Rainer Weiss

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
Image of Rainer Weiss

Credit: Matt Weber

Interviewed by
David Zierler
Interview dates
June 7, 14, 21 & 28, 2020
Location
Video conference
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 Rainer Weiss by David Zierler on June 7, 14, 21 & 28, 2020,
Niels Bohr Library & Archives, American Institute of Physics,
College Park, MD USA,
www.aip.org/history-programs/niels-bohr-library/oral-histories/47250

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

Abstract

Interview with Rainer Weiss, professor emeritus of physics at MIT. Weiss recounts his family history in pre-war Europe and the circumstances of his parents' marriage. He describes his childhood in New York City, and he explains his interests in experimenting and tinkering from an early age. Weiss explains the circumstances leading to his undergraduate study at MIT and his original plan to study electrical engineering before focusing on physics. He recounts his long and deep relationship with Jerrold Zacharias, who singularly championed Weiss's interests over the years. He discusses his graduate work on the hyperfine structure of hydrogen fluoride. Weiss describes his formative work with Bob Dicke at Princeton, and he explains how technological advances was beginning to offer new advances in general relativity. He explains how Dicke's influence served as an intellectual underpinning for the creation and success of LIGO. Weiss emphasizes the importance of Richard Isaacson as one of the founding heroes of LIGO, and he describes the fundamental importance of joining his research institutionally with Caltech. He describes his early research with John Mather, and the numerous administrative challenges in working with the NSF throughout the LIGO endeavor. Weiss describes the geographical decisions that went into building LIGO, the various episodes when LIGO's ongoing viability was in doubt, and how both Barry Barish and Kip Thorne contributed to ensuring its success. At the end of the interview, Weiss describes some of the sensitivities regarding who has been recognized in LIGO and who has not, in light of all the attention conferred by the Nobel prize, and he reflects on how LIGO will continue to push discoveries forward on the nature and origins of the universe.

Transcript

[Begin 200607_0142_D]

Zierler:

This is David Zierler, oral historian for the American Institute of Physics. It is June 7th, 2020. It is my great pleasure and privilege to be here with Professor Rainer Weiss. Rai, thank you so much for being with me today.

Weiss:

You're quite welcome.

Zierler:

To start, please tell me your title and institutional affiliation.

Weiss:

Well, I'm now an emeritus professor at MIT in Cambridge, Mass.

Zierler:

Now, I see LIGO is in your email. Is that part of an institutional affiliation as well?

Weiss:

That? No. It’s a project. I think that’s the best way to say it. It’s a project that Caltech and MIT are doing together. And now there is a huge collaboration of about a thousand people that have been drawn into that. But the thing is managed by Caltech, and MIT is a subcontractor. That’s a long story, how that happened.

Zierler:

Hopefully, we'll get to that. But for now—

Weiss:

I don’t know if we will! (laughter)

Zierler:

(Laughter) Let’s go right to the very beginning. First, it’s a very interesting story—tell me a little bit about your parents and where they came from.

Weiss:

Well, we're immigrants. My father was a doctor and a communist, a very persuasive communist, in Germany. He came from a family of very rich Jews- it turns out the Rathenau family. I don’t know if you know them, but they owned the Allgemeine Elektrizitäts-Gesellschaft, for example. That’s like the GE of Germany. And one of the Rathenaus was the foreign secretary during the Weimar Republic, and he got assassinated by Prussians. So, the family is rich. And my father came from a what I would call coupon-clipping environment. And he rebelled against it and became a Communist. He became a doctor and he worked in a communist hospital in Berlin. He met my mother, who was not Jewish and nowhere near the intellectual that he was, but she was quite attractive, I guess. On top of that, she was an actress. And she was on the Berlin stage and they got together. And I think on a New Year’s Eve, they conceived me, okay? (laughter) If you calculate back. But they weren’t married. They got married a couple years later. I was a bastard, okay? And what got him into trouble was that in his hospital, where he was the- he was a neurologist, by the way; that’s what his main profession was- there was a Nazi implant. And you have to understand a little bit about Berlin in the middle twenties; it was Balkanized. See, there was the Weimar Republic, but it wasn’t powerful enough and there were enclaves all over the city that were communist. And there were Weimar, and then there were people who were actually fascist, already. They weren’t quite Nazis yet, but they were fascists.

So, he happened to spot a Nazi doing a very bad operation in his own hospital, and then went to what was left of the Weimar to get that guy arrested. And then the Nazis took him prisoner, even though they were not yet in power. In other words, as he walked through an enclave where the Nazis were in control, they grabbed him. And the only reason he got out was because my mother’s family was still tied up in the government somehow. Okay? They were not activists, but they were in the Weimar, and they somehow got him out. And they shoved him across- the story is my mother and that family shoved him across the border into Czechoslovakia. So, he was the first to leave, and then after I was born, she followed. And I think they eventually got married.

And then he became a- he wasn’t a doctor in Czechoslovakia, because he would have had to go through the same kind of thing you do in the United States. You have to take the boards and show that you're a good doctor. And he didn't want to do that. So, he worked for a pharmaceutical company. And my memory of that was two of them. He wanted to go to Russia in ’36. My mother adamantly didn't want to do that. Because the Purge Trials had started. It was very clear to anybody that anybody who was an idealist communist would get killed in Russia if they went there. In other words, communism was a very- an idealist was the very last thing any communist in Russia wanted to see. And many of the Germans who went there, and a lot of the people from France and stuff, they wound up in Siberia.

Zierler:

So, your father was more attracted to the ideal of communism?

Weiss:

That was all it was. All communists. Every communist that you run into, in the United States as well, it was all the matter that that was a way to finally tame down capitalism so it would become sensible.

Zierler:

So, in 1936, he’s not looking to escape Czechoslovakia as a Jew? He’s not concerned about this?

Weiss:

No, absolutely not. No, no, no. That was not even on his mind. He wanted to go there and do good, and it was already too late to do good.

Zierler:

Was he aware of the truth of the communist regime?

Weiss:

I can’t tell you that. She was more emotionally aware of things, but not intellectually. But it was already true that people who had gone had not come back. That was the big thing. And she sensed that there was something wrong, and he didn't imagine that there could be anything wrong. And so that was one thing.

And then in 1938, he had gotten- I had a sister. In 1937, my sister was born, who by the way was a professor at NYU, in the Tisch School, I think. She’s retired now. What happened is that I remember as a boy, this is a vivid memory. I mean, this is when I- they went to the Tatra Mountains, which is a bunch of mountains in Slovakia. And there was a- a lot of Jews had gone there for vacation. And they were all- everybody was sitting in the foyer of the hotel, and there was a sort of gothic looking radio. You know those radios that looked like cathedrals? I don’t know if you remember. They're the plywood radios, and in the back, of course, are the tubes. There were tubes in there, and they were glowing. I mean, for a little boy, that was really very exciting, to look in that radio from the back. And of course, you don’t dare touch anything, but I could imagine all that stuff going on in there, and this voice coming out the front. And what was coming out of the front was Neville Chamberlain, this is September of ’38, giving the Sudetten Gebeit to Hitler. And that that caused a hell of a stir, within about four hours. Everybody raced to Prague to get the hell out there, because they knew that was the end. Including my parents. Then, the Jewish part came up.

Zierler:

Rai, was there any strain between your parents about the fact that—?

Weiss:

Oh, plenty. Plenty. That was a terrible marriage. And it never got any better.

Zierler:

I mean, did your mom ever think to your father, it’s like, “Well, you're in trouble, but I'm not”? Was that part of it?

Weiss:

I don’t think that—no, that was not the issue. He was, in the deepest sense—even though he was a communist, he was a German oligarch. I don’t know if you understand what I mean by that. I mean, his house was his, and democracy ended at the door. And that was part of the problem. And they wanted to get divorced, and I remember endlessly as a boy when we got to the United States, him complaining, even at the dinner table—“We wouldn't be married if it wasn’t for Hitler, the goddamn son of a bitch!” They just couldn't manage. Okay, but anyway! They lived together for sixty years. Okay? (laughter) But it was a very uncomfortable place. Anyway, so to get back to that thing, what came of it was that the people who got us out of Czechoslovakia and by the way, you think immigration problems are bad now, they were terrible then, also, at least vis a vis the United States and the rest of the world. It was mostly a Jewish problem. Nobody wanted Jews. But on top of that, you didn't want anybody from that part of Europe that had that smell of the East. In other words, the limits were placed on Italians, Russians, Poles, but there were not so many rigid limits on Germans. There may have been a limit on French, but the Germans and the British were favored.

Zierler:

So, your family’s German background wasn’t helpful in this regard?

Weiss:

Not really because- well, we'll get to it. What managed finally to get them out, and all of us out, was a family in Saint Louis. I have mentioned that in many interviews, and I think it’s amazing. Most people think of the Jews only coming to New York when they came to the United States. That’s not true. There were very many of them not, of course not obviously as many as went to New York, went to the middle of the country, especially along the Mississippi River, because you could do business on the Mississippi. And New Orleans was really a fantastic place. But Saint Louis was almost as fantastic.

So, there’s a very large group of Jews that had started businesses back in the middle 1800s, maybe, and many of them were doing things like Wanamaker or Macy’s, the same kind of big department store. And what happened is this woman, whose name is Stix—S-T-I-X—I've forgotten her first name. If it becomes important, I can tell you about it. I’d have to look it up. But anyway, she had taken on the idea that she would sign up and give bond effectively for 10,000 Jews who, as long as they had professions, now, she had conditions. Conditions. And she was going to give bond, meaning that if these people got into trouble when they got to the United States, she would take care of whatever might be the welfare payments. That’s a big deal, I thought. And my father was a doctor, and he got some money out of that, and we got out. We got out just in the nick of time.

Zierler:

Was your father ever able to get control of his family’s wealth, or he lost all of that?

Weiss:

No, he lost every bit of it. And yeah, that was- my aunt tried to do that. You know, trying to get the money from the Rathenau side. Nobody succeeded in doing that. The Germans used that up right away. On top of that, I don’t know the real reason. I certainly know my father didn't try it. He felt pissed off enough. And I think his sister Hilda tried, and she got something, but we never know exactly how much. You're right; reparations, it was an attempt at reparations, but not by him. So that got us in the United States, which was an enormous thing. And then, well, I don’t know how much more you want to know.

Zierler:

I want to know all of it. This is what we're here for.

Weiss:

Oh, god (laughter). All right! Well anyway, there’s not that much more. He eventually became a doctor in New York City, but not with a license, so he had to work as a doctor for some other doctor. And what he did eventually, he wound up in the Triboro Hospital in Queens. You come out of New York, right?

Zierler:

Yeah, yeah. How did they get from Saint- wait, did they go to Saint Louis, or they went to New York?

Weiss:

No, no, no. They came to New York. The lady in Saint Louis gave the bond, the affidavit.

Zierler:

So, there was no personal connection with her.

Weiss:

They never got to even see each other. Now, many, many years later, I went out there to thank her.

Zierler:

Oh, wow.

Weiss:

This was when I became an adult. In fact, many people in that family, the Stix family, don’t realize that Mrs. Stix did this.

Zierler:

Oh, wow.

Weiss:

When I talk to people out there—there’s two pieces or branches of the family, one in Cincinnati, and another in Saint Louis. And I think the name of the store was Baer, Fuller, & Stix. That was the big department store. And the people in the Stix family didn't really realize she was using- this was pocket change for her, and she was doing this for her own feeling about, “What the hell can I do to help this out?”

Anyway, let’s get back to New York. And what happened is that he, my father, then, never really got a—he had terrible trouble, as many Europeans had, with the American system of education. In other words, the idea of taking a short-answer exam, like you and I have probably done many, many times, you know, you get a question, then you get A, B, C, D, E, F, G, and you pick out the one that’s right, it doesn't work for medicine, at least for him it didn't. And the questions were like, “This patient has come into the office, has these symptoms, and then what is the illness?” Well of course, to him, he would write essays on the exam saying, “These are impossible questions. You don’t give me enough information. It could be this one or that one. You can’t answer this question the way you've posed it.” And so, he flunked out. He flunked it eight times. I mean, the family got very poor, and my mother went to work at various cafeterias. She did house cleaning for people.

Zierler:

Rai, what was your earliest memory of New York, when you got there?

Weiss:

Oh, that I remember. The first one was, yeah, the Statue of Liberty was a big deal. I remember that. But the next big thing was living on 86th Street near Broadway. But it wasn’t so much—I think the important part was that I was old enough—I had gone to school in Czechoslovakia some, and I was old enough—I think I was ready for second or third grade. See, I got here when I was—’39, ’40, so I was eight years old, so that’s third grade, something like that? Maybe third or second grade. Anyway, so I went to public schools. And the thing I remember vividly, my mother didn't understand how to- you have to remember, 1940 was about the time the Germans were running rampant over Europe at that time. In fact, they had taken over Holland. They had bombed Rotterdam. I remember going to school in lederhosen. I don’t know if you know what those are.

Zierler:

Sure, yeah.

Weiss:

You know, my mother sent me in that, and the teacher sort of effectively spit at me, because they had just bombed- the goddamn Germans had just bombed a city she was so much in love with. And that they were rampant, and doing all this terrible stuff. And the first German she could find, she tried to take it out on him.

Zierler:

Rai, I know you're a little kid, and you're probably not thinking in global terms, but—

Weiss:

No, I wasn’t. Not at all. I didn't know what was going on.

Zierler:

In terms of your identity, did you think of yourself as a refugee?

Weiss:

I don’t know that. I don’t remember that much. But I do remember the following, and there were a couple of nice stories in that, which you'll enjoy. But what happened is that, little by little, my father eventually did get through it, the exams, and he decided to go into psychoanalysis as a profession rather than neurology. It turns out you need an MD to do that. And so he started having patients. And by that time, I knew English better than he did, and of course I had made a lot of friends, Jewish friends, all over New York, in my classes. And so he, being a German Jew, he didn't understand half the stuff that most Jews in New York worried about. In other words, most of the German Jews, especially the intellectuals, want to be Germans, rather than Jews. I don’t know if you're aware of that.

Zierler:

Of course, of course.

Weiss:

Oh, you are. Okay. And so, he didn't know anything about the religion. I mean, he didn't know what a Bar Mitzvah was. He didn't know anything about all the problems of trying to marry outside of the religion, stuff like that. To him, that was complete nonsense. And so, he couldn't understand a lot of the problems that he was dealing with in his patients, and so he’d come to me a lot. You know, he’d tell me about this story about a Bar Mitzvah, that—he didn't understand what—what goes on at a Bar Mitzvah? So, I’d explain to him. And then, you know, what’s a bris? And then, what are all these different things? So, I knew some of that, and I knew more than he did. I don’t know how he survived with his patients, being a German Jew.

Anyway, eventually he survived. I don’t know. And then, okay, my mother then didn't have to work anymore, and she was then more involved with trying to run her family. She got me into a private school. We were living on the West Side in New York. We were living at about 107th Street and Central Park West, which in the forties, was still—much of Harlem—it was not white and lily white. So, what happened is I went to a school that was called the Columbia Grammar School. I don’t know if you know of that school. It’s on the West Side in New York. Turned out to be a very fancy private school. I had no idea. And I got a scholarship to it. And so there I was, being educated reasonably well. And I think that’s sort of where I began to get interested in- if you're interested in me, was where I began to get interested in science. I'm trying to think if there’s any other major event in that epoch that might be valuable to you. No, I don’t think there is.

Zierler:

In terms of developing the story of you as an inventor and a tinkerer, is it happening right around this time, as a young boy?

Weiss:

Yeah, it’s starting, then. It started just about when I was—seventh grade, maybe—in that school. I cannot tell you where the initial idea came from, but I was very much involved with reading Popular Mechanics, Popular Science, that sort of stuff. And electronics was absolutely fascinating, and I'll tell you why. And it got bigger and better as you went on. At the end of the war—and this is I’d say by the time ’45 had come along—the New York City streets, down and around the region where the World Trade Center was—there was a street called Cortlandt Street. I don’t know if you're familiar with that.

Zierler:

Sure, yeah.

Weiss:

And there, the sidewalks were loaded with junk. I mean, all this war surplus—tubes, vacuum tubes, capacitors, transformers, you name it. And so you could go down there—you could even buy a whole radar set for a buck, if you wanted. You could get an oscilloscope. If you bought the radar set, took the whole thing away, they’d give you—but, “Make sure to take all of it away, you know? Just don’t leave us anything” (laughter). And so, I had a room that was just full of stuff like that. And I started fixing radios for people. And one of the stories which I think you might enjoy, the one I was thinking about is that New York at that time in ’45, ’44, was still very much screwed up in their electrical system. I don’t know if you know this.

Zierler:

No.

Weiss:

The East Side was DC, and it had been that way because it was put there by ConEd, Edison’s company. He didn't believe in AC, by the way. And Tesla and the Westinghouse people were on the West Side, all the newer buildings (laughter). So, there was AC, sixty-hertz AC on the West Side. And as soon as the war was sort of over, people began to buy radios. These things became available. And what would happen is that people didn't know enough, and the plugs were so much the same—they were a little different, but people would buy a radio and then it wouldn't quite go into the wall, so they’d jam it in the wall, you know? And what happened is a guy bought an AC set—you know, a little Emerson radio, and take it to the—and on his Fifth Avenue apartment, the thing would go boom! The transformer would blow up, and then he had lost the radio, and everything like that.

And the other people, people on the other side, didn't get in so much trouble. You know, the people buying an old radio which was DC and putting it on AC, it’s not so bad. They didn't blow up stuff. They blew capacitors up, but not transformers. So anyway, what went on is I had a hell of a business going. I was in with these rich kids who were in my school, and their parents had bought a radio, and I told them I could fix it for them. So, what happened is I—

Zierler:

And were you self-taught? Did your dad help you out?

Weiss:

No, no. He knew nothing of this. I know the question is, “How did you know? How did you begin?” I don’t know the answer to that. The only answer I have is that it was these things I was reading Popular Science and there were things in there about electronics. And that seemed fascinating. So, a big thing was fixing these radios. I would carry radios under my arm coming out of the subway. And there were a lot of gangs around where we were living, up on 107th Street there, and they'd try to grab the goddamn radio from me! And that was a little—I was not very big—and sometimes they’d get it. And I would tell them, “Look, that doesn't belong to me. It belongs to somebody. I'm trying to fix it.” So, they got the idea that, well, all right—they'll go around, they'll steal radios, and I'll fix them for them. That was that. I didn’t realize that, but that was the strategy. And so they left me alone, and every once in a while, they’d slip me a radio to fix (laughter). And most of the stuff was sort of nonsense, like blown-up transformers or blown up electrolytic capacitors, because they had screwed up the wiring problems. Anyway, so that got me started. The thing that really got me started was something different, which is that I've always liked music, classical music. That I think comes from my parents.

Zierler:

You played the piano?

Weiss:

Well, no, I never got to, until I was an adult, and I wish I had started earlier. That’s to me one of the saddest things as a kid; I never learned. Because I don’t know if you know anything about playing any instrument, but—have you got your kids playing stuff, or not?

Zierler:

Yeah, I'm a musician. My kids play. Yeah, sure.

Weiss:

Well, when did you start?

Zierler:

When I was seven. I played cello when I was seven, because I wanted to play the bass, but you're too small to play the double bass as a seven-year-old, so they started me off on cello.

Weiss:

So, you played the cello? (laughter) And not a full-sized cello, either, right?

Zierler:

No, no, no. A little cello. Two years, I moved up to the bass. I've been playing classical bass ever since.

Weiss:

Oh, really!

Zierler:

Yeah.

Weiss:

And your kids? Anybody take the piano?

Zierler:

They take piano lessons now.

Weiss:

Okay, good. How old are they?

Zierler:

Four, six, eight, and ten.

Weiss:

Oh, wow. Good for you! Good for you!

Zierler:

Something like that!

Weiss:

Not the many kids, but I'm talking about that you started them early enough.

Zierler:

Right (laughter).

Weiss:

They will not have all the problems that I had.

Zierler:

Well, I don’t know; you turned out okay (laughter).

Weiss:

No, you don’t know. You never heard me play. I wouldn't even play for you. Okay? (laughter) But anyway, but the thing that I could do is—you have to remember the time. The time was, as I say, ’47 through ’50, about. That epoch, there was big things going on in the music business. For example, FM radio—frequency modulated radio was just coming in as a- it had been invented much earlier than that, but it was just coming in. And if you had a decent radio, you could pick up the New York Philharmonic. I mean, you could hear Bruno Walter, for example. Or you could hear the NBC Symphony, which was a—Toscanini was running that one. And you could hear a really first-class concert, full dynamic range, full frequency range, if you could get an FM radio. And very few people had a thing like that.

And so I started building FM radios. And I had a very good piece of luck. Two things. The British had started making a very good amplifier with vacuum tubes called the Williamson circuit. Since you know a little technical stuff, I'll just tell you what it is. I won’t spend a lot of time. It was a thing which allowed you to put feedback around the output transformer, which made it so that when a drum—when you plucked a cello, and you plucked it, let’s say—you heard actually the whole pluck, not a fuzzed-out version of the pluck. You heard the full impact. And a drum, it could shake you. It was really wonderful. It’s not just loud. The full frequency—the transient response was unbelievable. And that’s very important. It’s a thing people don’t often think about. So, the amplifier—I built these Williamson amplifiers. So not my design. I got the transformers from a friend of mine who bought them in England for me. I don’t know if you know—you say you're a New Yorker. Did you ever go to the Harvey Radio Company on 43rd Street?

Zierler:

No.

Weiss:

Never mind, okay. All right. Well that bunch of guys, they were a very good bunch of guys in there, and they kept me full of good stuff. They were really on top of this. And so they sold me the transformers; I built the rest of it. And then the big find was a problem in Brookaylyn. It turns out—I don’t know if you know that—not that you should know this, but there was a fire in one of the Brookaylyn theatres. I don’t know if it was the Paramount Theatre, somewhere near Borough Hall. Big fire behind the screen. And I got wind of it from these guys at Harvey Radio. So I raced down there to see what was going on, and I found out yeah, there had been a hell of a fire. It burnt out all the cones of the speakers, and they were trying to take the speakers down behind the screen. You know, behind the screen, there must have been a matrix of speakers. And I would climb ladders, and I got about seven of them down. They weighed much more than I could handle. But okay, a guy would help me out. But I got about seven of them off the burnt-out cones. But these were movie house loudspeakers. They were fourteen inches in diameter for the woofer, and then a small little hexagonal cell tweeter for the high frequencies. It was a very modern, at that time, loudspeaker. And I was able to get cones for them, and so I put together a system which consisted of one of these movie speakers, in my room, the Williamson amplifier, and an FM set. And I would bring in friends of my parents who were all interested in music. Most Europeans love classical music.

Zierler:

And your parents’ social circle were Europeans, mostly?

Weiss:

Oh, yeah, they didn't have—exactly that. You got it. And so, they would come—well, my mother is the one who instigated this. And so, what happened is that I’d sit them down in my room, and we’d listen to Bruno Walter, or we’d listen to Toscanini, one or the other. And they were absolutely—it sounded like they were right in the concert hall. They couldn't believe it. Neither could I; it was really spectacular. And so everyone who came wanted to have this, too. So, I got a whole bunch of orders for making these things from friends of the family. And I had a real business going. Not that I wanted to make money, but I wanted to give them what I had. They were willing to pay for the parts. But anyway, so that became a—I was already, by that time, in high school. It had grown to that.

Zierler:

And Rai, you stayed in the same school for high school?

Weiss:

Yeah. Right to the end, yeah.

