrate, basically. Ultimately, it worked. The first round was built. The original sales pitch had had the first spacecraft being replaced after five years with this second set of identical spacecraft; You never built the same spacecraft twice; that was a myth. The other thing that we knew was that spacecraft live with degraded performance far longer than their engineering lifetimes. In the political arena, it is difficult to admit that you’re having degraded performance, but (the instrument) still performs. You can’t go ahead and propose a 15-year program, because they will ask you, “After 10 years, how will things work?” And we will say, “There [is] bound to be degraded performance, but we’ll still get most of it,” and that just doesn’t go. It doesn’t go with the political arena, but it does go in the technical arena, and so the spacecraft that they’ve built have lasted that long with degraded performance and now are beginning to be retired. The real question is whether they can do it again ---building a set of spacecraft with integrated goals across different spacecraft, in which the integration takes place at the data analysis level and not on the spacecraft itself.

That was what the Earth Observing System was about, and that, I think — and then at that point, [climate change] was again like the Civil Rights issue (for me). I knew it (climate change) was there; I never thought deeply about it, but I had to at NASA, I had to see the complexities of just trying to deal with it, and also the awful political environment in which it exists. I couldn’t get it out of my head. I just couldn’t. So I went back to UCLA and went to the executive vice chancellor job, but I discovered — well, that was a good job for somebody, but when the Scripps job came up — Scripps being the best research group outside of the government in the country in these fields (Climate change and earth science) — I took it. Pay cut. Move down (from LA to San Diego). Changed location, and so forth. Never regretted it. I think it was a great experience to be at Scripps again (Having visited before). Of course, my thesis advisor [laughs] oiled the — he opened up the way again. Ed Frieman. But I think that it was — I also realized that it would not be possible at age 50-odd to get into the subject at a level of newness, novelty, with the psychological receptivity that the young have, that you could never, ever — when you enter a field like that — get to the heart of it, the emotional heart of that field. You just couldn’t. You could make lots of contributions. I could use my social and administrative and other skills and so on, but the thought that I could make a change in the way these (great) people think and feel? Not possible, I thought. But I took it (the job), and I went anyhow and did it.

Hearty:

Yeah. Looking back at that switch — well, actually, let me ask you this. What was Scripps like? You mentioned the legacy of Ed Frieman at Scripps, right, that he had kind of set up a space for a director who had a physics background, although I think the director before him also had a physics background.

Kennel:

That’s correct.

Hearty:

But as you mentioned here, the political climate when you took over Scripps as director was different than what Ed had gone through.

Kennel:

Yes, but Ed prepared the way. Remember now, Ed was one of the — a fine physicist but probably even more important as an advisor to the government. He had been vice chairman of the president’s science advisory council during the Reagan administration. He had been director of energy research at DOE, and it was at the — and he had also been at Scripps and influenced by Roger Revelle, who’s actually the emotional godfather of the climate change issue, as Al Gore will tell you. Scripps saw, in a way that most people didn’t — oh, and the other thing was that physicists were, at that point, dominant in the science advisory apparatus. They had an influence that came from the Los Alamos years and the Vannevar Bush influence and so forth. They had an influence far greater than they have now, and it was thought that physicists could just about do anything. It took a physicist to make — that’s why physicists (Billl Nierenberg, Ed Frieman, and me) became directors of Scripps, by the way: their influence in the government, the Navy, and so forth. But he (Frieman)Ramanathan saw that the government — the science game was changing, and that the central issue for America was going to shift from defense to climate change. And in that, he was right. It should have shifted. It’s no longer defense. But in any case, he saw that the primary rationale for the funding of Scripps was going to change, and that while we always have a good relationship with the U.S. Navy in particular, we would be left behind if we did not get into atmospheric, environmental sciences, and these issues that we did have leverage over. The greatest thing he did was to recruit Ramanathan to Scripps. That was the background of my appointment.