Zierler:

Did you consider one of the technical schools like a Bronx School of Science or Stuyvesant?

Weiss:

No, never. Never. I don’t think I would have, because I really didn't have to pay anything. I did a lot of stuff for the school. I built them a PA system for their gymnasium, and I started a ham radio club for them. So, I had a vested interest in that place.

Zierler:

And in terms of the education, they were advanced enough to keep you busy? You didn't feel like you were bored in the science classes there?

Weiss:

Well, I was never a good student. I'm going to tell you that, right off. I was a crappy student. And you'll see that in a minute.

Zierler:

So, you're no science prodigy in high school.

Weiss:

No. No, I was a technical- something. Maybe a technician who knew a lot. But no, no, there was no big anything. Science was not the big deal, okay, and nor was any of the other stuff. Mathematics was not a big deal either, because I wasn’t very good at that. I was not a good student. I certainly didn't give a damn about what grades I got. Just there were too many projects to do, too much interesting stuff. Look, I don’t know—were you a kid in New York?

Zierler:

Upstate New York.

Weiss:

Oh, upstate. Okay.

Zierler:

But my parents—I recognize—I understand.

Weiss:

Well, no, but you see, there was too much to do in New York.

Zierler:

Of course!

Weiss:

School is unimportant. There was just all this stuff. I mean, patents to look up in the New York Public Library, and stuff like that.

Zierler:

And your parents were never on you to get better grades?

Weiss:

They didn't really quite understand the whole system. They already thought it was awful that it was different than the German system of gymnasium, okay? So, nobody was pushing me one way or the other.

Zierler:

Rai, I want to ask you another question about your dad. Especially during McCarthyism, did he tamp down—?

Weiss:

Oh, man, was that a problem!

Zierler:

He didn't tamp down his ideology, ever?

Weiss:

Well, okay, I can tell you about that. I'll come back to this other thing in a minute. Yeah, no, that was a big problem. In fact, by the time McCarthy came around, his practice, his psychoanalytic practice, had become quite successful. Especially at the height of the McCarthy era. And two of his biggest worries were that some of the people who were in his group in Germany, in his communist cell—I'll call it the cell, but they didn't call it that, in those days—the group around him who were the activists—had managed to get out of Germany, and some of them were—two of them were at Columbia University. And they never talked to each other. Why? Because they were worried of one exposing the other.

All of them had lied. I'm sure my father lied. I can only—when he was asked at the immigration, “Were you a communist?” he probably said no. Although if they had chosen to look, they would have found him in pictures with Lenin. I mean, it was really sort of outrageous. So, it turns out in those days, it wasn’t like it is now, where you have instant [communication]—you could get away with lying, in a way, without getting caught too quickly, is my guess. And a lot of these people did that.

And so one of the guys—there was a guy in Chinese history who was at Columbia, who had been close to him in Germany. And there was this terrible worry that, especially since he was at a university like Columbia, which was not very good to its people, by the way, during the McCarthy—not many universities could stand up against McCarthy, it turned out. What came of it was that he was deathly worried—my father was worried that he, in order to save his job, would spill the beans, and they would try to get stuff out of him. I don’t know if it happened.

But much more serious was a different one. A lot of his patients, besides being Jewish, had been communists in New York City in the thirties. And a lot of that played in their histories, and part of their psychosis or neurosis they had was associated with that. Not that there was something wrong with it; it just came out. And he would take notes. He would sit—you know how a psychoanalytic session goes. The patient lies down. The doctor is behind the patient, and the patient doesn't see the doctor. And the doctor is taking notes the whole time. And these voluminous notes.

And my father had a—thank god he had lousy handwriting. But what he would do is once the McCarthy era came along, he asked me on Sundays to convert—he thought this was a way of—that was his—what I'm about to tell you now is the misgivings he had about American education, it will tell you right away what it was—he thought there was nobody in America who could understand Greek. Okay? And I mean by that Greek letters. Forget about the Greek language. Okay? So, what he asked me to do is to take his notes, and convert them into—leave them in whatever language he was writing, but put them all—where there was an “A” put an alpha. Where there was a “B” put a beta. Stuff like that. Okay? And he thought that would hide it well enough, because he couldn't imagine any American knowing any of that. Which was of course nonsense, but that was—okay. So, I would do this every Sunday. And I got to know a lot about his patients, which I never should have known (laughter). So, that was his strategy. Luckily, it never came that there was a real investigation of his friend or of his office. Sort of amazing.

Zierler:

What about his own personal views? Did he ever feel loyalty to the United States?

Weiss:

Oh, yeah. I'll tell you what did it to, of course he did. What happened—I'll tell you what did it for him and many others, is that he felt capitalism was the problem, and it’s greed. That’s the worst of it. And when he got to the United States, he saw this rampant capitalism. I mean, it was really rampant. The greed was pretty evident, but to him, he didn't see it as being as devastating as the problem in Germany. In other words, there were poor people, but they weren’t as destitute and left—now they are, but in those days, at least to him, it didn't look like they were as destitute as the German—let’s say the lower levels in the German society. I don’t think he understood well enough what was going on here.

So, the things that caused him to modify his things were two. First, he didn't find that it was as anxious in the United States as it was in Germany. That was number one. And the other one was that the thing that cooled him on it, especially the communism that was exercised in Russia, was that Hitler and Stalin had gotten together to make that pact to eat up Poland. And he thought that was an absolutely criminal act on the part of both the German and the Russian government. And that caused I think—in fact, in the United States, and many of his patients had exactly that feeling—what got them to leave communism was that particular step. The Hitler-Stalin pact was a real watershed for most American communists. And I think it was for him, also.

Zierler:

Meaning that the Soviets were rotten, also?

Weiss:

They were rotten, also. And that was already known to him a little bit, when he decided—or he was forced by my mother not to go to Russia, because Stalin was already killing doctors and killing Jews. But that was not so important to him. He thought he could fix it. But it got worse and worse as time went on. So, let me take you back, I guess, to the—

Zierler:

Back to school, your school.

Weiss:

Well, back to non-school (laughter). The hobbies and school. I know I can bore you to tears with that, but I'm not going to tell you everything. There’s one single thing that was really very important in that whole business of high fidelity. I could have made a killing had I stayed in it, and never gone to college or anything. Because the whole thing blew up—I don’t know, became a real business—about fifteen years later, ten years later. But at any rate, the problem I had was the following: I don’t know if you remember—you're probably too young to remember the old phonograph records?

Zierler:

I mean, I've seen them.

Weiss:

You've seen them. Have you ever listened to one?

Zierler:

Yeah, sure. We had a neighbor that had one.

Weiss:

Well, were you pleased with it, or were you horrified?

Zierler:

What did I know? I thought it was cool to look at.

Weiss:

Yeah, okay. Well, it was awful.

Zierler:

Okay (laughter).

Weiss:

(Laughter) I mean, after getting spoiled by the FM radio—and people wanted to have a phonograph because the things were recorded, and I could not tame that goddamn technology down so you enjoyed it. But the problem was very simple. The surface of the phonograph record was shellac. It was the smoothest thing that—people let’s say in the teens down to the twenties—they got a little better at it—that they could imagine. I don’t know if you remember, the way you got the sound up is by running a needle in the record in a groove, and that groove was modulated by the sound. So, a needle was wiggled and then you had a transducer where the needle was attached.

And the problem was very simple. There was a very good company in Britain called Decca FFRR—full frequency range and full dynamic range—and they sold these wonderful records which were great, until you got to quiet sections. I mean, for example, since I loved the piano even then, you listen to a Beethoven sonata, you listen to the first movement which is loud and quick, and the fact that there’s this terrible hissing noise in the background doesn't bother you. Take the second movement, where things are slow, and people are having big gaps, and you listen to the sustaining pedal more, what it’s doing, and all of a sudden, it really got at you! I mean, that horrible hiss.

And I said, “We've gotta fix this. This has gotta be fixed.” And that, to me, was a transition point. Because what happened is what I knew was street electronics, electronics you learn by reading in Popular Science. But I didn't know enough math. I didn't know enough about circuit theory. I didn't know enough about real electronics, professional electronics, to do what I was about—which I'm about to tell you. I had an idea—I think one way to get rid of that goddamn record scratch is to change the bandwidth of what you're listening to as a function of the amplitude. In other words, in a very quiet region, what you do is you don’t put all the high frequencies in. The thing that makes the hissing noise is the high frequencies, and you can cut them out. And then, when it gets louder, or when—yeah, louder or faster—you open the bandwidth up again. That was the solution I came up with. And I couldn't make a circuit really do that, that wasn’t worse than the hissing noise. I mean, it could go cchhhhhhh, chhhhaaaaa, chooooo, like that. You know? It was absolutely hopeless. Nothing I could think of would do it. So, I decided I’d better go to college.

Zierler:

And as far as you could tell, none of the engineers at Westinghouse or GE had come up with this on their own?

Weiss:

I didn't deal with Westinghouse engineers. I dealt with my buddies at Harvey Radio! The place who sold stuff to—consumer electronics. That’s about the most technical group that I knew, about this. And I did talk it over with my high school science teacher, and he didn't know enough about it. He didn't care about music, anyway, but he was a very sweet guy. Anyway, so that was the motivation for going to college. And then because of that, I got myself into even worse trouble once I got into college. But I mean (laughter)-

Zierler:

You must have been a pretty good student to get into MIT, though.

Weiss:

I don’t know how that happened. I really don’t know how it happened. I mean, I had not great numbers on the Regents. Here, I think it was the science teacher. Because I brought him home once to listen to this thing that I made. I brought him home to listen to the FM set and all that stuff. And although he didn't care much about music, I guess he was impressed by the amount of junk that was made to work. I don’t know. I would love to see the recommendation he wrote. But I suspect it was that. And by the way, all my life, I've always told people, “Fuck the grades. Work with somebody.” We'll get to that.

Zierler:

So, your game plan at MIT at first, I assume, was what? Electrical engineering?

Weiss:

That’s exactly right. I applied as an electrical engineer, with a single purpose not of getting educated broadly, but with the single purpose of being able to talk to somebody who could help me out on this problem with the goddamn phonograph records.

Zierler:

Mission accomplished? Did you find that person?

Weiss:

No, and I'll tell you what happened. That got me so pissed off after a while that I really got into trouble. What happened is MIT isn’t the place I thought it was (laughter). It was very rigorous. I mean, this is 1950, now, okay? It was very rigorous in the sense that they had a curriculum that everybody took if you came into the place, irrespective of what you were going to do later on. So, the freshman year was prescribed. Which is okay. I learned a little physics. I learned some math. I had terrible trouble with calculus, but I learned that eventually, by pictures. I don’t know how much calculus you learned in your life, but pictures work beautifully for calculus.

And I was waiting—so I got a job that summer working, yeah, for Sperry Gyroscope in Brooklyn, doing drafting. You know, drawing up stuff. That didn't teach you much. Came back the following—came back as a sophomore, and then I started electrical engineering. And the trouble is that did a job on me. In those days, the electrical engineering curriculum consisted of really most rudimentary electricity—making power. How to make power. How to make motors run. How to make generators go. Electronics was reserved for the seniors. It was a very esoteric topic. I didn't realize that what you were given was how to design a truss to hold a generator so it didn't collapse and fall off, or that the motor didn't twist out, out of its mounting. So, we learned statics. We learned all sorts of stuff.

I wasn’t going to—that wasn’t important. And in those days, they didn't have easy relationships for an undergraduate to couple to a faculty. Now, of course, it’s very, very different. And so I stuck it out for one term. No, either one term or two terms. I don’t know. But I was already looking now at the catalog. Where did you have less requirements and more ability to talk to people? And stuff like that. And I decided that if I was going to return to MIT, I was going to take physics.

And I worked that summer between freshman and sophomore year—no—yes, sophomore—that’s right, sophomore and junior year. Because sophomore year, I was still an electrical engineer. I decided to become a physicist for my junior year. And I worked for a guy who was a total—as far as I'm concerned—criminal. This is an important story. I stayed in Cambridge at the time, and exploited the fact that I knew electronics. I got a job pretty quickly. But still, I couldn't really do professional electronics. And I don’t know, this is an important history in our country—in 1951, it was still—or ’52—it was still considered reasonable that you could have an atomic war, and in fact that you could survive an atomic war.

Zierler:

Right, right.

Weiss:

And they were trying—the military was, and the government, at the time, was trying to convince the public to build fallout shelters, to have stuff in their houses that would allow them to deal with too much radioactivity. I mean, in the end, it turned out they didn't tell you, but they were very willing to decide, “Okay, we will put—we will estimate how many people will die. If it’s two million, that’s fine. If it’s one hundred million, it’s too many.” That kind of thing. To triage was their philosophy.

And this guy who I didn't realize it—there was a time in those days when almost any small business could start, if you entered into that kind of thinking. And his idea was to build an automatic blood cell counter. In other words, a thing that you could put a sample of blood into, and put the sample under a microscope into a rotating platform, and the microscope would look down on that, and it would count the size and the number of red blood cells, and also would do a differential estimate of the white cells. Okay? And that was all designed—he was supported by the Air Force to do this, and he had to deliver it within something like the year after I got hired.

But I began to—okay, so what happened is I worked on systems to look at signals on photodetectors through the microscope to gauge sizes of things, and built a little device that—they weren’t blood, but to look at things that you could look under the microscope and gauge the size and number of things that it was looking at. And built the counting electronics for it. And in the middle of all of that, because the Air Force was getting worried about this small company not delivering reports at a rate that they thought was adequate, he asked the three of us who were employees of that company to effectively say we were further along than we were. And I wasn’t willing to do that.

Zierler:

You saw the bigger picture of what was happening?

Weiss:

Eventually, I—well, I haven't quite yet—I caught on in the middle of working for this guy that this was a sham, and on top of that, the whole philosophy was crazy. And so I quit. And so OKAY. Well, that summer, because I had made some money, I went on a trip to Nantucket. You know where that is?

Zierler:

Yeah.

Weiss:

On a little bicycle trip. Nantucket. Took the ferry. And I ran into an absolutely wonderful woman. She was something I had never seen before. Was not an intellectual. She was an emotional person. But her emotions were very reasonable. I don’t know; that’s the best way I can explain it to you. I mean, she came to decisions not the way I did, but she came to decisions that were right despite that! In other words. And they were emotionally sensible. Turned out to be intellectually reasonable. A very strange combination. I would have never guessed something like that. And so I went nuts. I fell head over heels in love with this person. And so, she happened to be a musician, happened to be a pianist. She was a student at that time at Northwestern in Evanston, Illinois. And so when summer was over—we hadn’t spent the whole summer together, but probably a couple of weeks together in the summer—she came over to a house where there happened to be a piano, and she played—I don’t know if you know, but she played one of—now I see what it is—one of the easier impromptus, Schubert impromptus. But I was absolutely swept off my feet. Couldn't believe it! And she I mean just absolutely (I don’t know exactly what was said but the thought must have been learning that one could completely change the meaning of the music by subtle things such as playing legato or staccato or changes in tempo) and then be emotional—be right about it. And that you could put the thing—it wasn’t just playing notes; it was making it right in terms of the emotions associated with it was beyond my—I mean, I knew music a lot, but I didn't realize how it was made. So, my god!

So, we had many, many letters between each other from—I was still at MIT trying to now be in physics, and she was doing her stuff in Evanston. And we met at Christmas. I went to Bucks County to meet her family. Very interesting people. But I was probably not old enough for them (to be a serious candidate) or I wasn’t sufficiently aware. I was still too much of a nerd. And something didn't work on that little visit. And so, the letters didn't—there were less and less letters, and then all of a sudden, there were none. So, I decided, “Damnit, I have to go fix this!”

And so this is now in the second term of junior year. Yeah, Christmas—second term of junior year. I decided I’d go out to Chicago and find out what was wrong and figure out how to fix it. I mean, a typical nerd mentality. Well, she was very sweet about it, but she of course was beginning with another person. But that only came out a little bit—we had some interesting times. And I got very, very upset by that, and in fact started—I don’t know—how well do you know all the Schubert songs?

Zierler:

A little bit.

Weiss:

Do you know the “Winterreise”?

Zierler:

If I heard it, yeah, I’d recognize it, I'm sure.

Weiss:

Well, it’s not—yeah, the hearing of it is very important, because the music is absolutely gorgeous. But the poems are ridiculous adolescent love poems.

And what the “Winterreise” has in it, if you look—go look at it, and you'll see—the very first song is he has been thrown out by his girlfriend, and then he starts walking around the country with more and more feelings of wherever he looks, he sees her in everything. He sees her in the trees, or he sees her in the dusk, or in the waves on the water. She is everywhere. Can’t get away from her. And so that’s what all the different elements of the “Winterreise” are, at the end. And that was sort of what I was going through. It’s very adolescent, but enormously deep. So I hung around Chicago trying to fix this, but it didn't work. And I came back to MIT in May, having been there for probably two months. And like an idiot, I thought I could just take exams. No, no.

Zierler:

Just pick it right back up?

Weiss:

(Laughter) That’s not the way the thing works! You couldn't just take exams. You had to—they took attendance, but I didn't know that. And none of the homework was in, so you were screwed up. So, they flunked me out. I mean, I really flunked out. I flunked every course. Well, all right.

Zierler:

So, you're officially a college dropout at this point.

Weiss:

I had become a college dropout, yeah, the second term, junior year. Next would have been the senior year. My parents noticed that (laughter). “Why did you fail school?” Okay? That was something that was sort of obvious. Okay. So, what I did is I walked around—and this is the beginning of science. I don’t know if you want all this, but—

Zierler:

I want every bit of it!

Weiss:

(Laughter) Okay. What happened is that—and this is where I think the things that—whatever lessons come out of my life that are valuable to others would actually be worth something. Because what happened was that—I don’t know how well you know MIT, but MIT had a very old building in the back, which was built during the Second World War, where all the radar lab stuff was done. Huge building. It ran the length of a city block. It was made of asbestos siding on the outside, and all wood on the inside. Huge building. And it was loaded with—at the time, in 1953, it was loaded with many of the people doing experimental work in chemistry, electrical engineering, physics. Even Noam Chomsky was in that building. You probably know Noam, or you know of him.

Zierler:

Sure.

Weiss:

He and his group, a linguistics group, was in that building. Interesting bunch of people. And a lot of people doing neurophysiology were in that building. And so I walked around, and I walked around looking at—because you could look into the labs, because there were grillwork—you know, grillwork, you could look through it and see. And I saw some people—two guys working and screaming at each other. There was one guy sitting on a desk with a galvanometer on it, which is a device that has a little light that tells you—the light moves, and it’s like a meter. The light gets pushed by a magnetic field. Well, not the light, but the mirror, gets pushed by a magnetic field. And as you get more current from something, the light moves to the right, let’s say. It’s called an optical galvanometer.

He was sitting there looking at this, and the thing was going back and forth like that. The other guy was in a bosun’s chair, hung from the ceiling, sitting on top of an apparatus with two knobs. I didn't know what all this was all about. And they were yelling at each other. “Can’t you hold it steady?!” “No, no, no, less!” “No, no, go right!” Okay? I said, “Wow, I could help these guys out!” So, I walk in. They're mad as hell at each other, Okay? And I said, “But I can help you guys. I know some electronics.” And they said, “Okay, hang around.”

And then eventually what happened is I got a job with those guys. And it happened to be the lab of a guy who became my second father, actually more important father than the first one. And that was a guy named Jerrold Zacharias. And what was he working on? What were they all working on? Well, Jerrold had been working on military problems, because he was a very important person in the radar laboratory. He had worked at the bomb lab as well, in New Mexico. What the hell’s the name of that place?

Zierler:

Los Alamos.

Weiss:

Los Alamos, exactly. Los Alamos. And in fact, had been involved with—I don’t want to go into all what he had done, but he was at sort of the level of being both a very good experimenter, but also had moved into management. Management of experiment. And on top of that, he was a guy—and this is where the Vietnam War comes in, in a minute—who had taught the military that they needed to have more advice than what they were getting. In other words, that happened again during the Vietnam War, and at the end of that is when a lot of bad stuff happened, when Nixon stopped that all. It was Nixon’s fault, by the way, that it stopped.

But the military—for example, he would run summer studies. This may be important to you as a historian. He would run summer studies for the military. You know, they’d come along—for example at the time, the worry was, could the Russians bomb us, with bombers? This was before the intercontinental ballistic missiles. And the question was, what could you do to avoid—how could you find out about that earlier? Any moment you had earlier would help a lot, if they were going to charge on us, with a bunch of bombers.

Zierler:

Was this the work that was being done at Lincoln Lab?

Weiss:

We'll get to that. Lincoln Lab didn't quite exist yet, but that’s absolutely right. Lincoln Lab is the derivative of this. And he was deep into that. He and a guy named George Valley both were involved in making Lincoln Lab. But what came of it was that it was very interesting to listen to him talk, and I think it’s an interesting study. I think it has been done, but I don’t think it has been done adequately by historians, is that the military would still prefer to work in the following way: here’s a bunch of scientists on one side of the fence, here’s the military on the other side of the fence, and let the military do their strategizing on their side of the fence, come up with a set of problems, like how do you make a better bomber, how do you make a better radar, how do you make a whatever—instrument—and then they would try to sell the idea to the scientists. So, hop over the fence and tell them, “We need these things” but not tell them what it’s for.

Because they felt—the military was still thinking that those people on the other side of the fence were untrustworthy. You couldn't get them into strategic thinking. You could get them into scientific thinking and technical thinking, but we're not going to tell them what we're actually planning. And Jerrold decided, “Fuck this!” And in fact—and that became, by the way, very important in the Oppenheimer hearings and everything that followed.

What happened is he said that he—what happened is that he decided he would not take any military projects unless he knew exactly what they were for and the military hopped the fence and was with him. In other words, the military was involved in the project. And of course, the project wasn’t open, but it wasn’t controlled by the military. In other words—it’s a tricky thing. It wasn’t going to be something they could publish, but it would be something where the scientists had military consultants. The military consultants were the people who had asked for the study.

Zierler:

Was this before, Rai, when scientists had clearances?

Weiss:

Oh, yeah. They had to. Very much so.

Zierler:

No, so you're saying—so Zacharias, he had a clearance, but it’s the—

Weiss:

Oh, of course he had clearance.

Zierler:

So it’s the culture of the matter that was the problem?

Weiss:

That’s right. They had clearances.

Zierler:

But you're saying that wasn’t enough, to have the clearance?

Weiss:

No, no. Because look, even with clearance, there were people you trusted, and people you didn't. There were different classes of clearance. Otherwise you couldn't have built anything. So no, no. In the deepest sense, the military wanted to control why it was. And so what happened is that Jerrold had gotten caught several times in this first mode, and they gave advice, but the advice was worthless, because they didn't understand the whole problem. For example, how do you deal with conveys of stuff from Europe to the United States or vice versa? And they were given problems about how submarines should—submarine defense, but not the whole problem. And he decided that was worthless. He was not going to put his time into that anymore, and he demanded to know what the whole problem was.

And by the way, from then on in, there were many summer studies where the military and the scientists worked together on such problems. So, what happened is—if you're interested in this, I think there’s a book written by Jack Goldstein, about Jerrold Zacharias. If you're interested in that, you can find it in—it’s Jack Goldstein and it’s called Another Kind of Time, is the name of the book. Because Jerrold went on and did the same kind of thing for education, ten years later. We'll get to that.

So, what it was—the interesting thing—Jerrold had started this lab at MIT in that little interim right after the war when he wasn’t yet trying to solve the new military problems, the Cold War problems. He had started a group in nuclear physics doing something called atomic beam techniques. That’s a technique that was developed by I.I. Rabi at Columbia. And Jerrold had been a postdoc in that group all during the thirties, and understood the atomic beam technique, and he started what’s called an atomic beam laboratory at MIT. And that was the beginning.

And then when he went and started—and feeling that he had to still go back into dealing with the military, because they thought they were going—they were not getting good advice—he turned the lab over to a guy named Vincent Jaccarino, who’s still alive and at the University of California in Santa Barbara. But you can never get to talk to him, because he doesn't want to talk about this. It’s sort of interesting.

Zierler:

Okay (laughter).

Weiss:

So, what happened is he turned the lab over for a couple of years, and what happened to the lab—it became a really good nuclear physics lab. Virtually nothing wrong with it. So he comes back from his engagement with the military and looks around at what’s going on, and wants to start a new project on—and it’s just about the time I came into that lab he comes back from his Cold War effort and he decides that an interesting thing to do is to make an atomic clock.

Now, he got interested in that for a couple of reasons. He realized that the military needed a thing like that. But more importantly, he felt that it was a fundamental thing, to learn how you could make time measurements that were absolutely independent of all the vagaries that most clocks had. And I won’t tell you any more than that. Until the atomic clock, the clock that was determining the time was the rotation of the Earth. That determined the length of day by using stars as a way of when a star—you looked up, and if a star happened to come through your telescope at a moment, and it was right in the middle of the telescope, you set the clock. And then you wait a day, and then you look again, and the next time you saw that same star, that was equivalent to a day. And you set up your clocks so they would be synchronized to that.

And that turns out to be a clock that’s no better than a part in a hundred million. And for many, many things, like guiding missiles or if you want to think militarily, or if you're trying to think of precision experiments in science, that’s nowhere near good enough. So he went and started a whole program in that lab to try to make an atomic clock. And he asked me—he didn't ask me; I was an employee, and I built all sorts of stuff for the experiment. I got to know him pretty well. And I'll tell you a cute scene that happened. They had built—we had built all together—we had built a small clock, a clock which was probably as tall as you are. That was put at the Naval Research Laboratory, and several of them were built, and then they were given to industry to build. Industry built these clocks, and they showed up at various places that are responsible for keeping the time of the world. The Naval Research Laboratory was one of them. The Naval Observatory in Washington was another. And other places in the world. And so okay, here was a clock that was good not to a part—I'll have to use exponential notation. Do you mind?

Zierler:

Please.

Weiss:

The initial Earth clock is a part in ten to the eighth. This clock was a part in ten to the twelfth. No, a little better than that: a part in ten to the thirteen. No, twelve, twelve. Sorry, twelve. And Jerrold started thinking—and a copy of that clock, by the way, is in the Smithsonian. It’s sitting there. Completely mislabeled, by the way (laughter). But never mind. So, he had the idea, which is really where science really comes into this, that when you looked at how time is kept, there’s something deep in it, in Einstein’s theories of physics and nature. Turns out that time is another dimension of space and time. And by moving in space and time, you can convert motions in space to motions in time. Turns out that these four dimensions, in the deepest mathematical sense, were all equivalent. That’s not quite true, but that’s the way that it was written.

And so consequently, one of the things he set right away as an aim was to measure something which Einstein had predicted but had never really been verified adequately, and that is that if you make a clock out of—however you make it—and you put—let’s say you make two of them, and you make them as good as you can to be the same—you put one in a gravitational potential, or put one, let’s say, in a valley where there are mountains, and then take the other one and put it on top of the mountains. And that was in fact the experiment he proposed to me. And the idea was that if you do that, you will find out that the clock that’s in the valley is running more slowly than the clock that’s up on top of the mountain. Not by much, but by some.

Zierler:

Enough.

Weiss:

It turns out that’s an effect—now, that’s absolutely critical to get you home if you're using GPS. It’s a thing that you can’t get away—without knowing the Einstein redshift. Well, you won’t get home. You'll get some other place in the city, but you won’t (laughter) because the satellites have different clocks than the ones on the ground. You have to keep that in mind. There’s just no way out.

So, okay. So, he designs an experiment that he hopes will make a clock that’s good enough to do that. And in order to do that, you need a clock that’s good to a part in ten to the fourteen. To do that experiment. And so the idea was to make a clock that—what determines how well you can measure the time is—and here I have to go a little bit into the idea of how you do this—you have to find an atom—first of all, why is it called an atom clock? Atomic clock. It could be called a quantum clock. I don’t care what you call it. But every clock has to be the same as every other. In other words, when you take a clock and use a spring and a pendulum or whatever, you can’t make the thing the same. They're all a little bit different. And remember, your clock that’s on your wrist, probably you have to reset it—well, I don’t know if you have a clock on your wrist.

Zierler:

No, but I—yeah, I know what you mean. Right.

Weiss:

Yeah, you have to reset it, depending on how good it is, at least once a month. Well, no, that’s nowhere near good enough. That’s a part in ten to the six clock. That’s lousy. You can’t have a thing like that. Anyway, so what you do is you find out—the one thing that’s elegant is—and this is where I began to learn quantum theory, because I didn't know any of that—but here’s a real thing in front of me. I had to learn it, because it was fundamental to the whole damn thing.

Zierler:

By the way, Rai, so are you back in MIT? Are you matriculating again?

Weiss:

Am I matriculating? No, not yet. Oh, god, no. No, no, I was two years a technician. Anyway. But that was the most important experience of college, being that technician.

Zierler:

Outside of school.

Weiss:

In Jerrold Zacharias’s lab. It was absolutely more important than any other thing that happened!

Zierler:

And I'm curious, just as a mentor, was he encouraging you to go back to school, or was he thrilled that he had you full time?

Weiss:

No, no. I'll get to that. That’s coming. It’s a cute little story that relates to that. But the thing was that in an atomic clock, you used the cesium atom—it’s one of the alkalis—and inside the cesium atom, there’s an electron, and the nucleus, and they're interacting with each other. Effectively, you can think of the nucleus as oscillating in the magnetic field of the electron. Leave it go at that. It’s an internal oscillation inside the atom. And all cesium atoms are made the same. Man can’t tinker with those atoms. God made those atoms or whatever you want to call it. Nature made those atoms. And they're all identical. That’s fundamentally the idea behind all of this.

And so, the only problem you have is you don’t have to make this atom, but you have to watch the atom for as long as you can. And that was the problem with making a better clock. In other words, the first clock, the one you might see in the museum, has the atoms going vertically, and they're shooting up out of an oven, and they go fwoooot like this, and they last in your observation time for maybe only a millisecond, a few milliseconds. So you didn't have much time to look at them do their stuff.

On the other hand, the thing he was thinking about is to make a thing that looks like this, he wants to throw the atoms up slow enough so they will fall over like a baseball. Slow them down somehow so they go up like that—ffff—go up and fall down. That gives you a very much longer time to watch them. And that was, in the end, called the Zacharias fountain, the idea that now is absolutely working like a charm, with modern technology, but in those days, it was brand new. And to be honest with you, it didn't work, and we'll get to the reason why it didn't work.

And so he and I worked on this thing—we built a huge apparatus in this old building, and the old building had the right attitude for how to do this. What it is you could drill holes in the wall; nobody gave a damn. You could drill holes in the ceiling; nobody gave a damn. You could tear out windows. Nobody gave a damn. The building was a piece of junk; nobody gave a damn. It’s just exactly the sandbox for an experimental physicist.

And so we built the apparatus, wanted to see atoms do this. And, well, we didn't see anything. It turns out you're dealing with atoms that normally would go shooting up to about five kilometers up in the sky. You're dealing with very, very slow atoms, that theory says ought to exist, that do that. They're in the same population coming out of a source as the very fast ones. And so we made the apparatus longer. We cut a hole in the ceiling, made it longer, watched for this, still didn't see anything. And by this time Jerrold had gotten interested in education. I'll get to this. So now, some—yeah, it’s about ’55, ’56.

Zierler:

And Rai, what’s the work arrangement? Is he telling you what to do? Are you running this on your own with his oversight?

Weiss:

Well, it gets more and more that I'm running it myself. Why? Because he’s a busy guy. But I'll tell you, the way we got to be really buddies was during that epoch. And the thing I'll never forget, and it will probably make me cry when I think about it, is that I had started taking piano lessons at the Longy School, mostly because of that influence of that young lady. And I had gotten to the point where I could play with a violin. In fact, I was studying the—I don’t know if you know the “Spring Sonata” of Beethoven.

Zierler:

Mmhmm, sure.

Weiss:

It’s a violin-piano sonata. The first movement of that is sort of playable by an amateur. The rest of it—mm—but the first movement was. And so I had been practicing the first movement. I was going to play it with a violinist, some time around. And I'm up on the top of the apparatus. This is now three stories high, this apparatus. But I'm in the middle section of it, adjusting something. Jerrold is down on the ground and looking at something. And I don’t take any notice of him, and I start whistling the piano part for the “Spring Sonata.” And shit, he comes in right [clapping sound] where the violin comes!

Zierler:

Oh, wow! (laughter)

Weiss:

I couldn't believe it! (laughter)

Zierler:

You had no idea you had this connection.

Weiss:

Well, yeah, I had no idea. And so, I keep going, and suddenly I notice that he knows everything! Right down to—and so I come down, and I come down, and I say, “Jerrold, why do you know the ‘Spring Sonata’?” “Is that what they call this, the ‘Spring Sonata’?” he said. I said, “Yeah, it’s a Beethoven sonata.” He says, “Well—” Then he explains to me his background. I won’t go into all of it, but he was brought up in New Orleans by his mother and father. His mother was a professional violinist. And in the house was Beryl Rubinstein, who was a friend of the mother. And Beryl was a first-class pianist. He later became the head of the Cleveland School of Music. Beryl—B-E-R-Y-L. He’s actually not the famous Anton Rubinstein, or Arthur Rubinstein, but he’s in the same league as these guys. And so, what they were doing is while Jerrold was upstairs playing with his mechano set or whatever he was doing, they went through all the literature, and they kept playing things over and over again. They wanted to perform for them. And Jerrold had remembered it all.

Zierler:

Wow.

Weiss:

That’s just sort of amazing. Anything I was going to work on, he knew it (laughter).

Zierler:

Wow.

Zierler:

So, I mean, my god! And so, we got to be real buddies. Not only that; we had already become buddies. But that—that did it. And so, it was a very informal relationship. And by the time—you keep asking—there was no—I made money by punching a clock, but nobody kept track of me. And so I would spend fifteen hours in the lab; I wouldn't get more than eight hours of pay. Doesn't matter. That kind of thing. So, anyway, here comes the sad news. The saddest news is that Jerrold now gets—because he sits on various committees in Eisenhower’s regime. It’s Eisenhower? Yep. And so is Jerry Wiesner, the head of MIT at that time. They have now moved from military advice to general science and educational advice. And Eisenhower was actually not our best president, but at least he would listen, which is something—I don’t know how much you learned about him, but—

Zierler:

Sure. A lot. A lot.

Weiss:

A lot, yeah. So what Jerrold had decided by ’56 was that American science education at the grade school level and at the high school level was for the birds, and in fact if there was going to be a Cold War, we’d better get that fixed, because we don’t have enough people who think straight. That was his big thing.

Zierler:

And was your sense with Jerrold—was he a real Cold Warrior? Did he look at his science as in the pursuit of American defense, and things like that?

Weiss:

No, no. He was a patriot. For example, I think a lot now what he would do with Trump. I think he would try to blow him up. Because he would think of him as an evil, stupid asshole. And he thought of Nixon that way, by the way. I mean, he was not a patriot that is a gaga patriot, no. I mean, sorry; he was a patriot. He was worried all the time about the state of the country. He was not political in the sense that it was Republican or Democrat or Libertarian. But, “Is the thing smartly being put together? Is the smarts there? Are we getting ourselves into trouble?” That was his worry. And he felt that he knew how to think straight. I know a lot of people didn't like him for that reason. He was somewhat opinionated. But if he felt that he was right, he would go balls out to show people, “Here is the right way to do this.” And it was never on a doctrinaire thing. It was a rational—based on rationality, and it was based on some mathematics, usually, but sort of what I would call guesstimates. A guess made into an estimate. The kind of thing you give kids to worry about. When you ask somebody, “How many people are having sex at this moment?”—you can calculate that to a factor of probably twenty percent, thirty percent, okay? I mean, you don’t have to know a lot, but you have to put it together and think about it. You have to use some laws of physics if it’s necessary. And that—he was—no, he was not doctrinaire in that regard.

Okay, well anyway, so he sensed that there was a problem in American education, which was not adequate to meet the future. And that was also both for the economy—he sensed that the Russians, and the Europeans, by the way, both, would go running past us, in terms of economics. And so, he wanted to put the country back into a state where it was competitive with the rest of the world. And the military part was part of it, of course, but it was not the thing that drove it. Because look, he became later an advocate for understanding why there’s climate change, a whole bunch of things.

Okay, so what happened is that this apparatus—let’s get back to that, for a moment—the apparatus didn't work. It turned out that this thing I was telling you about, these atoms moving very, very slowly, they didn't do that. It turns out the books are wrong. The theory in books is wrong. Turns out the fast guys, the guys that are—many, many, many of the fast guys—they’d hit the slow ones in the ass and throw them out of the beam . That’s what would happen. And I couldn't sleep until I had established why that apparatus didn't work. So, he went off to do the education wars, and I stayed in the lab and I built a thing to show that that was the case. That’s what I discovered. That the fast ones were eating up the slow ones, and so there were no slow ones at all. But Jerrold let me do what I wanted. I mean, I kept on it. I loved it. I was eight years a graduate student, because I could do what I damn pleased.

And so, I spent the rest of my time as a graduate student trying to think of, “How do I get back in school?” You asked that? Jerrold has been my fairy godfather forever. I'm going to be actually blunt about it. He got me back into school, which didn't want me. They thought I was a real fuckoff, OKAY? He got me to finish the degree, bachelor’s degree. Then the only place I could go was MIT as a grad student. Why? Because Jerrold had the influence to get me back into there. My grades were so bad that (laughter)—I mean, nobody would take a gamble on me.

Zierler:

But he believed in you as a student, not just as a technician?

Weiss:

No, of course he did. He believed in me as an intellect. I mean, we got along, and we understood—I mean, I understood what he was doing. I'll tell you this—the important thing you tell somebody when you talk about a mentor is the following. You need not only to learn something from that mentor, but you also have to have the guy say, “You know, you're not as stupid as you look.” Okay? And that’s both things are absolutely necessary. And he performed that function admirably. I mean, I became a member of his family, in effect, which was tricky, but I became (laughter)- but about that epoch, I don’t know if you—well, you're too young; you couldn't possibly know this—there was a bunch of hurricanes that hit New England. One of them was called Hurricane Edna. And Jerrold had started a small business with Wiesner, because he was worried that Wiesner, who had had communist connections, and other friends of theirs who had had communist connections, would get into trouble, because MIT would not be forthright during the McCarthy era. They had started a business. And it was a business to build electronics. So he had made some money. He was not poor. And Jerry Wiesner was not poor, for the same reason. I don’t want to go into what they made, but they made stuff that the military could use, but also the civilian market could use. They were very successful at it.

So, it turned out he got well enough off, so he wanted to buy himself a house. He had a house in Belmont, but he wanted to buy a house on Cape Cod. And right after the hurricane—I was still working on that big clock at the time, but he was already spending more and more time on the education—he sends me down to the Cape to look at a house that he wants to buy. And he asks me, “Take a look and see if the plumbing is good, and the wiring’s okay.” So, I go down there, and I come to a place called Monument Beach, never mind where. And what was there was—here was a cliff, okay? There’s the cliff. And the house was sitting there like that. All of this area here—here’s the ocean on this side. All that had been eroded away. And so, the house was sort of sitting on solid stuff on this side. And as far as I could see—I was worried about walking over here, because it might go over, okay? So, I said to myself, “Holy crap. What’s he doing here?” So, I go down and look at the plumbing. It looks fine. I look at the electricity. It looks fine.

I come back to MIT and I said to Jerrold, “How the hell are you going to save that place?” He says, “Don’t worry about that.” He says, “I just want to know the rest of it’s okay.” And he already had a plan, and it worked. He put huge boulders under that thing. Built a whole other hill under it. And it cost him probably, I don’t know—I don’t know the number, but it cost him nothing compared to what that house is worth. And he lived there for a long time. It was his summer place. And a lot of the New Englanders—see, this was another problem. He was Jewish, and he got into trouble with some of his neighbors, who were very rich old Boston Brahmins who owned property there. But eventually they realized that he was a smart cookie, and they allowed him to be part of their society.

Anyway, where are we? (laughter) Oh, I meant to tell you the rest of how Jerrold was—and then we'll come back to the story a little—but how Jerrold was a member of—it was so actually. critical to everything in my life—not only did he do that mentoring right, which we talked about a second ago, but also into the actualities and realities of things. When I got too long of being a graduate student—eight years was too long for MIT, and I still hadn’t made my degree—he got me a job at Tufts University, as an instructor. Well, I did well enough, apparently—they offered me an assistant professorship there. But then I got my degree, and I decided to go to Princeton. He wrote the recommendation to get me a postdoc at Princeton. Then, I had been there two years; I get an invitation from MIT to become a professor. That’s all due to Jerrold. And I suspect it was all due to Jerrold when my tenure case came up, which was many years later, because I was a terrible tenure case. I'll get to that in maybe another interview. So in other words, wherever there was a decisive moment in my life, there was Jerrold. Yeah.

Zierler:

Amazing.

Weiss:

Yeah. Look, I'm yakking at you now for—well, we started a little late, but ten minutes more, and is that—do you mind splitting this up?

Zierler:

Yeah, we can do ten minutes. And then we can reconvene next Sunday. How does that sound?

Weiss:

That sounds fine.

Zierler:

Okay, so let’s keep going.

Weiss:

I mean, I don’t want to bore you to tears.

Zierler:

You are doing the opposite of that.

Weiss:

(Laughter) Okay, okay. Well anyway, so I finally—yeah. We find out why the clock doesn't work. I try all sorts of new ideas for clocks. Some of them are good. Some of them are—I won’t go into all of that. And eventually I graduate—my wife—I get married in the middle of that epoch. My wife gets pregnant. And I better do something about—you know, I have a responsibility all of a sudden. So, I decide to take one of the projects that I had been working on, which had a nice piece to it, which got published, was not all that impressive otherwise—I'll tell you what it was, if it becomes important. I had designed, during that time—trying to make better clocks, I had designed a device that could take any atom—doesn't matter what atom it is, or molecule—and bust it up into charged particles. It was called a universal ionizer was the name of it. And it turned out to be very useful for people in chemistry, a lot of people, physicists. But it was a tour du force experiment to do. And a lot of people wanted me to build them ionizers like that so they could go on in their research. And I did this for a couple people. But I eventually decided I’d better get myself a degree and get the hell out of there. So I did, and I did a very boring thesis. Uninteresting, very boring thesis. I won’t even bother telling you what it is.

Zierler:

No, you have to. I want to know.

Weiss:

Oh, I have to?

Zierler:

What’s the name of the thesis?

Weiss:

Oh. It’s the hyperfine structure of hydrogen fluoride. , I'll tell you one story out of that, if it’s important.

Zierler:

Please.

Weiss:

Okay. And this actually connects to something that’s going to be later. I built this apparatus with my own hands. Had this atomic beam apparatus that did things in a different way than they had been doing it. Instead of—all the early atomic beam apparatuses used magnetic fields to push the atoms around. I built an apparatus that pushed molecules around with electric fields. That was new for that lab. And I had to test the apparatus. I never—well, how could I get results of it? And the way you normally do that is you take an atom or a molecule, which has been very well studied, and you stick that into your apparatus and see if you get the same answers as other people do. That’s one way of at least establishing—you can do all sorts of initial work of carefully measuring everything, ab initio measuring of everything, but the final test should be, do you get the same answer as the other people or not? In this case, it’s one other person, who had a very, very precise measurement of the electric dipole moment of hydrogen fluoride. No, of hydrogen carbonyl sulfide HCS, which was a molecule you could stick into this apparatus, because you could detect anything.

And, well, I found out that—it’s listed in all the textbooks as a certain number. That’s unimportant. And I got a number that was about one and a half percent different, for a number that was good to five significant figures in the textbook, so ten to minus three percent. Yeah. And I got something that was no better than one and a half percent. What in the hell is going on? And I couldn't figure out what was wrong with my apparatus. So, I went to Jerrold. I said, “Jerrold—” I gave him this problem. “What do you recommend I do here?” And he says, “Are you sure you're right?” I said, “Yeah, I'm pretty sure I'm right. I can’t believe I—it’s so easy to make the measurement.” “All right, who did the experiment? Why don’t you go and visit them?” That thought would have never occurred to me. And so, I do that. I go down to Maryland, the University of Maryland. It turns out the guy who did the experiment was a guy named Joe Weber. Now, you may know that name in a completely different context.

Zierler:

Yeah.

Weiss:

He'll reappear later (laughter). So I go visit Joe, and I tell him that—I told him my problem. And he wasn’t very sweet about it. He said, “Look, you think we would screw it up?” I said, “I don’t know.” And he told his secretary, “Give him this graduate student’s thesis, and don’t let him out of there unless he gives it back to you.” Okay? So, I sat in a room. I read it. I read the thesis through. And sure as hell, I found out what they had done. They had screwed up. Okay? I don’t want to go into what it was. Well, I’ll tell you, but look, there’s too much to talk about. There is an error when you just assume that the electric field of something is the voltage between two plates divided by the separation of plates. That turns out to be much too crude. You have to know how long they are. You have to know—because there’s fringing field all around it, and they had left all of that out.

Zierler:

This is not just a detail. This is fundamental.

Weiss:

Well, it’s fundamental, yeah. If you go into the business of measuring electric things with precision, you've got to know and do it right. And here was a thing that was in the literature to five significant figures, and it was only good to two. Merely two. So, okay. So that was my first experience with Joe, and we will next time talk to Joe on some other day, talk about Joe. All right. So, I get my thesis. Tufts makes me an offer to stay, which is very sweet of them.

Zierler:

Well, wait. So first you bring this back to Jerrold, right? What does he say about what you found at Maryland?

Weiss:

No, Jerrold had disappeared. He didn't give a damn what I could do. I mean, later on, I’d tell him about it. He comes to my oral exam, but he is so busy with all his other stuff that he had forgotten all the details already. He had probably forgotten that he had given me the advice to see Weber, for example. No, pretty much, look, he left me alone. I wanted nothing better than that! (laughter) And by the way, he was very good about it. Any paper we wrote—we never wrote a paper together. We never wrote the—and it’s a mistake. We never wrote up the failed experiment with the big clock. That will come back later. People now talk about the Zacharias fountain, but that experiment is only mentioned in a footnote in places, that it didn't work. And by the way, just so you know, now it works.

Zierler:

Oh, really!

Weiss:

Yeah. And okay, maybe I'll just tell you that now. This is many, many years later. This is twenty-five, thirty years later. Nowadays, you can do something very clever. You can use lasers to take a gas of atoms, and you can take the lasers to slow them down. You can make them so they're sitting there, almost absolutely stationary. It turns out that’s called radiative cooling. And that’s something that was learned many—when the laser came to be important, that became a technique that very rapidly made very cold atoms. And a lot of the research that’s now going on is on atoms that have been slowed down, and so they're almost standing still. And with that, you can do this. In fact, it’s called the Zacharias fountain.