Hearty:

Was there a tension there when you started, that — I think you referred this in your writing too, a bit, that these were — Scripps was filled with mostly experimentalists, oceanographic backgrounds. Right? Multidisciplinary. But was there a sense that, okay, here comes Charlie, yet another physicist, or was there —

Kennel:

There was a definite sense there. In fact, the subsequent directors were all card-carrying oceanographers. Part of it was because of my own attitudes. Well, I don’t think these are central to that change, but when I arrived, I arrived at an institution that was at the top of its game. Its game was the deployment of scientific equipment in the ocean, and the analysis of the data. And also, under Frieman and Bill Nierenberg beforehand, they had gone into space observations of the ocean in a big way. I figured that it was already number one in its field, but that I had just been vice chancellor of a comprehensive university (UCLA), and I knew about all sorts of other things that might be important. At that time, which was 1998, I identified two fields that I thought would expand its (Scripps’) capabilities in oceanography and the climate, and thought that I could at least help them get into these fields — not get into them, but to expand their hold on the use of these fields —the two were molecular biology and information technology. And at that time, we managed — oh, and the other thing was that as a result of some things (my previous work with NASA and relationship with Dan Goldin?), I was awarded — well, I wasn’t, Scripps was awarded a (NASA) grant for a — it could manufacture its own spacecraft, basically. This was defeated by the Republicans in Congress. This was the spacecraft that’s now called DSCOVR, and it had (initially) been defeated for ideological reasons. It was to go out to the L1 point ahead of Earth; from that point, it had the unique vantage point, looking back at Earth, of seeing more of the atmosphere than any other spacecraft. Especially when you orbit Earth (ie, try to assemble a complete set of data from Low-Earth orbit), you have to suture together the data, and you can’t (do so) uniformly — it takes a lot of work to get a full picture. And here (From L1), we had the ability to observe the whole sunlit portion of Earth, and it would give us a better way of quantifying the rate of global warming (in colloquial terms). The problem was that Al Gore had seen that (how DSCOVR views the earth) and said, “Oh, God. This is wonderful. This takes pictures of the whole Earth. Just imagine what it’s like if all these pictures are beamed to every classroom in the United States.” That did it in.

Hearty:

Political.

Kennel:

Yeah, so it was defeated in Congress and put on the shelf and kept in mothballs for about 15 years, and then subsequently relaunched. And even during the relaunch discussion, I was told that some of the (Congressional) committee members that were discussing the project (ie, the proposal to take DSCOVR out ot mothballs)... wanted all the Earth-observing instruments disabled. It (DSCOVR) also had solar wind instruments on the other side (facing the Sun), looking — it was a Janus-like spacecraft. Ultimately, it was launched, and it’s now — it was years ahead of its time, and now it’s a few years behind its time, but it’s the best we could do.

(Imagine what my life would have been like if Gore had been elected, and I had brought Scripps its first spacecraft project in my first year as Director.)

Anyhow, I felt that I could enrich Scripps by bringing in other subjects, so I spent a lot of time doing that. Ultimately, it (bringing new subjects to SIO) proved to be a problem (because it did not fit with) the central ethos of the subject (oceanography); it was conveyed to me by a — we had a very tough experimentalist named Russ Davis, who was interested in measuring the temperature and global circulation of the ocean. He was basically a physical oceanographer; this was the subject that Scripps had pioneered or had been on the leading edge of in the late ’50s and ’60s, when we explored the circulation pattern of the Pacific Ocean. Now, Russ Davis said, “I’m really worried about this institution. We’re losing our core. People care about physical oceanography, but only as background to their experiments. They need our data, but they don’t need our ideas.” I thought this was a gentle way of saying that I was emphasizing these new things too much and that the thing that had made the institution great in the era of Walter Munk and Roger Revelle was now hurting; I think that was (a) fundamental conflict. But at the same time, we did bring IT and molecular biology in. We did make some forays that were moderately successful in computational modeling that have been outmoded by the fact that one doesn’t do those things (big computing) anymore as an institution. It’s out there in the cloud. Besides, we had the Qualcomm Institute here on campus that does a lot of that. That (emphasizing modeling and big data) was, I think, the right direction, but what I missed was that we shouldn’t do those things alone.

Hearty:

We’re nearing the end of our time here, so I’m trying to think of a good closing question. I guess it would be fitting to close with a question about what you’re doing now at Cambridge.

Kennel:

When I stepped down as director (of SIO), I had a sabbatical coming. Everyone who’s smart will negotiate one after an administrator job. During that time (around when I stepped down in 2006), I had a friend that I didn’t know very well from Cambridge (Bill Fitzgerald), but I had always wanted to go there. So I asked him, “Do you have something available?” I realized (later) he pulled off an organizational miracle by getting me a visiting appointment to Christ College; I went for winter term in 2007. They (the College) had a little emergency that I managed to help them with. They had an annual C. P. Snow Lecture, a lecture (named for C.P. Snow) and the speaker that they had lined up bailed out. So I looked like a willing candidate. I had done a bunch of things, had been in government, and had been in science, so they asked me to do the C.P. Snow Lecture, which I’m very proud [of]. Then several years passed, but I came back (to Cambridge) from time to time, mostly to see people. But sometime around 2009 or ’10, I was in Cambridge for something else, and I ran into Frank Kelly, who was then the new master of Christ College. Frank said, “Charlie, you’re not using your privileges.” [laughs] I said, “Professors have privileges?” And he said, “Oh, yes, we’ve voted you a distinguished visiting scholar, and so anytime you’re in town, you can come, and you’ll get five meals a week at Christ College when you’re in town. Besides, if you need a place to stay and we have a place after we’ve met our other obligations, you get first dibs.” I slowly but then continuously started to take advantage of that. I started coming to Cambridge much more regularly, and then beginning around 2013, I kind of made it a regular practice to go in the winter quarter, so January through March. And I’ve done so continuously every year, including this coming year (2020), and I gradually began to build up relationships. The question was: who was I going to work with? There were two people who played an important role in my integration into the life of the university. The first was the person who invited me to Christ College, Bill Fitzgerald, who was one of the original big data gurus. He’s now passed on, but Bill was a college man, in the old terms, and he introduced me to all of the subtleties and what it takes to be an actual fellow at one of these old universities, all the subtle delicacies. But the other one was Martin Rees, who was my colleague from astrophysics days. I’d visited Martin at Cambridge when I had written a paper that he liked (and everybody likes, it turns out), on the Crab Nebula and the relativistic plasmas around the Crab Nebula. And Martin took care — this is in my first couple of years — to introduce me to people all around the university, and subsequently in England — invited me to various Royal Society events and so forth, and he made sure that I got to meet a whole variety of very good people, which I did. (When I first arrived, Martin was Astronomer Royal, Master of Trinity College, and President of the Royal Society, very well connected indeed. ) Now, the question was: what was I supposed to do there. Cambridge is a place where you can survive conversationally. What matters is the — so I got connected up with something called the Center for Science and Policy, whose job it is to connect the academics of Cambridge with the analytic people in the government agencies; that’s a very productive relationship, it goes very well, and I just participated in that. But soon I found myself talking to people in a whole variety of departments, including social anthropology; the theme at that time was again climate change, but social anthropology had a series of projects involving how indigenous cultures are responding to the changes of climate in the Arctic, and also to the changes in the climate in the Himalaya area. There’s a Center for a Middle Asia studies unit (¬Mongolia and Inner Asia Study Unit, MIASU) that I visited quite regularly. We began to discuss the wider implications of climate change, not just the modeling, but the impacts on things like these (indigenous societies). We got interested in riverine cultures: What will happen to the Ganges when the Himalayan snows run out? This has turned out to be a surprisingly productive relationship, and I actually have a paper in one of those, the Cambridge Journal of Anthropology, and of course I’ve discussed with Martin —The other person that I met along the way was Partha Dasgupta. Partha is one of the world’s great resource economists; his main interest in [his] professional life has always been that economics fails to take into