And the experiment that Jerrold and I wanted to do, with him in the valley and me on the top of the mountain, sending the signals back and forth to show the Einstein redshift—the Einstein shift of the clock—is now being done by the following. You take this cold bunch of atoms, and you let them warm up a little bit, so they just go up about this high. And then they fall down again. Really, I'm talking about that. And they spend a long time doing this, because they're so slow to begin with. And so what happens is they now can measure—the effect that we were thinking of measuring between a valley and a mountain, you can take and raise a little apparatus the size of, I don’t know, a grapefruit, and you can raise it by a centimeter, and you can see the effect. One millimeter! You can see a clock that’s displaced by one millimeter with respect to the other is going at a different rate. That’s unbelievable to me, but just wonderful! And it’s called the Zacharias clock, still, or the Zacharias fountain. Jerrold unfortunately died before that happened. It would have wonderful had he been able to see that.

Zierler:

Was your committee—it sounds like it was Jerrold and then everybody else. It wasn’t a big deal who else was on the committee for you.

Weiss:

No, no. Well, there were committee members—look, what does a committee do? It’s very rare—except in my epoch later, I had some very hostile committee members. But Jerrold—look, he was a great man in the department. Nobody’s going to fight with him. And I didn't screw up. There was no tension associated with that committee. And so then, yeah, I go to—I'm already working at Tufts to make money so we can support a wife and a child.

Zierler:

Now, Tufts is what? It’s a postdoc? It’s a tenure line position?

Weiss:

At Tufts, I got an instructorship while I was a graduate student. That’s all due to Jerrold. And then, as soon as I graduated, they made me an assistant professor. But I didn't take it. I wanted to work with a guy named Bob Dicke, Robert Dicke, at Princeton. And we'll stop right there, because that’s a good place to stop.

Zierler:

Robert Dicke at Princeton, okay! I will hit the end of the recording here.

 

[End 200607_0142_D]

[Begin 200614_0154_D]

 

Zierler:

Okay, this is David Zierler, oral historian for the American Institute of Physics. It is June 14th, 2020. I am so happy to be back with Professor Rainer Weiss. Rai, thank you so much for joining me again.

Weiss:

Well, thank you for having me. Anyway, let me tell you a little bit about Bob Dicke. When I say “Bob,” I mean Bob Dicke.

Zierler:

So just for context here, we are picking up from last Sunday, and we just left off where you had a big moment in your life. You had defended your dissertation. You were thinking about an opportunity at Tufts. In comes Bob Dicke at Princeton, and let’s take it from there. How did that come together?

Weiss:

Well, the reason it happened at all was because of the work that I had done with Jerrold, my mentor, on trying to measure Einstein’s redshift, that time difference between gravitational potentials. And it was an interesting time in the history of physics, because what happened is that Einstein’s general relativity had become—in very early days, of course, it was extremely popular. In 1919, Eddington made the world aware that Newton had been displaced. I mean, that was a big thing. But in the end, it turned out there was very little you could do as a physicist with general relativity, because all the things it predicted were so tiny that the technology just wasn’t ready. And it became a branch of mathematics. In most universities, by the middle of the thirties, there was no longer general relativity taught in any physics department. That was certainly true of MIT.

Zierler:

So, it’s not so much that general relativity fell out of favor or fell out of style; it’s just that the technology wasn’t there yet, so there wasn’t so much that you could do with it?

Weiss:

Well, it wasn’t being done the way physics was being done. In other words, you don’t just do theory. You have to do experiment to prove that—the theory predicts something, and you try to find out if it’s there, in the real world. And from that, you learn that the theory isn’t quite right, and you modify it. It’s a wonderful game we play. Keeps going back and forth.

I mean, it’s amazing how well general relativity works, given that it all came out of Einstein’s brain. I mean, it’s really sort of amazing. Because every time you test it, it seems to be that nature obeys what he says. So, it’s really quite a—that alone is an interesting story. But nevertheless, what happened was that general relativity had become—and gravity had become—a boring topic for physicists. They couldn't do anything about it. And Bob decided—although he had done some very elegant stuff in the Rad Lab at MIT; he was a very good atomic physicist—but he decided that gravity was worth his time, and in fact began to study gravity as a theory, and then began to realize that the technology had changed enough so that there were things you could now test that were not able to be tested before. And that to me was very exciting. He was almost singular in that regard. He was making a revolution, bringing general relativity back into science.

Zierler:

How was Bob ahead of the curve? What had he realized before anybody else?

Weiss:

Well, what he had realized is—actually, he wasn’t really that far ahead of the curve. What had happened—the thing that was really the place that was a decisive moment in all of that was something called the Chapel Hill Conference of 1957. And that was an attempt by several theorists, some famous people—I mean, they're not like Einstein. Einstein was not yet—had died. He died in ’55. He would have gone had he been alive, of course. But it was to see what was the new developments in general relativity since 1916. Okay? (laughter) And on top of that, experimenters began to realize, there now things you could do to show that the theory was right in better ways than was known.

And there were three famous tests that were considered when Einstein drew up the theory. One of them was that the motion of Mercury around the sun had a peculiarity that you could explain. That was realized by Einstein right away, by the way. Einstein lost three weeks of his life to joy after he discovered that he could explain the perihelion advance of Mercury. I mean, it had not been explained. But what a tiny little itty bitty effect. But to people who were doing that kind of astronomy, and knowing Newton physics, it was a pain in the ass, because it was a real blunder. It didn't look like Newton explained this.

And in fact, let me tell you one thing—people went looking for another planet between the sun and Venus—and Mercury rather—the sun and Mercury—to try to find another planet that could explain this anomalous motion of the planet Mercury. And they even gave it a name; they called it Vulcan. So, we lived with Vulcan for a while. But it turned out Vulcan didn't materialize. And so, Einstein was super happy about that.

Then the other thing that Einstein predicted is there should be bending of light by the sun. And that wasn’t yet available, really, but it became so, but with a very I think murky beginning. I don’t want to go into that with you here unless we're going to get really into the theory of general relativity. Eddington claimed that he had measured that. And that’s what made Einstein famous in 1919.

Many people tried to repeat that eclipse measurement where you could see the motion of the other stars relative to the sun, during the eclipse, and it’s a stinking hard experiment to do. I mean, and everybody failed with it. And everybody after Eddington also failed with it. And so it turns out that was really only established properly as a thing where you could get error bars, and you could say, “Here’s the noise,” in the 1960s. That was important. It was done by radio astronomy, ultimately. And then you didn't have to wait for an eclipse.

Zierler:

Why do you not have to wait for an eclipse? What became available that made the eclipse unnecessary?

Weiss:

Oh, because the radio astronomers didn't worry about the sunlight. You see, in the days of 1919, you could not see a star with a telescope, because you were using your eye to do it, or film to do it. And all the brightness of the sun’s corona—the sun was too bright, and you could not see a star in the daytime, see, and if you want to see the bending of light by the sun, you have to do it during the day, because that’s where the sun is. Okay? (Laughter) So, you had this problem that you had this terrible, bright thing in front of you. You had to shield it from that. And it was a very hard thing to do. I mean, the stories are just wonderful about how people went everywhere in the world during an eclipse trying to do this, and their telescopes would change temperature, and everything would go crazy during the eclipse. I mean, everything went nuts. The film to make the image would expand. It was just a horrible, hard experiment to do.

And the third one was already something—the other prediction that was made was that there was this thing that Jerrold and I tried to measure, namely the fact that clocks that are in strong gravitational field go more slowly than the ones in weak gravitational field. And that had already been shown, but in a half-assed way, in some stars. So, the idea now was that are there new things you could do those experiments better? That was one idea. Was the theory in any way found to be wrong? That was the other thing that this Chapel Hill conference looked into.

And then there were experimenters who had come now to all sorts of new tools, just as you guessed—the electronics, the beginning of computers, mostly electronics. And also, just the technology had enormously evolved since the twenties of that century. And they came, and they became to talk about measuring gravitational waves, for example. That was one of the ideas that grew at that conference. And the reality of those waves was very much discussed in 1957. Bob Dicke went to that conference mostly because he had just gotten interested in gravity. He had been interested in atomic physics earlier. But he was already thinking about things that—there were funny things to him in the Einstein theory which you couldn't explain easily. And I don’t want to go into that, because that could keep us going for the day. Let me just say what it is, and then if you get interested, we could look it up, or you can look it up. He was worried that Mach’s principle, which is the idea that everything is relative—when you accelerate with respect to all the fixed stars, they accelerate a little bit, too. In other words, if you take all the stars out of the universe, there should be no resistance to acceleration. And Mach’s idea was that everything was relative, even acceleration, not just velocities.

And it turned out Bob Dicke felt that that was not properly put into the Einstein theory, although Einstein thought it had been. And in one of his very elegant books he wrote for the non-physicist reader, called General Relativity he said, “Here are the three things in my theory that show how Mach’s principle is incorporated into it.” And it turned out, one of them was wrong. But never mind. Let’s go on. So, Bob had that in his mind. But to him, more important was, could you now do experiments that really were precise enough to really say whether general relativity is right or wrong? And that was new.

Zierler:

What technology had come about specifically that allowed this to be a legitimate question?

Weiss:

Electronics. And the beginning of the computing, but electronics, mostly. And the experiment that he decided to do in the sixties—and that’s why I knew about him—was he wanted to show that—and stop me if this gets too technical, because it is not worth anything if it goes past you and you don’t understand—the thing is he wanted to show that something which we normally do in physics is to say that the ability of a mass, like a body, to make gravity, to make a gravitational field—remember, gravity is made by matter, mass, energy. The ability for something, matter, to make gravity, called the gravitational mass, the gravitational mass, turns out to be equivalent to something which is a completely different idea, namely the inertia of the mass, how hard it is to accelerate it. And so that was something that Newton assumed way back. He couldn't really explain it. But it was a core of physics for years, that the inertial mass and the gravitational mass were the same thing.

And in fact, that was then enshrined into something called the principle of equivalence, which is that—and this is where Einstein comes in; he enshrined it into a much broader idea—that when you accelerate in a local region, you could not distinguish whether you're being accelerated or you're living in a gravitational field. In other words, an acceleration—you being let’s say in an elevator, you're pulled up, you feel an acceleration, you feel yourself thrown to the floor by the elevator pulling you up—that is exactly the same effect as when the elevator isn’t moving, and not being accelerated, but you're being pulled down to the floor of the elevator by gravity. And those two are indistinguishable, on small scales.

Zierler:

Which tells you what?

Weiss:

It tells you that acceleration and gravity have something to do with each other. That was brand new. And all that was really known was that a gravitational mass, which makes gravity, and also responds to gravity—okay, I mean pulled on by gravity—that that is the same mass as resists acceleration. That already was the germ of the idea. But the full exploration of the idea was Einstein’s statement of an acceleration and a gravitational field are indistinguishable And that, in fact, was the basis of the beginning of the thinking in general relativity.

So, to Bob Dicke, that was so fundamental that he wanted to do an experiment to show that. And before I got to Princeton, he had done an experiment with modern techniques using vacuum systems, using suspensions, using amplifiers to drive things, feedback systems, a whole bunch of wonderful ideas, all things you could now do with vacuum tube technology and the beginning of transistors. Because just about that time when I got there, transistors were coming in.

Zierler:

What was the connection? Did you know of Bob’s work? Did Jerrold connect you?

Weiss:

Jerrold knew of him because they worked together in the radar lab. That’s how he got me a job there. But at the same time, he was known to me by the fact that he had done this experiment. And so, I desperately wanted to work with this guy. So I got a job there. First time I’d ever left MIT.

Zierler:

Rai, I want to ask, because you hear so much about Princeton and what a unique kind of place it is—how formal it is, how preppy, how old fashioned—you're not really a Princeton kind of guy, right? (laughter)

Weiss:

No, and I found that out pretty quick! (laughter)

Zierler:

So, I'm curious how that worked in terms of fitting in or not fitting in. What were your initial impressions?

Weiss:

It didn't matter at all. I'll tell you why. The fact that it was on that campus was an accident, as far as I was concerned. The important thing was that the people in the group represented people just like me. I'll tell you what the real problem was with Princeton. There was a problem with Princeton. There were two problems at that time. Remember, this is ’62. And there were no women there. That has nothing to do with me chasing women. The whole atmosphere is all screwed up when there aren’t women around. People race to Vassar to get laid—you know, that kind of thing—on the weekends. I mean, it’s ridiculous, okay? (laughter) So, you had this completely fake atmosphere. That’s number one. And then the other one was that it wasn’t really a place like MIT was. MIT was a junk store, a real junk store. I'll tell you what! Because it had all this stuff from the war. If you wanted to build an experiment, you could do it in an afternoon by just going to the warehouse and saying, “I want that piece of crap over here.” “Oh, give me that, too!” And these are stuff from ten years ago. You could make an experiment work in an afternoon.

At Princeton, there wasn’t any nice warehouse full of crap like that. You had just to buy all that stuff, and it took three months to get it, and it cost money. It was just—ech! It wasn’t the right place to do a quick experiment. So, you really thought things through very carefully. That’s not bad, but anyway, that comes later again, into the story. The nice thing Dicke had—he had group meetings every week—and they would discuss all these different experiments that were being done. What I got as a thing to work on was not a very important thing. It was important to Bob Dicke for a while. His new way of looking at gravity assumed that there would also be what’s called scalar waves of gravity. Not just the gravitational waves that Einstein’s theory has, which are called tensor waves, but also waves that—for example, when a body does—take a balloon. And it doesn't have much mass; too bad. But take a balloon and blow it up and let it shrink again. Keep blowing it up, shrinking it, blowing it up. That motion, that uniform motion, spherically uniform motion, in Dicke’s theory, made gravitational waves. In the Einstein theory, it doesn't. And so, he wanted to see, was there any evidence for scalar gravitational waves and fields.

And he gave me that as an experiment to work on. He said, “Why don’t you go find out if the Earth has a mode which does that.” The Earth, our Earth, sits there, and after each very, very deep earthquake, it expands uniformly, spherically, and contracts uniformly. It takes about twenty minutes to do that. So, it’s a funny mode. The whole Earth expands and then contracts, once every twenty minutes. Goes through a whole cycle. And the idea was, well, if there is such a mode, which people had thought they’d seen before, they might be excited by these waves. That was the whole idea.

So, we took that same experiment that he had built, the same place where the experiment had been built, to measure the equivalence of acceleration and gravity, which was called the Eötvös experiment, never mind that’s the name for it. And we converted that into a thing to look for scalar gravitational waves. I spent two years doing that, and learned an awful lot. Learned a lot from Dicke. And I'll tell you one Dicke story, which is one that I will never, ever forget. That’ll come in a sec. But the important thing was that the experiment failed. The experiment with the big clock at MIT; that had failed, also. So, I became a master at failed experiments!

Zierler:

There’s value in failed experiments.

Weiss:

There’s some value, no doubt (laughter). But the failure here was actually very funny. It was a failure—we got the experiment together. Another guy with me was a guy named Barry Block, who was a theorist, but he enjoyed making jewelry. There are crazy people in the world; he loved to make jewelry. And one of the things he had learned how to do is make little springs out of glass. You know, springs. He enjoyed that, because you could make earrings, you could make all sorts of stuff out of that. Well, we used some of his jewelry to make this apparatus that would measure these gravitational waves, these scalar gravitational waves. And then once we got the instrument built, he lost interest in the whole thing. The experiment didn't interest him as well. So, I was pretty much on my own. But the gadget involved some other person.

So anyway, what happened is that we got it all built, and on an Easter weekend, in 1964, my wife and I—we have just had a child—we go to Washington D.C. to visit her sister. And we quickly go to the site near the stadium at Princeton, is where we—this was in a hole in the ground—make sure everything is working. We get in the car, drive to Washington. Okay? Well, the biggest earthquake in many, many decades occurred on that weekend. And in fact, the Earth sang like a bell for the next four years. Hard to do the experiment after that! And so, what we have is a very nice—we wrote a little paper about the instrument and what the purpose was, but we couldn't set any decent limits, at least I didn't know how to do it at the time. I didn't know enough mathematics. We could have, if we had been smarter about it, but we didn't. We didn't set any decent limits for these scalar waves. So, we published only an instrumentation paper. One paper came out of the whole time I was at Princeton. That was it.

Zierler:

Rai, how much interest was there in these projects that you and Bob were doing? Was the general physics community aware of this? Was there a lot of excitement surrounding these experiments?

Weiss:

That’s a good question. That’s a very good question. At the moment when we were there, it was not yet. It was not yet. It was still very early in the whole thing. And that question that you just asked plagued us over and over again. It was not yet in the mainstream of physics.

Zierler:

Because Bob was ahead of his time? Was that part of it?

Weiss:

Oh, he was way ahead of his time, yeah. And so was Jerrold in the sense that he wanted to do this redshift experiment with the mountain and the valley. Anyway, there were much more vital areas of physics at that time. The big thing that was going on was how you explain the quantum theory of electricity and magnetism. That was the rage, at that time. Everybody was working on it. If you had worked on that, everybody would fuss over you. But not this. This was sort of backwater.

Anyway, here’s the story I want to tell you about Bob, and then we're going to leave that. I learned a lot, but the thing that I learned most was from this one experiment, this one experience. Bob, when he made his own little theory of gravity, noticed that it did the wrong thing for that thing in Mercury. In other words, that motion of Mercury would not have been the same as was later found it to be and would have predicted the fact that the motion of Mercury would have gone in a counter way than the way it does go. Let’s not get into the detail of it. And so—if this theory was right. So, he had an interesting problem. His theory was predicting that the Mercury would go in a different way, but it had already been measured to go the way Einstein predicted, and yet he wanted to make room for this theory. So, he thought of a way that maybe everybody had misjudged the role of the sun in the motion of the planet Mercury. If the sun had looked more like—instead of a sphere had looked more like an American football, it would have done pretty much what Einstein’s theory said, as well. And nobody had really thought about that very hard. It’s true.

I mean, I'll tell you, the people who had measured how round is the sun—the only people who had done that at all were the people in the Vatican. The Vatican Observatory was allowed to do certain non-controversial science, and one of them was they had a cave, and a little hole on the top of the cave, and the sun would come through that hole and illuminate the floor of the wave. And these priests would run around measuring with a meter stick effectively, measure the diameter of the image of the sun along its—let’s call it the equatorial direction and the north/south direction. They kept measuring the difference, and they kept saying there was very, very little difference between the equatorial diameter and the polar diameter. So, the sun was round as a ball, okay?

But that wasn’t good enough for Bob, because it turned out they couldn't do much better than about a part in a thousand. Maybe only one percent really. Somewhere between a part in a thousand and one percent. So, he decided to make an apparatus—he announced this at one of the meetings, at those weekly meetings, that he had a guy, one of the guys, one of his postdocs, to look up the literature to see, how round is the sun? They came to the conclusion that the Vatican experiments were the only ones around that really were trying to be precise. And that’s when he said, “Okay, guys, I'll see you in about a week or two.” He disappeared. Absolutely disappeared.

And what he was doing is he comes back about three weeks later to one of the meetings, and he has under his arm—he had a set of drawings under his arm. He already had made them into, well, these blueprint kind of drawings. They were a big fat thing about the size of your hand, all full of designs. And he threw it on the table and he said to two postdocs of his, “Build this!” and he went off and did something else. And so, I had looked at it. I couldn't believe it. And so, I sat there with these two guys, and we unrolled this thing, and it was unbelievable what he had done. He had thought through every little piece of the experiment. He wanted to make a telescope that looked to see if the sun was round. That's fundamentally what it was. And he had invented the various schemes for whirling little discs that would be indicators. If the sun were round, it would give you a signal, an electronic signal. If it wasn’t round, it would give you some other kind of signal. And then had designed the telescope, had designed the telescope mounts. Had thought of all the optics that would distort the image make something that’s round into not round. He had already compensated for all the effects of the optic. Never mind, I won’t go into the whole thing.

I had never seen a thing like that in my life. A guy sitting at home, developing something like fifty drawings, and having thought through every step in the experiment, like he was a guy who had thought, “Okay, I'm going to do this. I build this, then I build that, then I build this. Oh, that could go wrong. Let’s fix it right now.” Then kept on going. Unbelievable! I couldn't imagine a person with that kind of an intellect. In fact, the people who were working on this thing, they came to me over and over again and says, “Why do you think he put this mirror in here?” “Why do you think he put that lens over here?” “What’s the electronics doing?” “Why didn't the electronics—” So I had told him, “Look, why don’t you just build what you know, and then put these other things in that you don’t understand? Leave room for it.” And it turns out that many of the things that they left room for were necessary, but they were just not understood by the average person. For me, that was the work of genius.

Zierler:

I was just going to ask—genius gets thrown around way too often. It sounds like Dicke, as you describe him, was an unqualified genius.

Weiss:

He was an unqualified—no doubt. And he was later considered that by almost everybody. But the remarkable thing about it—how does a standard person do that kind of thing? A standard person designs the sort of things that are absolutely necessary, builds it, and then find out, “Oh my god, shit, it doesn't work!” And then you carefully look why doesn't it work, and you make one modification, the next modification, and so forth. You eventually go through the same process. But you get driven to it by what nature is telling you, not by what your mind is telling you. So, to me, it was unbelievable. So, that was a big experience—that anybody could do that.

Zierler:

Rai, you talked so much about the singular influence of Jerrold on your life, and obviously that influence helped form you as a scientist. So clearly Bob did not need to have that kind of role in your life, because you were very well formed. But I assume that he was, on some level, a mentor for you. So, I wonder if you could—

Weiss:

No, he was a very bad mentor. That was one of his problems.

Zierler:

He was a bad mentor.

Weiss:

He had favorites, and most of his favorites were people who were his students. I was a postdoc and sort of outside. I mean, we talked. It was nice enough. Look, I fixed his grand piano for him, okay? (Laughter) That’s another story, but we gotta go. So, we got to be buddies about that. But the thing is that, no, he was much closer to other people.

Zierler:

Because you were an outsider, essentially?

Weiss:

I was effectively an outsider, yeah. And he was at times—not prickly, but very hard to understand. I took a course in general relativity or tried to—that will come in a minute—from him. I sort of couldn't understand what the hell he was talking about. And then I took one from a guy named Wigner, who was a very famous mathematician, and I couldn't understand anything what he said! Because it was all mathematics and group theory. So anyway, that was a failure. I never really learned general relativity as a topic, at Princeton. That was one of my aims.

Zierler:

But as an experimentalist, you learned a great deal.

Weiss:

I learned a tremendous amount. And that’s something—I learned all sorts of new techniques from the people who had done things. And Bob was the guy behind all of that.

Zierler:

What was Bob’s style in experimentation that you learned from? What did you learn from that?

Weiss:

Well, his idea was fundamentally—and that comes over and over again later on—that you can do mechanical experiments. You can make a mechanical experiment, like do a pendulum, or do a thing which has wiggling parts in it. But what you do is you never let it move. You put a servo system around it, in other words a system that holds it in a fixed position, and then look at the signal that’s needed to hold it in a fixed position. You get the idea? It’s a quite subtle and very elegant idea. It’s the idea of using feedback to keep something from moving, but you want to find out how much force was there to try to make it move. And that’s what you're measuring. But that’s a way of doing all sorts of delicate experiments in a way that you don’t get into trouble with the things swinging around all the time, or being driven by noise that you're not controlling, and stuff like that. It was a powerful idea. In fact, all the way the next—in my life, and a lot of other people’s lives—were influenced by that. I mean, the fact that we succeeded detecting gravitational waves very much depended on that concept.