account the environment in the appropriate way; we’re using up the environment while enriching ourselves. In the terms in vogue today, we’re spending natural capital and converting it, at an adverse rate, into financial capital. Partha also became a friend, and I began to see his point of view. Then, I think at that point, a seminal thing happened. The first Cambridge research that I did, Martin and I went to the Gordon and Betty Moore Foundation, and got a grant for a home-and-home series on the impacts of climate change on river water supplies — mountain snows and river water supplies. This had been a big problem in California. You know, the Sierra Nevada — an assessment done of climate change in California in the 2003 period or so, said that the Sierra Nevada snows were liable to disappear in this century. And of course, 70 percent or 80 percent of California’s water comes from the Sierra Nevada. So we had already had a significant effort at the state level, very sophisticated for government managers. Very sophisticated water managers. And the California Energy Commission had commissioned a number of research programs on the impacts of losing the mountain snows, forecasting the river runoff, and so forth. Cambridge had this program involving the Himalayas, so we decided on a home-and-home series; also Cambridge had on its staff the retired director of the U.K. Meteorological Agency, Lord Julian Hunt. Julian was interested in water resources in Africa. So we had a home-and-home program on the impacts of climate change on water resources in Africa, and ours — at UCSD, we invited the folks from the Himalayas (people who worked on the Himalayan snows) to attend. We (UCSD) invited people from Beijing (to hear_ how the Chinese were handling the threat — you remember the Yangtze has its headwaters in the Himalaya mountains, as does the Yellow River, so how they were handling it — we also had people from India, the Ganges, and of course we had the very sophisticated California water people there. (as well as Nepal). And that was a good (meeting). And then Julian cosponsored the one at Cambridge, which some of us went to, on the difficulties of managing water in Africa. Totally different. Much of it was (about the difficulty of) gathering data. But anyhow, we had this home-and-home. (There were follow-ons to this meeting) Ramanathan was sick (at the time of ours at SIO), but he knew about all of the water resource problems, and Ramanathan (later) sponsored — he was a member of the Pontifical Academy, and he was allowed (a number of) workshops that he can actually design. He designed one on the mountain snows and climate change for the Vatican Academy of Sciences in about 2011. I went to that. The important thing for the future was that during all of this, Ramanathan got to meet my then-new colleague, Partha Dasgupta. Partha was prominent in the Pontifical Academy of Social Sciences. They’re both on the Council of the Pontifical Academy; their story is that they got together one night at a good Italian restaurant. They got drunk on wine and decided how nice it would be to have a meeting in which real scientists, if you will, and real social scientists got together and actually discussed the implications of climate change. That led to a great series of workshops, of which I went to the first, at the Pontifical Academy. There was joined together in one place — for the first time for me — the social implications of climate change as well as the physical part. We had discussions of human trafficking and, you know, the social change stimulated by climate; also There (were a number of) Nobel Laureate economists there. It was quite a crew that Partha had gathered together; Ram had gathered together some of his friends, including me and Martin Rees. We had a debate. And this was a week-long seminar that the two (Ram and Partha) orchestrated basically. The week-long seminar actually went longer, because we had a weekend in between. But at the end of the day — and it was wonderful that each invited speaker had a whole morning. You couldn’t use up the whole morning talking about your work, but you could use up half of it talking, and then there would be a recorded debate from all the other people about the issues. So we went through all of that; that meeting was the one that provided some of the intellectual substance that underlies the Pope’s great encyclical on climate change, Laudato Si’. Ram from Scripps and Partha from Cambridge co-chaired that meeting. That’s been a very important relationship. The question was that we were trying to think out how you get through this climate crisis.