So, the thing is that was one thing. He knew electronics very well. He ran a company of his own that made certain kinds of devices. I learned a lot—see, at that time, there was a revolution going on in electronics, too. The transition from vacuum tubes, which I knew well, because that was my whole shtick when I was in high school and in college. But now, the transistor had come along, and the integrated circuit was about to come. So, he was at the forefront of all of that. And so a lot of the modern electronics that we now have was being exercised by Bob Dicke right away. And I learned it there. So no, he was a very, very important person in my life. Not that he came and taught you technique; you just saw what he did, and you marveled at it.

Zierler:

And to do a little bit of foreshadowing here, is this really—not necessarily specifically, but generally, in terms of what you're learning, is this sort of one of the intellectual origins of LIGO? Does it all sort of trace back to this work?

Weiss:

A lot of it does, and I'll point that out to you when we get there. Yeah. In fact, I would give Dicke the most credit of anybody for the—it’s not the only thing—but for the success of LIGO. In my life. Other people might not agree with that. Because by the time LIGO got going, these ideas had permeated the experimental world, and many people were trained on the ways—they didn't realize that what they were doing was they were using ideas that Bob had invented or Bob had made popular. But in my life, since I'm generally the oldest person in groups like that—and that happened to be the case over and over again, that I had that connection to Bob, and they didn't, and that make a difference. But I attribute it to Bob, much of it.

So okay, I think that’s the Princeton years. My wife contracted MS when we were at Princeton. That was a big step and big misery. And then there was the whole question of whether to have another child. That was important. We had one. Should you have another? That took a little bit of studying and stuff like that. But we came back to Boston at the insistence of Jerrold. And I don’t know—

Zierler:

Did you stay in close touch with Jerrold while you were in Princeton?

Weiss:

No, not really. We would occasionally talk on the phone, but he was so busy with his revolution in education, and he was more and more dropping out of physics directly. So, things to him that were important were could I help him—it was mostly sometimes things about they were trying to build a gadget to show a movie of a certain principle in physics. A lot of the things that Jerrold was doing in those times was he had made a system approach to education—I won’t go into it deeper than that—and he needed things like books to be written, problems to be given to the students, and movies. A lot of what he wanted to do, he used movies as a way of showing. And I remember him calling me on several occasions about “If you wanted to make a movie to show the pressure of light, how would you do it?” Things like that.

Anyway, so I came to MIT, and the very interesting thing—it was in 1964, ’65—the beginning of ’65—and it turned out I didn't have to get any money. And this is a very important piece of my life at the time. I had never written a proposal. I mean, that’s unheard of nowadays. But this was still a relic of the Second World War. What was happening—Princeton had this, Columbia had it, Berkeley had it—a lot of the big research universities had what’s called the Joint Services Program. Do you know about this or not?

Zierler:

I don’t. No.

Weiss:

Oh, wow. Okay, good. Because this will interest you, I think. And what it was was that—this was all a result of the Second World War and the fact that by the end of the Second World War—and Jerrold told me a lot about this—they couldn't populate the labs anymore. For example, Harvard was running underwater sound as a thing for the Navy. MIT was running radar and doing all the radar development for the Navy and the Air Force and the Army. And then of course the big bomb project, which was going on in Los Alamos. And Johns Hopkins was doing proximity fuse. Everybody who had any talent in science and engineering was working on the war effort, intellectually.

And it turned out that by the time the war was over at radar lab, they couldn't find people anymore, and they would hire psychologists or they would hire maybe even historians, okay, but anybody who could think straight, and teaching them enough physics so they could do the development of the radar. And that was a very complicated business, because they really had run out of engineers, physicists, electrical engineers. There were no more left.

And so, a very intelligent man in the Naval Research Lab named Emanuel Piore—P-I-O-R-E—had come up with the idea that what the military should do is they should have a program to put money into education of scientists and engineers. And don’t just give them the money; they had one requirement—to graduate them. They didn't give a damn what they worked on. And MIT at that time was still being supported that way. So, what I was going to do is I was going to start a whole group on two topics, cosmology and gravitation.

Zierler:

And Rai, this is like what? A second postdoc? This is an assistant professor?

Weiss:

I was an assistant professor. I had one postdoc. And yeah, it was in fact really quite a deal. I was invited to come back into the lab, what’s called the Research Lab of Electronics, which is the lab that Jerrold’s lab was in. Over—sort of supervisory lab. And I was allowed to make a new group. And they gave me some number—I don’t remember the dollars; it was unimportant. I could take a couple of graduate students—no other faculty yet—and a couple graduate students and as many undergraduates as I could get my hands on and do something interesting.

Zierler:

Were you able to work with Jerrold at all?

Weiss:

He was no longer part of this. He had an interest in what I did, but he was no longer—it was maybe once every couple of months we would see each other, or something like that. So, what I decided was two things. Since I didn't have to worry about money and writing proposals—I got this money by being part of this research lab of electronics—I thought of only, “Let’s do some very interesting long-range programs.” And one of them was—in fact, the first one was to try and see an effect that may or may not be there, which was a prediction of a couple theories of gravity, which was to see, does the Newtonian constant of gravity—G, big G—the strength of gravity, does that change with time? That was a project I laid out for myself.

And what it required was measuring—the way I thought we would do it is by measuring the strength of gravity g (the attraction of a mass on the surface of the Earth to the Earth) on the ground, and measuring it very carefully, but measuring it in many places around the earth. It was an ambitious project for an assistant professor. And then also measure whether the Earth changed shape. Not only measure g everywhere but see if the Earth changed shape by using lasers as a way of—using the wavelength of lasers as a way of measuring between positions in the ground, whether the Earth was expanding or contracting. Because that would also affect g. So, two things would affect g—mass coming in from the outside, which does come—dust comes in from everywhere—and mass also—atmosphere goes away from the Earth, so you want to keep track of it. And the other thing was the expansion or contraction of the Earth itself (Note: changes in g when corrected for these effects can be related to changes in G the Newtonian constant).

So, this was a whole program with new instruments, some ultra-stable gravimeters, very stable lasers, all sorts of stuff that had not been done before. And so I got very deep into this. And what happens is that—a change happened. And the change happened, which is the same one that you studied, namely that the Vietnam War had—the corruption from this Vietnam War had screwed up the society. And it’s true, but I'm not—there yet—I'll come to that in a second. What that leads to is that it changed the whole business of this program that the military was running. It killed that program. We'll get to that. But there was another thing that happened before that. That’s very important in the story. I was doing this work, and there was some progress in it—making the instruments, not doing the final thing to see if big G was changing. And I was asked by the department chairman whether I—it wasn’t an ask; it was a demand. He said, “We need somebody to teach general relativity.”

Zierler:

Uh-oh (laughter).

Weiss:

(Laughter). And that was a real problem. Because I couldn't admit that I didn't know it. Here is a lab dedicated to gravity and cosmology, Okay?

Zierler:

But would you ascribe to the idea that the best way to learn a subject is to teach it?

Weiss:

Well, you'll get to that. You're way ahead of me! What happened is that (laughter)- I couldn't lie that I didn't know it. Or that I—you know. No, I mean, it’s not a lie; it would have been the truth. But I lied as if I did know it. That’s the other way around. Gotta be careful. And so, I didn't make a big objection to it. Let’s put it this way. Because I felt I couldn't. And so, what happened—just what you said, namely I started a new course, which hadn’t been in the physics curriculum for thirty years at MIT—it had been in the math curriculum, but not in the physics curriculum—on the general theory of relativity.

But mostly first were the things I understood best. What was the experimental basis of gravity? What was the experimental basis of general relativity? In other words, turn the thing around. The things I knew best, I would start with. That gave me time to learn the mathematics. That was an approach that worked fine, and the students—and they were mostly graduate students—began to realize that I was not a bullshit artist, but there were things that they could ask me that I couldn't answer, and I would tell them, “I don’t know the answer.” And that didn't lose me too many students it turns out, which is sort of interesting. They were willing to play this game.

And the place where it finally hit hardest was in a place—I had learned enough tensor analysis to explain why the theory required tensors, how they actually curved space and time so that the theory looks like gravity. I got that far. But I couldn't do anything subtle. And right in the middle of the course, a very interesting thing was going on. And this is where Joe Weber comes in. Joe Weber—you're probably only a thousand feet away from Joe Weber’s craziness. Do you know that? I mean, you're in College Park, right?

Zierler:

Yeah (laughter).

Weiss:

Do you live there, or not?

Zierler:

I'm in Silver Spring, but close enough.

Weiss:

Well, when you go to work, you're very close.

Zierler:

Right.

Weiss:

And you haven't gone to work for a while. And you ought to go look at it. There’s a garden which has now been dedicated to what I'm about to tell you. I'll get back to the military thing in a minute, because this plays a role only after this. What happened was that Weber had gone to that meeting in North Carolina, the 1957 meeting, and he and Wheeler, a very famous theorist, had concocted an idea of how you might detect gravitational waves, of Einstein’s variety. Not of the Dicke variety. Their idea was, if you take a great big bar, a big hunk of metal, and what a gravitational wave will do is it will come through that, and it’ll stretch the bar a little bit in one dimension, and compress it in the other dimension, both dimensions being perpendicular to the direction in which the gravitational wave is moving. In other words, it’s a transverse wave, but it stretches space in one dimension and contracts space in the other. Can you imagine it? Okay. And it affects clocks. It affects space. I mean, all of these things get moved. But the big effect, if you look at it in one particular metric, was that it would stretch space and contract.

And he exploited that idea by saying, “Look, why don’t we make a great big aluminum bar?” Your size, the size of a person. And weighing much more than you, of course. And then the idea that Weber had was to mount little sensors on it to see if a gravitational wave went through and stretched it and starting it singing as though it had been hit by a hammer. So, it would be like a xylophone. You pop it and it sings on, due to gravitational wave excitation, after that. That was his idea and Wheeler’s idea, that you could make a measurement of gravitational waves.

And it turns out that he had, by 1964, set up three of these—one in his lab at University of Maryland, another one on a golf course that’s somewhere near the college eight miles away from it, and another one in Chicago at the Argonne National Laboratory in Chicago. He had three of these things operating. And the students in the course had heard about this. I didn't know much about it. And they asked me, “How does Weber hope—how does a gravitational wave actually interact with a big bar like that, to do what he’s hoping to see?” And I had a hell of a time explaining it. In other words, it’s the actual mechanics. How does this theory, which is pure geometry, a geometric theory, make forces that pull the bar around? That to me was—from the little I had learned of general relativity, completely the wrong way to think about it. Because the way I had now learned general relativity in the process is that it’s the properties of space and time that get changed. There are no forces. Gravity force is gone. There’s no more gravitational forces.

So, I spent a weekend thinking about how in the hell I was going to explain this to the students. And I said, “Here.” I come in—and this is important to the future [of this story], Okay? I said, “Look, let’s imagine you have a clock. A very good clock. And you have one clock here, right there, and you have a laser, which was coming in, and you send light from a laser to another mass over there and reflect it back again. And imagine the time it took the light to go from here to there and back again. Now, if space is being stretched by a gravitational wave, you'll see that time change on a good clock. All the measurements are being made in the same place. So, there’s no question of making measurements in different places.” That’s always a problem in general relativity. Everything is done in one place. And so that was my explanation. “Here’s how I would detect gravitational waves.” It’s a gedankenexperiment. You know what a gedankenexperiment is?

Zierler:

Yeah.

Weiss:

It’s a thought experiment. So there was no clock anywhere near good enough to do this, but the basic idea was send light, start the clock, send the light down to the other mass, have it come back, and stop the clock when the thing came back, and see if that stays fixed. And if it doesn't, it means a gravitational wave came along between these things and stretched space. Okay, well, that was a problem I could give to them as a thing they could do themselves. How to calculate that. It was a fairly straightforward problem, and I gave it as a homework problem, after having explained it. Every kid in the class was able to do it, even—and I felt very proud of myself that here was a pedagogic scheme for explaining something.

But I couldn't explain the Weber experiment where there was a force on the bar that pulled it apart. I don’t know where that came from. So, all right, we then go on in the course to cosmology, which was an absolutely beautiful topic. Have you looked ever into a little bit of the theory of the universe as a whole?

Zierler:

Yeah, sure.

Weiss:

It’s a wonderful topic. And we all fell in love with cosmology. So anyway—

Zierler:

Rai, how new did cosmology feel to you at this point?

Weiss:

To me, it was all brand new. I had never studied any of it. And the thing that was about to happen or had just happened—the course was in 1966. In 1965, the group at Bell Labs had discovered the cosmic background radiation, which is that radiation that comes from the big explosion that’s associated with the start of the universe. And by the way, the Dicke group, after I had left, wanted to do that experiment, to actually measure that. That’s another place where Dicke becomes very important. And to me, it’s an outrage that Dicke never won the Nobel Prize for this. In other words, Penzias and Wilson, who were at Bell Labs, won the Nobel Prize for this thing. They didn't understand what it was that they were measuring, but they measured it, and they showed that—and there’s endless lovely stories associated with that measurement, not only their measurement, but the engineers who made the antenna that made the measurement as well. Maybe that’s a story to tell you a little later, though. Because it’s fun. That’s a fun story.

But the thing is that Dicke and his group was trying to measure this. They had wanted to go ahead and measure it. And well, it turned out they learned from a mutual acquaintance when they were starting the experiment that the Bell Labs people had already effectively done the experiment, although they didn't know what they were measuring. And so that got together in two journal articles, and that was a revolution in cosmology. But that was not the only thing we talked about. That was one of the things. And that, by the way, became very important to me later on in my life. But the important thing was that I had never looked at the whole idea of the universe being a fluid, and you could calculate all sorts of properties of it, as a single system, it turns out much simpler than many of the other systems you can study with the general theory of relativity. So that was a real charming end to that course. I learned an enormous amount out of doing that.

All right, the next thing that happened was that about 1969, Weber makes a huge announcement. This is three years after that course. He finally pulls all his data together, announces that he had discovered gravitational waves. And it didn't take long for everybody else to start doing that experiment, a lot of people, all over the world, probably I’d say ten groups of people. It was not such a difficult experiment to get one of these aluminum bars and mount sensors on it and sit there and listen to it. And by the way, the thing that you will see if they ever open the doors for you again—when you go to the University of Maryland, you'll see a garden of all of the different Weber bars that were—they're put in the ground like trees. And it was quite a ceremony. I came to that.

Zierler:

Oh, wow. Now I'm excited to get back to the office!

Weiss:

(Laughter) Yeah, it’s really quite fun. They had to do it, because people didn't know what to do with all that aluminum. They began to throw it out. And here’s this heirloom that that university is very proud of, because he started the field of gravitational wave research. Anyway, but the sad part was that by ’71, it was very clear to almost everybody who had been doing these experiments that they were seeing nothing. In other words, a lot of people had followed Weber. They did what he had done. Some people did it other ways. Some people did it in better ways. They understood their apparatus better. But the basic idea was still, look at a resonating bar and see if it sings after it has been hit by a gravitational wave. That was basically the idea. The differences were only in how you made the measurement of the bar singing, where you might put the sensors that looked at where—I mean, these are subtle differences. But fundamentally, even the people who did it other ways and better ways did not see anything. Nobody saw anything.

Zierler:

What was the problem? What was the limitation there?

Weiss:

Well, that’s a good question, because nobody—up to now, we still don’t really know. We now know that Weber couldn't have seen anything, and let me tell you—the real problems were that—in fact, at the time, there were all sorts of reasons—some of the most famous people in science thought it was cheating, that he was cheating, that he was making it all up. And I felt that that was not a fair thing to Weber. I knew him pretty well by that time, and I felt that he was a very imaginative man. He was not in the catalog of—as Dicke. Not even in the same group as what I would call Jerrold. He was a different kind of person, but not in the same—he was not of that ilk. But he was just not a careful enough experimenter. That’s what I felt. But he was very imaginative. I mean, imaginative guy, came up with all sorts of wonderful ideas. In fact, as a scientist, he came up with the idea of making a maser, which is the predecessor to a laser. He wrote it up but then never built it. And he came up with a lot of different very interesting ideas. So, I can tell you what I thought was what he was seeing. I thought he was seeing magnetic pulses. But look, let’s not get into that. There is nobody that accepts a real theory of what he was doing differently than others.

Zierler:

But you're saying that really there was nothing for him to actually be looking at.

Weiss:

No, that’s not true. That’s not true. Now, we know that he was nowhere near sensitive enough to have seen anything. But that was not known at that time.

Zierler:

So, the problem was with the instrumentation? That was the issue?

Weiss:

The problem was with the instrumentation and the frequency he had chosen and the sources he was thinking about. In other words—I mean, the problem was already known that there was something funny going on, before other people could not find his signals. And I'll tell you this much. If you took the signals he was seeing and calculated using the Einstein theory—using the Einstein theory of how much energy was in those gravitational waves—and he had a sort of very crude way of saying, “Look, most of these pulses that I'm seeing come from the center of our galaxy.” And that was not a bad guess. But there was enough information allegedly in his data to tell him that when the bar—here’s the bar, and when it’s perpendicular to the bar, it faces the galactic center, it stretches the bar better than when the bar long axis faces the center. In other words, there was a directionality to the detection and his instrument. And so he came to this guess. And then people said—some very famous people, one of them being Freeman Dyson, who you may have—have you ever talked to Freeman, or not?

Zierler:

No.

Weiss:

That’s a shame. You would have had a good time.

Zierler:

I have no doubt.

Weiss:

But Freeman did a little calculation and said, “If Weber’s right, he has made a very special thing about the whole universe all coming from our galaxy. You can take all the rest mass energy, all the energy in every star, all the electromagnetic energy, every piece of gravity, everything, and completely eliminate it in a million years.” In other words, with the rate that Weber was seeing, he was seeing enough pulses so that—that they carry so much energy, these pulses, and they do very, very little, but they carry an enormous amount of energy, that the amount of energy that would have been carried away from the center of our galaxy would be enough to have converted every piece of mass into gravitational waves in a million years. And we know the galaxy is ten billion years old. So that was already noticed; there was something funny going on. But you couldn't say—maybe now was a very special moment in the history of the galaxy. God knows. But the big problem was it was not seen by anybody else.

So, what I decided—I was already busy with this crazy experiment to measure G changes—that I would look at that gedankenexperiment that I had in that course, and maybe I could convert it into a real experiment. In other words, do what Dicke had done: think it through, and could you actually make a device, use that principle of the timing of light to measure the existence of a gravitational wave? The gedankenexperiment couldn't do it. That was just not possible. But on the other hand, if you made it so that you could take differences in time—for example, if the gravitational wave is coming at you and it’s stretching space, compressing space in this dimension and expanding it in that—well, if you made a thing that looked at the compression and the expansion simultaneously and took differences, you could probably not need a superclock. You could do it without measuring—you could take differences in time. And that’s a lot easier to do than measuring the absolute time.

And that was the beginning of it. And then I realized that, oh my god, you could make this bigger and bigger, much bigger than a bar. Because the wavelength of—the bar’s size is limited by the velocity of sound. You can only make it so big. But you could make these things the size of—the velocity of light. So, you could—hundreds, a thousand times more sensitivity than what Weber had, if you wanted to. And so very, very clear to me that—and I spent a whole summer, I think the summer of ’71, sitting in a little room at MIT, away from everybody, calculating—I didn't realize I was doing what Bob Dicke had done, but not as well as he did. That’s why I had to tell you the Bob Dicke story. I was trying to think through what all the different noises were that might affect this measurement. Came up with a great big list, and then drew up an apparatus as a prototype.

Zierler:

Did you do drawings like Bob, also?

Weiss:

No, I had one drawing (laughter). It was mostly calculations. And then I came in, and here’s where the interesting thing comes. I wrote it up—and here’s the other part that’s a little wacky. I wrote it up as a progress report for my laboratory, sort of a twenty-five-page thing with all these calculations and stuff, and this concept for making such a detector using interferometry. That’s the name of the fancy word that goes with it. And I put the concept in a progress report of the laboratory, but never published it. It was never published. In fact, my whole life—and this is another piece of me that’s very strange—I never published enough. It recurs and recurs and recurs. I mean, I never thought of—if it wasn’t a completed experiment, and if it wasn’t finished as a completed experiment, I wasn’t going to publish it. A lot of the earlier stuff was never published. And I learned that from Jerrold.

Now, that was maybe a bad lesson. But to him, his argument was the following: suppose you have an idea about something, and you publish it, and you don’t do it—I mean, you don’t vote with your feet to do it yourself. Well, then, why should you try to get any credit for it? The guys who do it—or the girls and the guys who do it, the ones who slaved over this thing, they deserve the credit. Let them publish. They don’t have to refer to you for the idea. And I felt very strongly about that. Now, that’s not the way the world works (laughter) but okay. And that has caused a lot of trouble as life went on.

Anyway, so I went to the head of my lab, that lab, the Research Lab of Electronics. I showed the guy who was running it, who was an electrical engineer at the time, I showed him this idea. And he was charmed by it. Not because he was a physicist; he just liked the idea of all this technology that would have to be developed to make this go, and he thought some of it would be great. Making these highly sensitive interferometers and making these highly good vibration isolation systems and all this stuff that was needed to do this would be just the thing that he thought that lab should do, his lab should do.

And so, I asked him for some money, and he gave me what sounds like a lot of money then, but not so much now—it’s about $50,000—to go build a prototype. I had laid out a prototype. And we had gotten some distance into it when the axe fell, and that’s why I told you about the military. Because what happened is that—this is now in ’73—it’s around ’73. I don’t know if you remember Mansfield. You do? You remember Mike Mansfield?

Zierler:

Yeah, of course.

Weiss:

Oh, yeah. Well, anyway, he had tried already in the sixties to try to get rid of the Vietnam War. And there were two Mansfield amendments to the military procurement bill. The first one—I don’t know exact the date, but it’s under Nixon’s—it’s Nixon’s time. I think its Nixon’s time. I'm not sure. It’s late sixties. okay, you probably know better. Who’s that? Is that Johnson already?

Zierler:

Yeah, Johnson until ’68, and then Nixon comes in, in ’69.

Weiss:

Okay. It’s at Nixon’s time, the beginning of—he writes an amendment called the first Mansfield amendment. It says, “Let’s declare victory and get the hell outta there.” Well, that didn't work. And the country got more and more a mess. I mean, talk about revolutions. You know, people coming from Harvard wanted to burn down my building, for example. The old building I was in. And in fact, I would have joined them (laughter).

But the thing was that the whole country had gone nuts, so the left—the right was pissed off at scientists because they didn't support Nixon. And Nixon closed off all scientific advice—I don’t know if you know that—whereas Eisenhower had a physical sciences study committee, which included most of the American important scientists like Jerrold and Wiesner and others. He listened to his scientists. None of them would vote—none of those guys would vote for Nixon, and he knew that, and he abolished it. Very much like Trump. And what happened was—so he got no advice from anybody, and so the scientists got pissed off at the war, and him. In the meantime, the right got pissed off because he wasn’t being aggressive enough. That was another problem.

So, it turns out there was a confluence of the right and the left to do what I'm about to tell you, which was Mansfield’s second amendment, which was that the military should not support research unless it’s in disciplines that’s important to their interests. In other words, engineers and scientists should not be supported for science and engineering that was not relevant to the military mission. And they didn't define what the military mission was, but that was locally interpreted. So, for example, at MIT, it was interpreted as that we didn't want to jeopardize the money coming in by the Joint Services Program, so they played it very carefully. They got rid of all the things that didn't look like might have application that they thought. So, a lot of solid-state physics was brought in. A lot of—mostly solid-state physics. A lot of what I would call computer electronics was starting. Things that looked like they would apply to the national defense. And things that had to do with pure science were eliminated. And one of the things—my group got eliminated. So, we got no more support.