In about 2014 I gave a series of 10 lectures (at Christ’s College) — I think 10; I can’t remember — they were all recorded. They’re podcasts… on the impacts of climate change. Not why it is happening, not greenhouse gases, but what various regions, countries, what you can expect from river resources, urban heat islands, all of the various things about adaptation. A year later, people — it (was) a series of lectures for the general public; the lectures were nonspecialized. But I think a lot of experts turned up, too. They (the Christ’s Climate Lectures) were held from 5:30 to 7 on Thursdays (in Winter Term); the college would give a little wine reception, which assured that people would stay till the end. It went pretty well, but after the first year, I couldn’t give another 10 original lectures. [laughs] I used up my last 10 years of research on all these meetings on river resources and so forth. Then we (basically, Rob Doubleday and I) decided we’d invite other people (to speak), and as we did that, the whole emphasis changed, and for the better. The science was no longer the real issue. (In the U.S., it may still be, but certainly not there.) It wasn’t a question — unless you had something really new to say about how climate change would evolve, it wasn’t the real issue, and it wasn’t the real issue that a university like Cambridge were interested in. They wanted really to lead — remember, they think they lead the government — they wanted to lead a whole discussion about what to do. Gradually, we found that we got (good ) speakers. We got speakers from engineering, green technologies, the prospects of green power. That started off (a consideration of) green urban planning, and gradually it evolved toward the whole discussion of what to do, and ultimately what Cambridge should do. So about 2016, we chose the topic for the seminars (to be): what is the responsibility of universities in climate change? I kicked that off with a discussion of the really remarkable position that the University of California has taken, not only to commit to zero carbon behavior and green power; I think 2025 we’re (UC) supposed to be 100 percent green power. That was the simple part. The really difficult part was to use the capacity of the world’s largest research organization, if you include Lawrence Laboratory, Livermore, and Los Alamos and so on, and Lawrence Berkeley Laboratory, plus the research capacity of 10 campuses, including medicine and public health. What, then, should the university do, and how should it use its intellectual resources to deal with the solutions to the climate problem? They (UC) came up with a remarkable document. Ramanathan chaired a 50-person University committee, and they came up with a 10-pronged effort, of which (only) two really are science and technology, but many of them are going to be (important going forward) — one that my colleague, David Victor, was instrumental in — is basically to reform and develop new ways of promoting international collaboration. A deeper collaboration at all levels is required if we’re going to try to get through this problem. Finally, we understood that the full intellectual — it’s a challenge for all of society, and if that is so, then the full intellectual resources of the university are somehow to be mobilized in finding the solutions. That was the 2016 meeting: a series of seminars. At that point, I was just sort of avuncular; I would introduce the discussion, and then (after the lectures) I would lead a general discussion. I remember finally at the end of the ’16 one, I said, “It may be impertinent for an outsider to ask, but what is the right thing for Cambridge to do?” And that started off a very big discussion in the audience, which didn’t — it wasn’t decisive in what happened, but it was part of the atmosphere. Later on, Cambridge was in the process of selecting a new vice chancellor; remember, the present chancellor is Lord Sainsbury, a figurehead, but the vice chancellor is the true administrative leader. They were selecting a new one; my friends —Martin and Partha and others — were all involved in shaping the climate of opinion (I don’t know how that happens at Cambridge, actually). In any event, they (Cambridge) selected a gentleman from the University of British Columbia, which is one of the world’s leading universities in the environmental area. And they’ve (UBC) had had a green forest management program for a long time. Stephen Toope (UBC vice-chancellor) accepted the (Cambridge) job, but he had to delay his arrival; during this preparation period, he asked for a Cambridge-wide committee to advise him on what they should be doing on climate change. My Christ’s College colleague, Ian Leslie, who had been their pro-vice chancellor for research but was no longer, was on that committee. Margaret Leinen, present director of Scripps, and I contrived to invite Ian to campus to see what we at USCD were doing. We had had a very advanced program of (energy management) — we have our own microgrid. We’re 75 or 80 percent carbon-free electricity. The Cambridge campus estate management is pretty well advanced, but we also have Scripps and all sorts of other things happening. So Ian came, and he spent a week checking us out. Talked to David Victor, Ram, and various luminaries on campus, and then went back. In the event, Toope proclaimed at the beginning of last year that climate was going to be a primary focus of the University’s, all the resources. Now they’ve got an initiative called “Cambridge Zero,” which is to bring (Cambridge University) to zero carbon-emission operations, I think, by some date.