Zierler:

Did that end the research, or you were looking for ways to keep it going with other sources of funding?

Weiss:

I was too naïve to realize this was going to hit us, and I started too late. But I had never been able to hire new people into this—never, for example, a faculty member.

Zierler:

Did you feel, Rai, that it was unfair that you were swept up in all of these things?

Weiss:

I felt so angry about the Vietnam War that I almost let anything happen.

Zierler:

Right.

Weiss:

I felt a little injured by it. Yes, of course. Because radio astronomy survived, but why did it survive? It survived because they were making receivers for the radars. The thing I was in was already in trouble, with MIT itself. That was the thing that you have to—see, cosmology and—so two things happened because of that. Gravitation was not considered relevant to the military, absolutely. And cosmology was not considered relevant either, but if there was communications aspects of it that involved—maybe they would contemplate it.

So, what I did was I did two things. And this is a little bit out of order; I should have left something out, but that’s okay. I had already gotten into trouble with MIT, the physics department, by the time this all happened. In other words, they looked at my program. It’s a very long-range problem looking at how G changed, and now this new program of looking at gravitational waves. And they said to me, “You'll never get tenure out of this. You're too far into it.” And the guy who was head of the division said to me, “If you continue in this, you can forget about it. You'll never be able to stay here.”

Zierler:

Wow. Why did they say that? What was behind that?

Weiss:

Because I wasn’t publishing a thing! Nothing scientific. If there was any publications going to come out of those two things, it was going to be publications that had to do with the technology. And that was appropriate for an engineering department, maybe, but not for a physics department.

Zierler:

What was the issue? Were you just so deep in the research that you didn't feel like you had time to publish? Or did you want to wait to publish until you really had something to say?

Weiss:

Well, both. And I never—

Zierler:

Rai, let me ask a naïve question, then. Those seem like two very legitimate reasons not to publish.

Weiss:

Yeah, but you can’t do that, in this world.

Zierler:

(Laughter) Okay.

Weiss:

(Laughter) I mean, it wasn’t even a matter of stubbornness. It was a matter of I hadn’t considered that that was a necessary thing to do. It was a naivete of a very deep sort. Anyway, the reality came in, and a guy whose name was Bernard Burke suggested to me, “Look, why don’t you put those experiments on the back burner until you're tenured, if you ever get tenured, and do something else? You said you're interested in cosmology. Aren’t there things in cosmology that you could publish?”

And there was a thing. And one of the things that had come out of my course that I taught was a very good student who had joined me, and he was deeply interested in cosmology. And we decided together that maybe we had talked about this a little already, but it never was pushed as hard as after that dictum had come down. See, I backed off a little bit. I'll have to tell you—there was—this little piece was missing, before I told you about losing all the money. This happened before I lost all the money. I already had this trouble within the department.

So, what a guy named Dirk Muelner—who was one of the best students I ever had—and I did is we looked at why—because the cosmic background radiation had just been discovered, probably three years before. Yeah, ’68. And maybe ’67? Or two years before. And one of the important things was, could you show that it was really thermal radiation? In other words, a definitive sign that this cosmic background radiation, which had been discovered by Penzias and Wilson, and wanted to have been discovered by Bob Dicke and Dave Wilkinson, if that was a thermal spectrum. A thermal spectrum has a specific shape to the amount of energy at each wavelength or each color. And the measurements were not doing that. The measurements were all at low narrow band of frequencies, and the question was, could you come and see, for example, what a thermal spectrum is supposed to look like? A thermal spectrum is supposed to get brighter and brighter and brighter as you go to higher and higher frequency, and then it’s supposed to turn over. There’s a maximum. So, there’s a shape of a curve for a thermal spectrum. And the idea was, well, couldn't we show that the thing turned over? That the thing had a peak at the right place?

And the trouble with that was one had to do this from either a satellite or do it from a balloon. Using balloons, going high up in the atmosphere. And Dirk Muehlner and I decided that we would try to build an apparatus that could measure this from a balloon. So, we had already decided on that. We had already—and that’s where I had gotten some money from the laboratory to do that. That had also been killed when the Mansfield amendment came along. But I had already started looking for money for that from NASA. So, I wasn’t right when I told you I had never tried to get money. The money was right after that—right before that, I had tried to get some money from NASA to do what’s called the cosmic background experiment measurement. And then within a couple of months after that, the axe fell for everything, for the other stuff in the lab.. And actually, I then wrote a proposal to keep that going, to the National Science Foundation.

Zierler:

What was NASA’s response initially?

Weiss:

NASA’s response was very positive. And the reason why was at that time—now, they just celebrated her with giving her the name of the—I think they called her—the Hubble telescope after her. Yeah, Nancy Grace Roman. Is that a name that means anything to you?

Zierler:

No.

Weiss:

Okay. So now we go on to NASA a little bit. NASA was in its heyday. This was about 1973, about then. ’72, ’73. Wait a minute—no, that’s wrong. The beginning of the ballooning started earlier than that. It started in the late sixties, and then continued into the seventies, and ends at the eighties. Let me make sure the sequence is right. The course was in 1966. Dirk and I started the ballooning stuff in about 1967 or ’68. The gravity wave antenna started about 1971. And then the axe fell in about ’73, ’74. That sequence.

So Dirk and I had already, by the time that the axe had fallen, made a balloon flight with NASA support, and we had discovered that it did look like the spectrum looked thermal, but we weren’t able to do it as well as to really be out of question. There were too many things that still existed as questions, and they had to do with what the atmosphere did to the measurement at a balloon altitude. And a balloon altitude is about forty kilometers up. You're up at about just at the edge of the atmosphere. One way of saying it—the pressure of air at that point is one torr, about. One millimeter of Mercury. So, you're over most of the atmosphere which is 760 millimeters. The universe—space looks black to you, when you're at that altitude.

Now, you don’t fly with a balloon. I don’t know; maybe I've gone too fast. You send equipment up there and then you don’t—yeah, Okay. And so, this equipment was kind of fancy, needed liquid helium to cool it. We had to invent all sorts of detectors. And we effectively showed, in three bands, that it looked like the radiation was close to thermal, but it wasn’t enough to convince everybody. So, we then had to make another program with a new apparatus which then did it better. But the final way it was ultimately done was later on in our lives, in about 1989, by the COBE satellite. And I'll get to that. That’s an important chapter in my life.

So, at any rate, the ballooning started. It was going. It got supported by NASA. And then it was taken out of the program that the military was supporting. It was no longer needed in that. So, when the axe fell, the ballooning program still existed. On the other hand, I wanted to continue the gravity waves stuff, and so I wrote a proposal to the NSF for that. And that took a long time to get the money for, and the reason was that a lot of people thought the idea was nuts, the idea of using light timing—not light timing, but the wavelength of light to measure the distortion of space. Because their argument was that they didn't understand enough general relativity, a lot of the reviewers. And what they thought is that when you stretch space with a gravitational wave, the wavelength stretches along with it, and so consequently, there would be no change in an interferometer’s output when a gravitational wave came by. In other words, the idea didn't work. And they would write that in their reports, and the NSF didn't know quite what to do with that. It turns out there was a guy at the NSF who plays a very important role in all of what followed. Have you ever heard of Richard Isaacson?

Zierler:

Sure.

Weiss:

How do you know him?

Zierler:

You know, it’s a name that’s out there.

Weiss:

Okay. Well, in fact, he was the real hero of LIGO. He was the hero of this whole field. And Richard, who had just taken over being the director of gravity at the NSF, which was a tiny little office compared to all the other offices, he decided that he would send the proposal to a bunch of people who were in Europe. And it turned out that the Max Planck Society in Munich, in Garching, they had a very good group there who had been making bar detectors like Weber’s, and they were doing the same thing a lot of other people were thinking about. They had put a huge investment of their time into checking on Weber, finding nothing. But then they decided if they could, they would like to stay in the thing they had developed, but if they wanted to do that, they would have to shift technology. In other words, a lot of people who were doing that, checking up on Weber, tried to make better Weber bars, more sensitive ones, by making them cryogenic, by making them cold.

One of the noises in the Weber bar is the thermal noise, the Brownian motion of the bar itself. It shakes, due to just the fact that it’s at finite temperature. So, what came of it was that they saw my proposal and they didn't want to go into cryogenics anyway. I mean, it’s very strange. What they did is they decided that—the head of the group called me up, which is a little bit out of what you're supposed to do if you review a proposal. You're not supposed to talk to the people who wrote the proposal, okay? And they called me up and they said, “Look, this sounds interesting to us. Do you mind if we work on it?” And so, I was somewhat taken aback, but at the same time I said, “No, by all means.” But they asked me, were there any people in my group yet who had been experienced enough that they could hire? And that wasn’t yet true, because I was not able to put graduate students on the thing, because it was not a project that was going to produce a piece of science. So, somewhat later, we did have an exchange of people, and that was very productive.

Anyway, they had a lot of money to begin with, and they had a lot of excellent people, and so they started working on a three-meter prototype of the kind of thing that I had just cooked up as a one-and-a-half-meter prototype. And within two years, they had one running, and in fact found all sorts of things that I had not thought of. In that Dicke-esque session that I had, I hadn’t been smart enough, okay? They found things that I hadn’t even thought about. So, they found out—and I won’t go into all the different things, but they found out that scattered light was going to play a very big role in it. Light that wasn’t going along the path that you had determined. And I had neglected that. That was actually a big mistake. But anyway, they solved a lot of the problems of this thing, and by about 1978—yeah—they published a paper showing the results of their three-meter system, and it looked very promising.

And by that time, I had gotten some money from the NSF through Richard Isaacson, and I was trying to build up the group into getting a postdoc, and allow me to take—first I wanted the postdoc so I didn't have to fight with the department about giving him a PhD, or her a PhD. But then I began to think of putting graduate students on it and throwing caution to the winds. You had to do it. You couldn't do it with undergraduates entirely. And by the way, I way I managed to put graduate students on the gravity program was by starting them on the gravity program and then putting them onto the balloon program to get their PhDs. In other words, they all were told this—“Look, why don’t you help us get the gravity wave detector working, but you're going to have to do your degree on measuring some new attribute of the cosmic background.” And most people were perfectly happy to do that. It wasn’t that many students, but some. And okay.

What’s next? We have some money. We have money from both the three-degree cosmic background program and the gravity program. And the next big development is in 1975. That was a watershed moment, again. Now, this is slightly a political piece. I don’t know—when you were a historian, did you have to get money from anybody but your job?

Zierler:

Never.

Weiss:

You never had to get research money? Yeah, OKAY.

Zierler:

Not at state. That’s one of the beauties of the federal government.

Weiss:

That’s the beauties—yeah, yeah, right.

Zierler:

It’s all there.

Weiss:

Anyway (laughter), but when you get money from the NSF, or you get money from NASA—there’s nothing wrong with what I'm about to tell you, but you have an obligation. And the obligation is if they need help, you have to give it to them.

Zierler:

Oh, you mean like a one hand kind of washes the other thing?

Weiss:

Well, it’s not as bad as that. The two examples are, once I started getting money from NASA, every time they had a committee that worried about cosmology, I was asked to be on it so they would not do something crazy. It’s a nice thing to do, but at the same time, it takes a lot of time. At the NSF, what happened is that—in fact, this was the NASA program was the one that got me into this sort of 1975 situation, which was that the lady—Nancy Grace Roman—who is this lady who is now being celebrated—she died, but who is being celebrated as the first astronomy and astrophysics division head of NASA—had asked me to do something for NASA, which was to look and see, with a group of people—they were willing to pay for this—they wanted to know whether there was things in the space program that could be done for the science of gravity and cosmology. In other words, what can the space program offer to that discipline? A perfectly reasonable question.

Zierler:

Was it not also a self-evident question?

Weiss:

No, it wasn’t. Not to the people. See, it’s an interesting thing—most astronomers don’t know anything about cosmology, in those days, and they knew even less about gravity. And she was an astronomer. And of course, the thing that NASA was doing that was scientific was mostly planetology and astronomy. They were doing less physics in space and stuff like that. So okay, I pulled a committee together, which consisted of Bob Pound, who had done the redshift experiment at Harvard by a tricky technique. He had gotten to be known in this. A guy named Peter Bender, who was at the University of Colorado, who was very interested in planet motions around the sun as a way of getting information about relativity. And Charlie Misner, who was a theorist, in fact, at University of Maryland. And he’s still alive, by the way. And I think all—no, Bob Pound is dead. Misner’s alive. Yeah, okay. And then somebody from one of their—a guy from Huntsville, who was a NASA scientist, whose name I don’t remember, but he was important in another experiment. Never mind. So, what happens is that one of the most important people in that business of gravity—it’s mostly gravity—what can you—directions gravity should be taking in 1975 was Kip Thorne. You have read any of his stuff?

Zierler:

Sure, sure.

Weiss:

Yeah, okay. I didn't know Kip, but I asked him to be on the committee, but he was too busy. But he did offer to give one day to it. He would come and give testimony.

Zierler:

What was Kip involved in, at that time?

Weiss:

Okay, Kip was deeply involved in several things. He was always interested in black holes. That goes way back to the early days of his interests. But the other thing—as a practical matter, since Caltech was associated with the Jet Propulsion Laboratory. And the Jet Propulsion Laboratory was interested in using the planetary system as a way of doing all kinds of different kinds of measurements, and one of them was actually to see, could you find out more about general relativity if you had better measurements of the motions of the planets? In other words, could you look for small deviations and perturbations in the motions of the planets and in the sun that would tell you about general relativity in a deeper way? And they were also interested in gravitational waves. That was another thing JPL was interested in at that time.

And so Kip, a lot of his group, what he was deeply involved with at that time was something called taking the Newtonian formalism, which is what NASA was using for spacecraft motions and—just Newtonian gravity—and adding to Newtonian gravity corrections that were relativistic corrections, corrections that would come out of the general theory. Now, you could do this in two ways. You could do this by grafting general relativity onto Newton, or you could do it by taking the general relativity and looking at very weak cases, cases which have very, very small gravity, tiny gravity. Non strong, weak gravity. And that’s the case—that’s the direction in which Kip went. Kip went and he developed a whole—a really very beautiful way of making it so you could test general relativity at different levels by looking at classical experiments like planetary orbits, especially planetary orbits, and perturbations in the orbits. And he did it by expanding General Relativity in terms of increasing values of v/c as a sequence of post-Newtonian corrections that are made. And that was a business he was deeply into, at the time.

He was also of course interested in the basic theory, and he developed much understanding we have of strong gravity—much of the original work on black holes was done by Kip. Not exclusively by him, but much of it. So, he came to this mostly because he was wanting to see for applications of what his post-Newtonian approximations could be used for, and what new things you could learn. So, he gave a lecture to the committee about that. But the important thing was what happened after the committee meeting that day (laughter). And it’s always what happens not at the committee meeting, but what happens in the elevator or wherever, as you know. And what happened is he couldn't find a hotel room. It was the middle of the summer. And he had not gotten a hotel room in Washington in the middle of the summer, and that’s a big mistake (laughter). So, every high school kid in the country is running around—

Zierler:

(laughter) Right.

Weiss:

—as you know. So, he had a plane the next morning, and so we spent the whole night in my room discussing one thing only. He had already got one of the most prestigious groups doing theoretical gravity at Caltech, and he had talked the president of Caltech into starting an experimental group in gravity. And the question was, what should that experimental group focus on?

And so, we sat around a little table, and we kept writing, making tables and lists and pictures and god knows what. And finally, Kip said that he had already been thinking a little—anyway, we talked about what new things you could do in cosmic background measurements, what new things you could do in measuring large-scale structure of the universe, all these cosmological measurements. But then we got to what specific measurements could you do in gravity directly that were new. And could you, for example, measure the curvature of gravity? Well, that was an idea we talked a lot about. And we finally got to gravitational waves. And that’s the one thing he had thought about already. And he had come to the conclusion, with the help of a Russian whose name was Vladimir Braginsky, that if you were going to try and do a measurement of gravitational waves, you would be confronted by the quantum theory of physics. That would get in the way of making those measurements. You couldn't—if the measurements—you were measuring such tiny motions that you were going to be dealing with having to try to make a measurement at the level of a quantum system.

And that was kind of interesting to Kip, and he thought that a group that would do that would be quite amenable to what they did at Caltech. And he was thinking of that as a bar group, a Weber bar group, working on how do you make a bar so it’s limited only by quantum theory. That’s effectively what he was hoping to do. And I told him, “Well, that’s an interesting topic, but have you thought about interferometry?” And he hadn’t. And in fact, in his textbook, the big textbook that he wrote in 1973—I don’t know if you know the book. It’s gigantic. It looks like the New York City telephone book. It’s called Gravity. It’s all black, has an apple on the cover. Well, in there, in that book are examples of all sorts of things which Kip thought were bad ideas for how to detect gravitational waves, among them measuring light using timing of light to do it. Now, he had never seen my little—because it never was published, so he had never seen the thing I’d written. And for some reason or another, he had never gotten the proposal, either. Anyway, I didn't know he had written that negative thing about this in his book. But later on, I found out (laughter). Because I realized that when we started talking about this, he was really quite naïve about it. And it didn't take me more than about ten minutes to get him to say, “Oh my god” (laughter).

And so, after that ten minutes was over, he goes, “Well, who’s working on that stuff?” And so, I said, “Well, there’s a guy—there’s this German group that has now made a lot of progress in it. And then there’s a new group which I don’t know well, but the head of it seems like he has a lot of good ideas.” A guy named Ronald Drever, who was in Scotland at the time. And so, I told him, “Look, why don’t you investigate? See if either of the members of those groups are suitable for being a Caltech faculty member.” And he did that right off. And in fact, that changed the whole history of the topic, as it happened.

So, what happens next is that he finds that the—this is now ’75. By ’77, Caltech makes an offer to Ron Drever to come to Caltech to start a gravity group to do interferometry. And that was a complete breakthrough moment in the following sense. Caltech was willing to put—I don’t know the amount of money, but of the order of three to ten million dollars, somewhere in there—into making a new group, and the laboratory for them, and enough postdocs, and an assistant professor, to get into this field, which is considered at MIT like—MIT thought of it as a waste of—never gonna work, and also the science wasn’t interesting. That was my problem, okay? Well, at Caltech, it was completely different. Caltech, they were enthusiastic for this, and they were willing to make an investment. And that changed the whole character of the field.

Zierler:

Ray, did you think about just joining Caltech? Was that a possibility?

Weiss:

No, I didn't, but I can tell you that story, because that’s very cute (laughter). That’s a cute story again.

Zierler:

I mean, not a big boost of confidence you're getting at MIT, if they're telling you, “Forget this if you want to be tenured.”

Weiss:

Yeah, well, not only that, but they didn't think the science was—they didn't believe in black holes at MIT. At the time. Gotta be careful.

Zierler:

Right, right.

Weiss:

Things change. Administrators change. That’s the important part! So, what came of it was very interesting. Kip, before he made his offer to Ron Drever, called me up and he said, “Would you be interested in coming to Caltech?” So, I said, “Well, I'm reasonably happy where I am, but what does it require?” And he says, “Well, send me your CV.” So, I sent him my publications list and the CV. And then he sends me back an email. He says, “Well, I only got one third of what I think you were supposed to send me.” And so, I said, “What do you mean?” He says, “Well, you know, you’ve only published this much.” I don’t know—four or five papers. And he says, “What else can you tell me about yourself?” (laughter) And that was the end of it. In other words, my record was so terrible. That was the problem.

Zierler:

By the way, did Jerrold not tell you that you had to publish? Really, you didn't know this?

Weiss:

I told you, I was naïve, okay? Because to me, the important thing was get the thing done! So there was an attempt made, but I don’t think it was very deep. So, the next thing that happened was—there are sort of two parallel things going on. So, Drever gets hired by Caltech, starts a group. And two big things happen in my life that are the—maybe I should—I have to keep flipping back and forth between these things. I'm sorry, because it’s confusing, but yeah. And that is that I don’t know if you know John Mather. Do you know John?

Zierler:

Sure, I do.

Weiss:

That’s right! Of course, you know John, because you interviewed him.

Zierler:

John connected me to you!

Weiss:

Yeah, that’s right! Okay. John has graduated from Berkeley. This is a little before the ’75 meeting, but maybe ’73, something like that. ’72. It was ’72. But he already—and what happened is that he wanted to—the COBE satellite, he invented that. John did. And I think you probably learned a lot about that from him. I hope you did.

Zierler:

Oh, yes, I did.

Weiss:

And he wrote a very wonderful book about all the problems he had with it. The Very First Light or something like that is the name of it. Anyway, so he came to me, and since I was already established in the ballooning business and had published two papers on the spectrum, and I'm a good bit older than he is, he as a graduate student had cooked up the COBE satellite, and he was looking for people to join him. So, he came to MIT and told me about this, and told me about the whole idea that he had, namely you can make a satellite with those three different instruments on it. The three instruments were: an absolute spectrometer to measure the spectrum of the cosmic radiation, a set of four differential radiometers to measure the isotropy of the radiation at different frequencies and an instrument to measure the infrared to submillimeter background from all the galaxies in the universe. And I told him that it sounded interesting, but look, I would help him but he needed more people. He couldn't do this by himself. Not even with me. You needed more than that. So, I convinced him to talk to David Wilkinson. He may have thought of it himself, but I certainly pushed it very hard. And eventually, there was the COBE project, and I was part of that, with him. And that lasted until, Christ knows—I mean, I think it started in the early seventies, and it finally flew in ’89. And I'm not going to go through all of that, since John went through it with you. It went through I don’t know how many iterations. But my function in the end on that program was to try to sell Congress on it, sell NASA administration on it, and also, I was head of the science working group, the group of twenty people in the end.

So, that was all going on concurrently with what I'm about to tell you now. That went on until about the end of the eighties. And much of my group was mixed up in that, because it was hard to get enough people working on the gravitational wave detector until this thing with Caltech actually happened. And then all of a sudden, more people wanted to work on the gravity work, even at MIT, although it still was not fully accepted by the MIT administration. And I'll go on with that a little longer. God, we're almost again at the end. That’s terrible. What came next was that I made a rash decision in 1978, ’77, with the help of Richard Isaacson. Now, Richard Isaacson was by this time at NSF selling gravitational wave science as the next most important thing in science. He had convinced the whole physics division that this was an important thing for them to do. So, he and I strategized a lot together. We got to be really good friends, and still are. And what came of it was that I felt that if Caltech was going to go to do this, we had made some progress on the prototype, solved a lot of problems.

The Germans had solved—and they were going to go on now to build a thirty-meter system, which they were busy building a thirty-meter instead of a three-meter, so that you could show that the scaling laws for the detector worked. That was very important. So, they were working on thirty-meter. Caltech had made this big investment. Glasgow was sort of not the best, but it was making some progress. I decided the right thing to do now, if they're giving me this hard time at MIT that there’s no science in this, the best thing I could do is to work on—the prototype was running, but to make a big push to see what it would take to actually make an instrument that would measure gravitational waves.

In other words, it was a complete change of attitude. Namely, the prototype is working; I was not going to build the next prototype. I was going to see—I wanted to plan on building the LIGO. And I felt confident about that because of the fact that Caltech had given it a little bit of credibility, and also because Rich Isaacson thought it was a good idea to see, what might this cost? And so that was the next big step. And I got a bunch of money from the NSF, at the same time as the group at Caltech began, to study with industry, what are the problems of making a let’s say five-kilometer-long LIGO, with two arms, five kilometers long. Two of them separated by thousands of kilometers. Could you find places in the United States that could do it? That’s number one. How much might the whole system cost? And get industry to help you in doing the costing, and not worry so much about the detector. The detector was not yet at a state where you could freeze it. But you could certainly look at the design of the vacuum system, the architecture, worries about what the environment would do, stuff like that.

So, we did a study which lasted about three years, which resulted in a thing called the Blue Book. It’s a book that was intended to be a general book given to the NSF with all these facts and numbers and scaling laws and god knows. I mean, all the noise terms that were now known, what the sources might look like. But that wasn’t important. The most important part was, how would you build a thing that was sensitive enough, and these are the things that are needed. Here are technical things that are not quite ready. Here are things that are ready. And that kind of thing. And then a cost estimate.

And that turned into a thing with two companies—Arthur D. Little, which was a small company in Boston, here; and Stone & Webster, which was a big company that made nuclear reactors, and they had made very big systems. And that study would join—I got new people to come to the MIT lab. Peter Saulson came, who later became a very important person in this field. He became a professor at Syracuse. And then a guy named Paul Linsay, who was a postdoc. And Linsay—the three of us, with some help at the end from Stan Whitcomb, who was the young professor who was hired at Caltech—so we had these four people, myself included, who were writing this Blue Book and pulling together all this stuff from industry.

And so, by 1983, we gave the Blue Book to Rich Isaacson. And he used that, then, as a way of having us be part of the meeting of NSF big projects in science, in December of ’83. Three big projects were being considered by the physics division of the NSF. One of them was enhancement of the synchrotron at Cornell; another one was a new nuclear facility at University of Urbana, Illinois; and this thing called the Caltech-MIT LIGO System. Because what we had done before that, Kip and I, who had gotten to know each other, had managed to try to make a collaboration. Now, it was a very tenuous collaboration, because Ron Drever was not interested in collaboration.

Zierler:

Why not? What was Ron’s issue?

Weiss:

Well, that’s a long, long story. Ron was a very complicated man. Now that I know much more about him—and unfortunately, he died, just recently. But he was an Asperger’s syndrome person. I don’t know if you—have you run into people like that?

Zierler:

I have. Yeah.

Weiss:

Singular focus. Couldn't really get on. Was very sweet but not really—could never project himself into another person to try to figure out what was going on. So, he was fundamentally scared to death of a collaboration. I didn't know that. But that was the problem. And poor Kip, who was caught in the middle of it all, because it turned out that Kip didn't realize this at all. And Kip also thought it was essential to do a collaboration. Richard Isaacson finally convinced Drever. And the only way you could get to him—by saying, “Look, I can’t get the money for this unless it’s a collaboration.” So, “Forget about anything that—you're not going to get the money to do it on your own.” And so that was as blunt as you could get, and that got through. So anyway, what happened is we—Kip, Ron Drever, and I gave a report about the Blue Book and also the future of the field. Kip gave a wonderful presentation on the science. And what we got—I will send you this if you don’t—I think I'll send you this, rather than read it to you—

Zierler:

Please.

Weiss:

—the recommendation was unbelievable. And it was a recommendation that said, “Look, this is absolutely the right thing for the NSF to do, the LIGO. It’s risky, tricky, hard, but this is exactly the thing that might pay off in a dramatic way. It’s the thing in science that you want the NSF to do.”

Zierler:

Do you have any idea who was behind the letter?

Weiss:

I know every bit of it. I can tell you who—by this time, Rich Isaacson had got the NSF in the palm of his hand, and he had convinced a guy named Marcel Bardon, who was the head of—in fact, Marcel played a big role in getting rid of the Star Wars and a whole bunch of other crazy ideas we've had. But he’s dead now, unfortunately. But Marcel was head of the physics division at the NSF. And he got so interested in it that his son called me up—Oliver Bardon called me up—wanted to work on LIGO in my lab. I said, “Fine.” I felt a little funny about that. But in the end, he went into high-energy physics, because it looked like there was too much of a—too much time yet before this thing would become a reality.

Anyway, the guy who is responsible for the wonderful words is Stan Deser. I know that for a fact. And Stan was put on the committee expressly by Bardon and Isaacson to do this. It was a put-up deal. In other words, they wanted to make sure that they got a good recommendation. And Stan is a famous theorist in gravity and I don’t know, he was always interested in pushing something new in gravitation. Anyway. So, they got the right guy, and he wrote this wonderful thing; I'll send it to you.

Well, okay. So, we get through this meeting and the people at MIT had thought that we would fail, which was—that was their assumption. Because it was so risky and so completely—see, what they really complained about was the fact that the technology wasn’t ready in their minds, and secondly, that the sources were not known. You could build a huge thing, and it’d be a total failure. And they didn't want to be associated with a failure. And the Weber experiments had already exposed the field to failure, you see. And that wasn’t that distant in the past history. So, I can understand their motivation, but I'm not sympathetic to it. And in fact, the people who killed it at MIT were a very famous physicist, Francis Low, and a chemist who is still there, John Deutch, who you may know from his CIA days.

Zierler:

That’s right.

Weiss:

He turns out to be a total disaster. Everything he touches. I mean, everything! I don’t know of a man who has done more mischief in this world, than John Deutch! I can tell you endless stories about him (laughter). You must know him a little bit from Washington.

Zierler:

Of course, of course.

Weiss:

I mean, he’s a bull moose. That’s what he is. Okay, well, let me—I'll tell you this story right off.

Zierler:

Please.

Weiss:

This is completely out of context. This happened very much later, but since John, you know him—this is way into this thing, after MIT hadn’t yet put faculty into it. NSF was very worried about this. And Caltech was still putting a lot of money into it. MIT had not made any changes yet. This is sort of within the five years that we're talking about. And they were mostly distressed by the fact that NSF sent some people to MIT to see if they could somehow convince them that this project was for real, and they were not fooling around anymore, and they were going to go forward with it, and they shouldn't worry so much.

And I remember sitting in John Deutch’s office. He was provost, at the time. And on the left of me was the head of my physics department. On my right was John Deutch. Across the table were two guys from the NSF. And they come and they sort of very nicely go forward with telling John Deutch that, “Well, this is going to happen. It would be better if MIT really gets more interested in this thing.” And it doesn't go very far into that meeting, and John—who, as you remember, he’s about six foot three; I mean, he’s pretty tall—he gets up, he looks at me, and he says, “Give me a piece of paper.” So, I give him a piece of paper. He takes the paper, takes his pen out of his pocket, and writes a great big zero on it, and shoves it in the face of the NSF guys. “This is what we're gonna do!” And he walks out. That story is all over the NSF. I mean, in those days. And that was his contribution (laughter).

I'll tell you where the problem came with Deutch on that one, is that he felt the NSF—a lot of the people at MIT had done so much work with the military that had been very much more generous than the NSF. See, taking money from the military during that Joint Services Program was a piece of cake. They made no conditions. Very few conditions. On the other hand, NSF made all sorts of conditions. When you took money from the NSF, you couldn't pay certain things, you had to come up with corresponding amounts of money from your own institution, you had to pay your postdocs only a certain amount and so forth. A whole bunch of rules came with it. And the thing that bothered John was that all these rules would make it impossible, and he had been fighting those rules as provost for a while.

I don’t know; I'm giving him the benefit of the doubt. The real problem was that he and Francis Low, who was the provost after him—or no, before him, sorry—had come to the conclusion that this was nutty stuff to work on, and that was that, and that was the end of it. It completely changed, by the way, as soon as we got the next president of MIT, who is Charles Vest. Charles Vest was a mechanical engineer who thought this was absolutely the right thing for MIT to get into. Couldn't think of a better thing! (laughter)

Zierler:

And where is MIT administration? When do they start coming around?

Weiss:

They came around nineteen…well, they started coming around in 1991, and then they really came around when—the physics department got a new head also, who was interested in this field of gravity, named Edmund Bertschinger. That was a little later, maybe ’93. And then all of a sudden, they became just totally captivated by this. But it’s a shame that they went through this terrible—they got a very bad reputation with a lot of people in the field for that.

Well, anyway, let me take you back to what happened immediately after the 1983 meeting with the physics division committee meeting where we got this wonderful recommendation. Isaacson advised all of us—Kip and Drever and myself—to now try to make an organization of a project, in other words, and try to get a project manager. Because you had a lot of things to think about. You had infrastructure that you had to build, or things that were far more than just a few people who were working in those two groups both at Caltech and MIT. There wasn’t enough people to do this. You had to add people, and people who had experience with large projects. Isaacson saw that right away. But they didn't have yet the money. And this is the problem. I don’t know if you know how the NSF works. You know about the National Science Board?

Zierler:

Sure.

Weiss:

Okay, good. Well, the National Science Board had not yet been consulted on this at all, and the physics division was running sort of freewheeling on this a little bit. Although they gave it this wonderful recommendation—they had a price tag for it, but they didn't have the money—they had never gone to their upper level within the foundation to see if they could get the money. So, they were in a preparatory stage. And so, they couldn't yet put money into the project. They could give us money, as research groups, for things that were communal, like site measurements and stuff like that, but it had to be buried as part of the science. It was complicated. It was not yet a project.

And so, they said to both Kip and me that, “You've got to get a project manager, but we can’t pay for him right away. We will, once we get the money.” That was the problem. And MIT, I went to those guys at MIT—in particular, John, who was still—John was at that time dean of science, and the provost was Francis Low. They were about to shift. But I went to John Deutch first. I didn't realize he was deeply against this, but I said, “Look, we need some money for a project manager.” And he gave me the goddammdest runaround about that, that you can imagine. He told me all his bad experiences with physicists who lied to him when he was at the Department of Energy. I don’t know if you know this, but before the CIA job, he was the undersecretary for science at the Department of Energy. I don’t know how he ever got himself into all of those fantastic positions, but that’s a whole other thing. And he just told me that he wouldn't dream of giving money to this.

And very quickly Caltech, who had JPL—you know, Jet Propulsion Lab—they found a guy who would be a project manager, and they didn't even think about how to pay him; they just did it. And so, at that moment, the project left MIT, being run by MIT—not that I had ambitions for that—but it became a Caltech project. And the whole thing became also quite polarized. The NSF wanted to deal only with one institution, so Caltech became the primary, and we became subcontractors. And then [pause] what came of it—and this is the place where we failed, the group itself—the Caltech/MIT group failed: that is, we could not make decisions. See, Kip—you don’t know Kip as a person, do you?

Zierler:

Not yet. I hope to get to him, though.

Weiss:

I hope you do, yeah. Very, very sweet and gentle person. Wonderful guy. And he got himself trapped in a terrible situation. Here was Drever who was somewhat stubborn. And as I say, we didn't realize at that point what the trouble was. And then here was I who was experienced somewhat, because I had tried to get COBE together, had done other things. I had some experience. Had run committees for the government. So, I had some experience and I knew what was at stake, and I wasn’t going to give in to a bunch of crazy ideas. But there was no way to do it, because I was not in charge, and I didn't want to be in charge, but I couldn't go along with having the thing directed by Ron Drever, because he was just not reliable. And Kip was a theorist and the best he could do is talk to us. He couldn't be the person in charge of things, either. So, it turns out we made a thing which was unholy: a troika (laughter).

Zierler:

Right. Bound to fail.

Weiss:

And it failed (laughter).

Zierler:

Where three people are in charge, that means no one is charge.

Weiss:

That’s exactly right. You got it. And so, what came of it was the troika was a disaster, and the guy who had been chosen as project manager by the Caltech people tried to do something which was perfectly sensible, and he couldn't manage it. He wanted to get at least the two groups—the one at Caltech and the one at MIT—to at least have the same research plan. In other words, if they're going to build this big thing, you ought to at least have a plan that you divide the work up between each other, not try to do everything on your own. I was perfectly amenable to that, but Drever wasn’t. He didn't want to be controlled. And so, it turns out this project manager couldn't do very much. And what he did in the end is he worked with me a lot on trying to find sites and work on the vacuum system, on things that were not controversial. And we're going to run out of time again, I'm afraid. But let me tell you what’s ahead. We're getting near the end, but it gets—I will send you what I've written on this.

Zierler:

Please.

Weiss:

Be careful with it, because it’s something that’s being—let me tell you what this is for, but then you hold onto it on your own. What it is is that Kip, Barry Barish, and I are trying to write—and Peter Fritschel, a younger person, and a guy named Peter Shawhan, who you might know because he’s at the University of Maryland—are trying to write a technical history of LIGO. That’s why I can remember so much of it, because I've been just involved—

Zierler:

Right. This is wonderful.

Weiss:

Well, what I'm going to do is send you the chapter that I wrote, which is not yet by any means published. But it has a lot of the facts in it. But you can’t use this in any other way than in notes or something like that.

Zierler:

Sure, sure.

Weiss:

They want to publish it. I'll send you that.

Zierler:

Okay.

Weiss:

Okay? And maybe we should have another time. Why don’t you read it first, and then see if you want another time?

Zierler:

Oh, what do you mean? I know we need another time.

Weiss:

Okay, yeah. Okay. Well, let’s—I'm perfectly happy to do it, but why don’t you take a look at what I send you? I'll send it to you tonight.

Zierler:

So, why don’t we make this the cutting point for today, and then we'll pick up? All right, so we'll end the recording here.

 

[End 200614_0154_D]

[Begin 200621_0165_D]

 

Zierler:

Okay, this is David Zierler, oral historian for the American Institute of Physics. It is June 21, 2020. I am delighted to be back with Professor Rainer Weiss of MIT. Rai, thank you so much for joining me again!

Weiss:

Okay! You're quite welcome! We left off the last time at something called the troika, but I'm backing a little bit from that, so you can understand a little more what was the tension that was associated with presenting LIGO, this new project, at the NSF, at the time we did in 1983. What happened was that—and this will recur later in this story again—but the setting in 1983 was that Weber’s experiments had not been verified by anybody else. And one of the most important people not to verify them was a guy named Richard Garwin. Richard Garwin was the Buddha of physics; that’s the way I would call him. He’s still alive, by the way. You probably know him.

Zierler:

I know him. I interviewed him.

Weiss:

You interviewed him? Oh, excellent!

Zierler:

I did, I did.

Weiss:

But probably not about this. Probably about more important things!

Zierler:

He wanted to talk mostly about coronavirus, because that’s what he’s working on right now.

Weiss:

Oh, really! Interesting. Good for him. Good for him!

Zierler:

I mean, he solved everything else! What’s left but coronavirus?

Weiss:

(Laughter) Well, he tried to solve the gravity business, too (laughter). He probably didn't tell you that. It came twice. Once is ahead of us, and once is behind this story a little, early in this story. What happened was that during the Weber time, he was the chief scientist at IBM in the Watson Lab in New York City. It was associated with Columbia University at the time, and it hadn’t yet moved to Yorktown Heights in New York. That was later on, IBM moved out of the city. And it was hard on him. I don’t think he ever moved out into Westchester County himself.

Anyway, what happened is that being chief scientist, he got interested in—he read about the Weber discovery, and he and a guy named Judah Levine, who happened to be a research scientist at IBM at the time, decided they would—“Well, let’s quickly see, can we see something?” And they built a quite different version of the Weber bar. You won’t find that in the garden at Maryland, where I told you about last time. This one I think is so hostile, they won’t have anything to do with it. But what it was—it was maybe one quarter of the size of what Weber had built, but perfectly, beautifully instrumented. In other words, they had thought the thing through more carefully than Joe Weber had, and they had made changes in the design, and they had actually calibrated the system in a better way than Joe had.

And so within about—I don’t remember the exact time, but ’69 was the paper, the discovery paper. I think by ’71, they had their system running, and Garwin and Levine were seeing absolutely nothing. I mean, really nothing. And they would challenge Weber. They would go to Physical Society meetings, especially Garwin, and would give a little report on their experiment, say—they showed the statistics. They did the analysis really very well. And Weber said, “I see things. You don’t see things. The only reason you don’t see things is you didn't do it my way.” And that’s something that made Garwin so pissed off. I mean, this was against his whole instinct about how you deal with physicists. Namely, that’s not an argument. You calibrate it, you measure the noise, you understand the system, and you don’t say “You didn't do it my way.” “I think I did it better than you did!” He may not have said that, but I'm sure that was on his mind. And he let that go for a while.

Weber—nobody else had seen anything either, by that time. And so, Weber—actually there was an open fight between Weber and Garwin at one of the later Physical Society meetings, where finally what happened is that Garwin accuses him of faking it. I don’t think that’s the right word. Of—well, yes, he effectively accused him of doing something pathologically wrong.

Zierler:

Worse than massaging the data.

Weiss:

Oh, yeah. Yeah, yeah, yeah. That he was wishing rather than doing, OKAY? I mean, he did the experiment, but I think—and he—the whole thing got very nasty. And Garwin, who has a certain righteousness, as you probably can well believe, just felt he had dominated, and he had managed to discredit this. And he made it stick. So very quickly after that, a lot of people who were in the business also, who had done it the way Weber had done it, didn't see anything either, and the thing stuck. And Weber got discredited in a big way.

Zierler:

Rai, where are you on this? You're on the sidelines? You're watching this?

Weiss:

I'm on the sidelines, because I'm teaching relativity. It’s between the time I came up—see, I taught this relativity course in ’66, and this was ’71. I had already come up with this gedankenexperiment, but I hadn’t yet converted it into a real experiment. That was about to happen. In fact, Garwin may have had some influence in that. But others did too, namely that nobody had seen anything, and I began to realize maybe there was something really wrong with the Weber idea.

And I spent that summer—and I don’t want to go back on that—but thinking about how you really make a thing with an interferometer. But anyway, the important thing is that the ambiance—that’s the thing I'm trying to tell you—the ambiance had been set, that this was a field that was either full of crazy people and was full of nuts—nuts and crazy people. That’s sort of the impression. And that was certainly the impression that people at MIT had, still, about this, people all over the world, after they failed to show that Weber—they could not confirm Weber. And so it was a very big deal. A very big deal, for the National Science Foundation. This is now many years later. This is now ’83. And the mood has not changed, OKAY?

This is—well, there were two big deals. One was Caltech jumping into the field which was discredited. That was in ’78 or ’79, the meeting that Kip Thorne and I had back in Washington. That was a very big deal. I think the biggest deal of all was really when Caltech put a multimillion-dollar investment into this field. But then, that was a very local thing. But a very public thing was for them to encourage me to look into how you would make a big one. And that was the Blue Book study, and that’s what the Blue Book study was talking about, when we went in front of the NSF committee.

Zierler:

Who’s “they,” Rai? Who encouraged you?

Weiss:

Richard Isaacson. And Richard Isaacson—I suggested that, that we do this, because I felt that we ought to find out if it’s even reasonable to think of building something that would ultimately detect something. That was my idea. But I really got encouraged by the fact that Caltech had put their own investment into it. To me, that was—MIT would not have even encouraged that. They were dead against the whole idea, as I think I told you. But we'll again come to that in a minute. So, what happened is that I got this encouragement from the NSF when I put my proposal in to do the study, and we then did the study with industry. And I think I described that to you already.

Zierler:

Mmhmm.

Weiss:

But what came of it was the thing that we hadn’t yet talked about was the end result of that was two things. We had done the Blue Book study. All during the time of the Blue Book study, Kip Thorne and I—and occasionally Ron Drever, but mostly Kip and I—would talk about, “What’s the Blue Book for?” It’s really to document for the NSF so they can make a decision about whether they want to go—for the next big thing. And the hope was that it would foster a large collaboration, especially of those people who want to stick in the field of gravitational waves.

Now, Kip convinced me eventually that—and that’s something I maybe hadn’t told you—that it wasn’t so smart to do that. It was probably better to try to make a collaboration now between Caltech and MIT. Because the big investment that Caltech had made into this particular field of interferometry, which was the only other people doing that at the time were doing—was MIT and Caltech, and also now a little bit at Glasgow. And of course, the Germans. But in the United States, it was Caltech and MIT. And so, he felt that that was a better deal. And that was a very hard thing to do. We met—Kip, I, and Ron Drever, and my son happened to be with me at the time—we met in Padua, at one of these general relativity conferences. I don’t think I told you about this.

Zierler:

No. Not this.

Weiss:

And he was twelve, eleven, something like that, and he wanted to see Venice. And I took him with me to this meeting, and he sat in the back. We met in a hotel in Padua, I think. And there was Kip, Ron Drever, and my son and I, all sitting in a room together. And he was reading this book or whatever it was, but he couldn't help hearing what was going on. And what it was all about was Kip trying to convince, and me trying to convince, Ron Drever, who had just made the decision to join Caltech and leave Scotland, but that this was going to be not a thing he could do on his own at Caltech. That it had to be a collaboration. It was just too big a deal. No single group could take a thing on that would look to everybody like a $100 million-dollar proposition.

Well, Ron—and we'll have lots to say about Ron as we go on—but Ron didn't see it that way, and he complained bitterly to Kip that he wanted nothing to do with me, and he didn't take that job at Caltech to be made part of a collaboration. And Kip had to backtrack. It was very tough for Kip. He had done a hell of a selling job on getting Ron to come to Caltech. But in the end, the thing that convinced Ron in the end that he had to collaborate was Rich Isaacson. That was not at the Padua meeting. My son, after we left that Padua meeting, looked at me and he said, “What in the hell are you trying to do? That guy wants nothing to do with you!” I said, “Yeah, yeah, I understand that. I heard that just like you did. But it doesn't make sense. It doesn't work in the world of reality.”

Zierler:

Any idea why he was allergic to collaborating with you?

Weiss:

Oh, god, yeah. Now I know everything. But that’s a whole story on its own. Maybe I should divert there. I learned a lot of this.

Zierler:

I think so.

Weiss:

I learned a lot of this only very recently, okay? I learned that from his brother, who I met, who is very a completely different person, Ron was an Asperger syndrome kid and had terrible trouble with people. But he was extremely smart, but in a way that was different than most people. I don’t know; have you ever met Asperger syndrome people?

Zierler:

Yeah, I have.

Weiss:

They have their own way of looking. But Ron had lived enough in the society so that he covered much of that. But he had a streak in him, which I couldn't understand, and I thought it was sheer egotism. I completely misread the man. I mean, every time you mentioned something, he—“I invented that!” “Well, how did you invent it, Ron?” “Well, I thought of it, and this is—” And then he had it not quite right. There were many, many instances where—and it turns out that he was saying something he believed. It wasn’t lying! It was one of these really terrible situations. But most of us did not understand it.

For example, a wonderful example is something that happened to a guy named Roland Schilling, who was one of the members of the German group who had done some beautiful work. And as Ron got into this, he had some ideas, and one of the ideas he had—well, first he went to visit the German group. And I'll give you an example. This is a typical example of Ron. He went to go visit the German group. This is just as he was getting started. And he talked with Schilling, and Schilling explained to him on the blackboard—or on a piece of paper, but it was written down in equations, okay—a particular idea where you could use the light that was in the interferometer over again. I can explain all of that to you, but there’s a way of making it so that when you have no light going to the photodetector in one of these instruments, all the light returns back to the laser, and if you put a mirror between the laser and the beam splitter in the interferometer, you can cause the whole interferometer to be a resonant system. And so, what happens is the light goes in and builds up inside the interferometer, and you can make it look like you have a laser that may be one hundred times more powerful than the laser you actually have. In other words, it builds up the light dramatically.