My essential point is that at that lecture series, I was able, for a while, to use what I had learned at UCSD and Scripps, which were leading Cambridge, and provoke a discussion (of what Cambridge should do). Now, just before you (Ryan) came, we (Rob Doubleday and I) had a discussion of this year’s lecture series. The (UNFCCC) climate negotiations for next year are going to be in Glasgow, which is in the United Kingdom. And they’re also the important ones, because they’re the five-year review of the great Paris Agreement, which the Pope’s encyclical helped to prompt. But while everybody in Paris believed that by 2020, the world would be able to increase its commitments to reduce CO2, in fact there’s been serious backsliding. That (Paris) was the high point. So this (UNFCCC COP) meeting will be an important meeting, to take stock and reformulate what needs to have happened. We (Rob Doubleday from Cambridge and I) thought that we could hitch (the lecture series to that Glasgow meeting) — we had reduced the lecture series to four, because Trinity College now gives a similar series, and we connect with them. But our four would be devoted to the question: what should England do, what should the U.K. do at Glasgow, and what should Cambridge do for the U.K. in thinking this through? So we’re getting some lecturers together, but in the meantime then, the thing that’s now got me kind of interested is my conversations with Martin Rees, Ramanathan, David Victor, and Partha Dasgupta, and that we thought — we really had to think out, once again, keeping going. It turned out that we first had a discussion with Partha and Martin — should the Pontifical Academy have another joint meeting of natural and social sciences? We had a long discussion about that, and finally we came up with some thoughts; this time we thought: maybe this is not — they’re not the right venue for the present time. So we’ve connected up with Peter Raven, who is a famous ecologist, and we’ve got a proposal going forward to the American Academy of Arts and Sciences, in particular their journal Daedalus; we’re proposing a whole series of contributed articles on what we would call “the Anthropocene crisis.” It goes beyond the climate crisis, but it has to do with the observations of the great ecologist E.O. Wilson. Wilson and Peter Raven and others have noted, of course, that the greatest environmental risks are being created just by population pressure — just activities of human beings. We’re using up the environment, and we’re not paying for the repair. Paying for the repair is Partha’s job. Right? Figure out how we do that. But the crisis comes because the demographers predict that the world’s population is going to peak around 2075 or so. We don’t quite know the level at which it would peak, and of course, that’s the key issue. But the facts of the matter are that climate variability and ecological vulnerability will coincide in the same era, the era in which the number of human beings at risk is the largest. If you go forward from — after that, the presumption that E.O. Wilson made was that we would learn how to deal with the environment much better, and besides, there would be fewer human beings, post the demographic maximum. So really, the whole crisis doesn’t last forever, but we’ve got to get through the next 50 or so years. That changes it. It changes it quite a bit, because it makes the problem look enormous, but solvable. The other question that we’ve been trying to deal with is: why is it that after 150 years of research and 40 years of negotiations, why is it that people will pay — even the most willing people in the world will pay lip service to dealing with the climate change and then come up inadequate response? And they’re leading us into a very dangerous territory. So we think that it’s not a question — people have tried political persuasion. People have written books. There have been all sorts of scientific assessments. Every scientist in the world, except maybe four are — [laughs] I exaggerate---. But the uniformity of scientific opinion around the world is there. You would think that finally you would have moved society into recognizing the danger and doing something at the scale that’s required. The question is: why aren’t we doing it? Ultimately, I think it resides deep in human psychology, that we don’t really deal well with long-term goals. We need short-term rewards. At the same time, it’s absolutely clear that if you are going to transform the environment to get through this crisis — we’re in the Anthropocene already. We’ve already changed the environment. The things that we’re doing are heading (the planet) towards a major crisis. We have the capacity to change it, but we have to change ourselves first. We have to have the capacity to change the direction, but we have to change ourselves and our societies first. People on the academic side are thinking very carefully about how to go to low-carbon economies. The transition to zero-carbon economies is a very profound one. It requires all sorts of adjustments and changes, in obviously technology, but finance, attitudes, economic attitudes, and so forth. So society needs to change along with the change in the environment; if society doesn’t change, we will continue to get the problems that we currently have, because the two (society and the environment) are tightly coupled now. We’ve got this project going with American Academy to see if we can’t get a whole series of commitments (for articles) from them on the crisis of the Anthropocene. (The Academy eventually refused the project.)

Hearty:

Very interesting.

Kennel:

But the point is, it’s sort of a new way of assembling all sorts of old thoughts (the concept of the Anthropocene Crisis), but put them together, and you get a new way of thinking about it.

Hearty:

Yeah.

Kennel:

It (the ecological and climatic crisis) isn’t an eternal problem, but it may defeat us anyhow. But it needn’t. There is an escape hatch. The other thing I think that’s important is that I think the rise of all the social media and the power of big data will play an instrumental role in creating a planet that human beings manage. I guess we have to.

Hearty:

Right.

Kennel:

We don’t have any choice anymore. Careless management will land us up in the dump.

Hearty:

Right. Well, thank you so much, Charlie. This has been great.

Kennel:

Yeah. Well, thank you.

[END]