And Schilling had come up with this idea, a very clever idea, I thought. I didn't know about all of this, but I learned the story only many years later. And what happened is Ron couldn't quite follow the argument. Why? Because he did not follow an equation well. He understood things best in pictures, okay? And a lot of very good scientists do that. I have the same problem, but not as badly as Ron did. I mean, I eventually get myself to think about the equation and draw pictures of the equation. Then, it works fine. But the point is Ron was even more inhibited about that than I was. And so, he sat on the airplane thinking about this. And Roland probably had not drawn any pictures for him. And he started putting the thing together in his head and drew some pictures of how it was possible that you could build up the light in the interferometer by putting another mirror blocking the laser light. That was sort of a completely nutty idea, but you do—you can get—so he finally caught the idea, somewhere on the plane back from Munich to Glasgow. A picture that he drew gave him that idea. And he could see it as a picture. And so, he thought he invented it, because he wasn’t told it that way. OKAY? And he goes to the people in the Glasgow group and explains this new idea that he has had. And they say, “Well, you should publish that.” And he goes ahead and publishes it. And Schilling has an absolute shit fit. I mean, he is ready to never let Ron in that lab again, okay?

Zierler:

But Ron wasn’t trying to be sneaky. He thought he was doing something right.

Weiss:

He thought he had invented it, because he did it. He had this way of pictorially looking at this process, which he didn't get from Schilling. You see? And he didn't choose to try to understand the equations. So, this was a thing that happened over and over and over again. And I mean, many people after a while wouldn't deal with him because of this kind of thing. He got a terrible reputation with a subset of the individuals. A lot of other people thought he was absolutely so brilliant that you could not imagine working without him in this field. It was a very complicated mix.

And so, one of his problems, as I learned later from his brother Ian—and I only learned this when he—I mean, I lived through a lot of bad experiences with Ron, which we'll get to in a minute. That’s why the troika didn't work, okay, and why nothing worked for a long, long time. And that’s why Robbie Vogt, who is a bit of a complicated man himself, later on threw him out of the project. This caused an unbelievable sensation and almost killed the project again, for the third or fourth time. That’s something that goes on into the future. But the thing is, he threw him out of the project. He didn't want to deal with him anymore. And the Caltech faculty just about blew up. A lot of people.

But eventually—look, let’s go back to the troika a minute. So, Ron is a very complicated figure. We'll be coming back to him periodically. But the thing I learned from his brother, who I didn't know, and I wish I had known him, and also his niece, is that Ron was always difficult as a child, even, because of this Asperger thing. And his brother was two or three years younger than he was, but his mother—this is now Ron’s mother, who—the family tradition in the Drever family was always to have a doctor in each generation. It goes way back into the 1700s, near Glasgow. It’s either in Edinburgh or Glasgow. But the Drever family was famous for having all these doctors. And that’s what Ian was, was a doctor, and his daughter is a doctor. Those two. So, they had a different view of this whole person than most of us had. And so, Ian told me that when he was three and Ron was six or something like that, his mother, who was a doctor, didn't have time to take care of Ron all the time. Because he had these strange things. He would disappear and run off, then he wouldn't show up for a day or two. Because he was looking at something that got him all interested, and then he forgot what he was doing, and he would sit there just looking at it. And so, what happened is that he was given the job of minding his older brother. And this is when it turns out that Ian would take Ron, bring him back to the house, and that they would then—everything would come back. And it turned out, most of the rest of his life, the connections that Ron made with people were first vetted by his brother, Ian. In other words, Ian became his mentor forever and ever.

And so, Ron, one of the reasons he could—one of the problems with Ron is he could never make a decision. And I don’t know; that’s part of Asperger, I believe. I hear that from people. And so, he would mull this over and over and over again. How should this—anything that limited his—took away a contingency, that eliminated a possibility, he would resist it, because it would reduce his ability to invent. So, his feeling was every time a rule was put up, every time a decision was made, it closed off an opportunity. So, it was very, very hard for him to be involved in making a decision. With anything! And this is what we'll get to in a second. So, I always thought this was a flaw in the man as egotism, as a way of saying, “If I'm going to do this, I want to do it myself, and get all the glory.” Which is bullshit. That’s not the way he was operating. It was out of a fear of losing an opportunity, fear of being dominated in some way.

Zierler:

And Rai, part of his condition was probably that he was unaware that this is how he was coming off to other people.

Weiss:

That’s probably true. Probably true. And a lot of people who met him once would say, “Oh, he’s just a charming and wonderful Scotsman with a wonderful little accent.” You know, a brogue. But then if they tried to work with him, they would have trouble. I know virtually a whole list of—there’s very few people in the world who, in the end, could work with Ron. They eventually just tried to run away from him as fast as they could. But look, I don’t want to denigrate him. He came up with some very important ideas, and I will point them out to you as we go along.

So okay, now I can take you back to the—that’s a little diversion on Ron, but there will be more. But what comes of it is that here we have Garwin, who has effectively doomed the field. NSF is going forward anyway, and here comes this very public thing where we make a presentation of the Blue Book. And this is now done jointly by Caltech and MIT. And the reason that happened is because Isaacson told Ron, “No, you can’t do it alone. I can’t get the money for the thing if you're going to be alone.”

Zierler:

Alone means collaborating with MIT?

Weiss:

“You have to collaborate. This is too big a deal for one person.” And one thing that you could always get Ron with is if it had to deal with money (laughter). If somehow it jeopardized money, that was something he listened to. It was sort of a very interesting situation (laughter). There was no honor associated with money, and there was no inventiveness associated with it. That was a given condition. So “Goddammit, all right, I'll go along with it!” So, the thing is—yeah, okay, so we give this presentation together—Kip, Ron, and myself—in 1983. And it comes up—the people on the committee—remember, the time was—this was the field that was full of crazies. No sources were really known about. We hadn’t yet developed the technology completely to the sensitivity that we needed. But nevertheless, we were pushing to build a big system. And that was my fault. And Kip’s, a little bit, but mostly mine. Not so much Ron’s.

So, then I expected that we would get very nice words, maybe, but not what I got as a recommendation from that committee that looked at this. And that’s what you'll find in that writing I sent you. And in there—and I'll just paraphrase it, but it’s written beautifully much better than I can paraphrase it—it says, “Look, this is a new field for the NSF. NSFP is the only people doing this. It looks kind of risky. We don’t know enough about the technology. Will it really work? We don’t know about the sources. Are they really there? But it’s of so much fundamental importance that it’s exactly the right thing for the National Science Foundation to make an investment into.” That’s in fact the paraphrase of the whole thing. You couldn't ask for more better. And the thing was written by a very famous guy who was a theorist in gravity. I later learned that, and I put that in that paper. It was written by a guy named—oh, boy. Hold on. I'll tell you in a minute; I've drawn a block. Oh, yeah—by a guy named Stanley Deser, Professor Stanley Deser, who was a theorist at Brandeis University, and a very strong advocate for anything new in gravitational theory, gravitational experiment. He thought that that would be a wonderful thing for the NSF to do. And I think he was a plant, to be honest with you. I think he was planted by Isaacson and by his boss, into this committee, so he would write that. I can’t prove it, but I've asked Stan that, and he wouldn't deny it.

Anyway, so that brings us to the troika again. Because what happened—this is sort of where it ended last time—was that we immediately after that recommendation were told by Isaacson, “Now you've got to pull a project together. You've got to get all the things that a project has—a project manager, organization. You have to do the research together. You have to make an organization that is blended for this project.” Blended meaning you now are working on this together into the future.

Well, Kip saw this as a wonderful thing, too. Ron was very dubious about this. OKAY? And what happened is that what I did, because I wanted to keep the thing going, I went to my administration—now this is too big a deal to do on your own, but now you're talking about a thing which might cost hundreds of millions, or let’s say—later I made an estimate; in fact it’s in that Blue Book—it was around $100 million total, $60 million for the apparatus only. But at that time, $100 million—the idea was to build, even in the Blue Book, was to build two detectors, one several thousand kilometers away from the other, and build them and look for coincidences. I mean, pretty much the LIGO idea.

And the idea was not to build one—you could not build one. You had to build two of them, because that's the only way you could do the science. And to me, that was absolutely critical, because my problem was the following. As I told you the other time, I could not put graduate students on this in my department, because the department didn't see any science in it, in the prototype. They saw a lot of technology development, but there was no scientific result that would come out of it. And they were very rigid about that. So, what I did is I would take the students and some of them who were interested in this thing and start them on the prototype. But as soon as it got to the point where they said, “Look, I gotta get out of here and get a degree,” I would put them on the cosmic background work, which was going on simultaneously in the lab. And there, a ballooning experiment or COBE, whatever, which, okay? But it was mostly ballooning experiments, because COBE took almost twenty-five years to get done, also, as you learned from John. But anyway, so it was mostly balloon experiments. So, they could write a paper with some new, very interesting astronomical discoveries, but they weren’t as fundamental as gravitational waves. And many of them had very little to do with the cosmic background.

That’s a story I should also be telling you, but that’s too much stuff, is what we were discovering with the balloon program was that we were going to get—as I'll say it right—we got fucked by the dust. Okay? And I mean that. Fucked by the dust. We discovered that, to our horror. Because what happened is that we started trying to look at the cosmic background carefully. We measured the spectrum. But then we wanted to do the next step. Could we see what was called the dipole, which was a proper motion with respect to the sources of that cosmic background? Which is a doppler effect. And we couldn't do it with the instrument we made, because we had unfortunately discovered it’s all the dust emission in our own galaxy. And it took us years before we mapped enough of that dust to be able to get a cosmic background measurement out of that particular wavelength in the spectrum. Man! It turns out what unfortunately we had chosen is the wavelength band that’s right at the peak of the black body, and that turns out to be not only the strongest of the cosmic background, but also quite strong for the dust. And the dust gets stronger and stronger as you go to shorter wavelengths, still. But it was a mess. And so, in fact it was very important for COBE that we had done that, so that they didn't decide to add an instrument in COBE that would try to measure exactly at the peak for the anisotropy.

So, the first piece was this problem I had of not being able to put graduate students on the prototype, so I felt I had to do something to make it so that it least looked to my faculty as this had a prospect for doing science. And that was the other thing that was so important to me about getting the study done. Because here was now finally a document that says, “If you do these things, you will be able to get to a sensitivity which is good enough so that you can detect something.” Or at least we thought that. We thought so. And that was fundamental. Although you aren’t going to do it right away, but if you could show that you had a proposal that went in that direction, I thought I could flip the coin at MIT.

It didn't work. What happened is that—and here’s where that came in. It happened when right after leaving Washington, Kip and Ron went back to Caltech. I went back to MIT. And both of us—I think Kip and I quite separately, tried to get the next step, which was to find a project manager. In other words, for LIGO. And I went to the dean of science and the provost of MIT. The highest level, the president, doesn't have much to do with local research so much, but the provost has a lot to say about it. And if it’s in the School of Science, it’s the dean of science. And the guy who was—we talked a little bit about this last time. But the guy who was dean of science at the time was John Deutch, and the guy who was provost at the time was Francis Low, who was a physicist, very famous physicist. And both of them, although they were not absolutely blunt with me, said, “Look, we're not going to pay for even starting this project. We're not going to even hire anybody. Don’t try to ask us for the money. We're not going to give it to you.” See, the NSF didn't yet have the money. MIT would have had to make a small investment of a salary of project manager for maybe a year. See, the reason why that was the case—even though NSF, we had gotten through the committee, the next step would be to go to the National Science Board, which is above—a high-level board in the National Science Foundation, and they would have had to approve this LIGO thing, which had not yet happened for any kind of real money to come out of the NSF.

So as preparation for that, they told us to go to our institutions—this is now the people at the NSF—and see if you can’t get a project manager, even though we might have to pay for him for some length of—or a girl, a woman; doesn't matter—man or woman—pay for them for a little bit of time. That was too big a deal for MIT, and they were very blunt with me. Later on, I found out why, and that was they just didn't believe in me. I was sort of a cuckoo character from Building twenty, which was often some nutty place. And the other thing was that they didn't believe in the science. There was nobody at MIT—

Zierler:

All you were asking for was the salary of a project manager for a year. So this is, what, an investment of $50,000, $60,000?

Weiss:

You're right on. That’s what it was. Exactly that.

Zierler:

And they wouldn't give you that?

Weiss:

They wouldn't dream of it! And it’s not because they didn't have the money. It has to do with what I was just going to say to you—they didn't think I was the right person to do a thing like this. And second, they didn't think that the science was there. In other words, there was nobody at MIT—except a guy named Phil Morrison, who you may know, who wrote a lot of stuff. And Phil was dead against the idea that there should be black holes, by the way. He was also dead against the cosmic background, but that’s a whole other story. Then he flipped. He was a dangerous intellectual to have as a very strong opinion maker at MIT. Okay? (Laughter) And he I think set the tone that they wanted nothing to do with this.

Zierler:

What about Caltech funding the project manager?

Weiss:

That was easy! You see, now what happened is—so I went to my people, and they said no, and Kip went to his people, which was much easier for them. They already had a whole installation called Jet Propulsion Laboratory. And they just took this person and I don’t know how they arranged for the money. That was not important. They had already put several million or maybe $10 million into this already. MIT had not put a penny into it. It was military money that got it going at MIT. But Caltech had used their own money to get Ron Drever going. So, a little more didn't bother them, and so they immediately had somebody.

And that was a threshold for me. Because what happened is at that juncture, I effectively had to leave—I didn't leave MIT, but I had to now effectively do all my work through Caltech. In other words, what happened—although there were still independent groups, which was not the right thing to do—but the contract, the deal with the NSF would now be coming through Caltech, because Caltech would be the prime contractor with the NSF. That’s when that happened. In other words, they took the project over; Caltech did. And thank god they did, because MIT wouldn't go near it.

And so, now comes new stuff, which you haven't heard. So, we had to try to figure out how to manage such a project. Here’s Drever who doesn't want to collaborate; me, who wants to make it go forward, but I was leery of Drever in any position of making decisions, because he couldn't make decisions. I knew that, by that time. And the other person was Kip. Kip, who was a wonderful theorist, had never really been involved with a large project. I had been involved with—the ballooning project was moderately big. But I had become deeply involved with COBE, with John, and tried to sell that project and pull it together administratively. John had done more than I did, but I did some, also. So, I had some experience with that. But that was something that Drever wouldn't tolerate. He couldn't imagine that I should be director of the project. And I didn't push it, because I knew it would fail.

And so, what happened is we did this very silly thing. There was no simple solution. We made this troika, which consisted of Drever, Thorne, and myself, as running LIGO, with a Caltech project manager. That was the setup that was operating by 1984.

Zierler:

Rai, did you ever think about just decamping for Caltech altogether?

Weiss:

No, we talked about that a little bit. My record is so bad, of every sort, that—I mean, I told you that. Kip tried to hire me, although not with any seriousness. But I sent him my CV, and he said, “Where’s the other half of it?” You know, that kind of thing. So, I knew that. Look, I told Kip, “You'd never be able to convince anybody to hire me.” So, the thing is, many, many years later, I had a really very good offer from Princeton. When the thing was going—when was it? It was after the second big ordeal. When we get there, I'll tell you about that. But my kids, by that time, wanted not to move. I mean, you will have the same problem. Your kids are eleven, twelve, something like that?

Zierler:

A little younger, but yeah.

Weiss:

Okay. Wait until they become teenagers. They become the most conservative thing you ever met! (laughter)

Any rate, so the troika—what happened was that [sigh] after the—two very big things happened. We tried to run this together. The first thing that the project manager gets asked to do by Caltech, in this case by Drever, is to do the Blue Book over again, on his own terms. In other words, the Blue Book had all this stuff on costing and how you build, and Drever felt that it was too expensive. Well, later on, we found it’s too cheap. That’s the real problem we ran into. So, he felt that he could do it more cheaply if he had another study done. And for reasons I can’t tell you, I was not involved in that. There was a JPL mini study done, and they didn't come out with any cheaper way of doing it.

And so little by little, this guy named Frank Schutz who was the project manager began to realize that he was dealing with a naïve. That was with Drever. And he began to realize he couldn't really deal with him. Every day, there would be—the trouble with Ron was, every day there would be a new idea to work on. And there was no consistency, no constancy. And there was certainly no way to make a decision. You couldn't make—“No, no, no. We're not ready for a decision.” And so consequently, this was very bad for a project manager who wasn’t brought up—I mean, a project manager is supposed to make those decisions. Otherwise, he or she fails. Otherwise, no progress is made.

And so, what happened is that I had spent by that time some—this is now—some time at MIT—for complicated reasons, I found a site where we could do this that was in Maine. In Cherryfield, Maine. I don’t know how well you know Maine, but it’s way up north of Bangor. Okay? It’s in Washington County, Maine. And there is a place there which you'll never believe. It looks—the state geologist of mine pointed it out to me—it’s almost as flat as some of the places out west. I mean, you had 180 degrees—you could see the 180-degree—you could see, no, a 270-degree horizon. Which is sort of—it was flatter than you can imagine. Everything is so hilly, everywhere. No, this was flat. Certainly, for four kilometers, five kilometers even.

So, I got very interested in that site, and in fact I talked to University of Maine in Orono to see if they were interested in becoming part of this collaboration. Because if we went there, we’d need a local base. They were interested. So, I was doing that, independent of the project manager, because he never paid any attention to me. Everything was still—and eventually it turned out that the project manager—the only thing that was left for him to work on that Drever didn't want to control completely was the choice of the sites. And so, he came east for a while. He looked at this site in Maine. He thought it was a good site. And he started working with me on what kind of things you need so you could learn how much it might cost to build there, and stuff like that. We did a survey. We did some geophysical studies to see what the ground was like underneath. Stuff like that. Everything looked good. And then I worked with him trying to find a site in California, and that was very much more difficult. Why? Because there’s too many choices. Too many flat places all around in Nevada, California, New Mex- well, not as far east as New Mexico, but certainly Nevada and maybe Arizona.

Zierler:

Rai, what’s so important about having a flat site?

Weiss:

Well—oh! Because it’s important to save money. In fact, we're in that again now. Over four kilometers—we wanted to build something—I was pushing for something with ten-kilometers arms. I would accept five-kilometer arms. And we could find many sites that could take five-kilometer arms, but you had to do a lot of—there were certain conditions on the site. And that’s a whole story on its own. The best thing is hopefully it would have been government-held land—you know, BLM land, or government land, and we wouldn't have to deal with private owners. In the end, that only worked for one site. But leave it go.

It had to be environmentally flat. Why did it have to be flat? Because the cost of cut and fill are very different than the costs of boring, OKAY? If you had big tunnels, that’s vastly more expensive. You wanted to make a surface site of enough surface that you didn't have to do much digging. In other words, the cost of cut and fill is fundamentally the thing that drives the costs. And so, you wanted a flat site, if you had any choice whatsoever. In fact, what you really wanted is a site that looks slightly like a bowl. And we are now in that, deeply now, because we're thinking of a forty-kilometer system.

And there, it turns out—let me say—and this is sort of cute—that if you build on four kilometers, and let’s say you can’t make it so that it looks straight—because if it looks straight, it’s following the Earth’s curvature. And that difference in height between the ends and the middle is thirty centimeters due to the Earth’s curvature over four kilometers. So, it’s about two feet. Yeah, about two feet. In other words, the Earth is doing this, but you want to make a chord that runs across it. Because the light doesn't travel around that arc; it travels along a straight line. So, to avoid that cost was looking for a flat site. And that would have been an extra—somewhere between $50 to $100 million more, depending on how bad the site was. And we have plenty of sites that are flat in the United States. We have a few in the east, but most of them are in the west. And the initial idea was with this project manager that—which was not a necessarily bad idea—was couldn't we find some sites near the two universities—near MIT and near Caltech, in other words—that would make it easier for the people who were working there. And that’s why we started looking around our local areas, where there are flat sites around. And it was easy to find flat sites all around California; it was very hard to find flat sites in the East Coast.

So anyway, so what happened with this manager was he and I became quite friendly. Why? Because he could deal with me. Ron Drever wouldn't deal with him, and Ron Drever wouldn't trust him. Because what the project manager should have been doing is the following: the first step he should have taken was to take the two research groups and blend them into a single effort. Because there were still very disparate things going on, and there should have been a central plan for how we build the detector, how we do the sites and everything. It shouldn't have been piecewise, two separate plans. There should have been one plan. And he couldn't bring Ron Drever around to that notion. And so consequently, he did what he could do.

And so, there was slow progress in the two groups. Ron Drever now took what he had in Glasgow and built a larger version of it at Caltech. That became called the forty-kilometer prototype. And in my case, I wanted to build a prototype that would let us build things that we could transfer to the site. In other words, instead of a thing that was one and a half meter on a table, I wanted to have things with big tanks, large tanks, so that you could build vibration isolation systems and optics that was appropriate to a four or five—well, between four- and ten-kilometer system. That was all I was thinking about, then. I was no longer thinking about trying to make a one-and-a-half-meter system better and better and better. I wanted to build stuff I could put into the big site. And so, I cooked up a plan, and the NSF supported it, to build a five-meter system in a garage we had at MIT, where I had big tanks, and I could build things then on the scale for the final LIGO. That’s sort of the situation we were in.

And up to the next big event—in other words, there was progress. I don’t want to go into all the little technical things that were progress, because that could take us for the rest of our lives. But no, there were improvements in all these little sensors in seismic isolation and interferometry, all sorts of improvements, made by a lot of different people.

Zierler:

But the big questions remained the same. You're talking about just tweaking.

Weiss:

It’s mostly tweaking. What you needed to get to was this—everybody knew you had to get to a sensitivity that if you put it into a thing that was four kilometers big or five kilometers or ten kilometers big, it could get to a strain sensitivity of about ten to the minus twenty-one. That was the mantra. In other words—and that was the gauge. If you could make a small system where if you could get the position sensitivity, which is different than strain sensitivity—your position sensitivity tells you how well you can measure a small displacement. The strain sensitivity is taking that displacement measurement and by dividing it by a big length. That’s a strain measurement. Okay? So everybody knew that if you wanted to measure a strain of ten to the minus twenty-one on four kilometers, you had to get to a position sensitivity of ten to the minus eight meters. That was the laboratory effort.

And when we started all of this, we were not—the Germans were not hopelessly far away from that, in their thirty-meter system. And we were pretty hopelessly—we were about a factor of one hundred away from that. And the Scots were probably a factor of 300 away from that. And then the Caltech system eventually got to being—and we had to—well, that comes. The Caltech system became the centerpiece for the project when we had solved the problem. We'll get to that now. Okay? But progress was being made on all elements of things by all the groups. That didn't get stopped by all this nonsense I'm telling you, but it could have gone a lot faster if it had been organized. Now, okay. So, the next big event that happens is—and I sent you that—that’s still in the document I sent you. And that was good old Dick Garwin, okay?

Zierler:

Had he been paying attention this whole time?

Weiss:

Well, that’s the point, and I'll tell you that right off. Dick was very close to Marcel Bardon, who was at that time the boss of Rich Isaacson. Marcel Bardon was the director of the physics division of the NSF. He himself had been the director of the Pupin laboratory of Columbia. That was a nuclear laboratory in Westchester. Bardon had been the head of that. He took the job at the NSF as director of physics. I don’t know why, but he did. And Rich Isaacson had worked him over to try to convince him—for years, they worked together—to try to convince him and a lot of other—but mostly Marcel—on convincing that the right thing for the NSF to do is to get big-time into gravity waves. So that's where Rich keeps reappearing over and over again in this story. That’s why Kip and I made an award for him at the APS. I mean, he’s the singular person that kept driving and driving.

Zierler:

And were you careful this whole time—did the issue with Weber—did that loom large, in terms of—?

Weiss:

Yes, it loomed very large. And it loomed large still at MIT, because they were so skeptical. But it also loomed large in Washington and other places. So, it was very difficult, and I think it was really quite a marvelous thing that the NSF—well, that review that was so important that I told you about a few minutes ago—when the Blue Book was used as a way of getting the LIGO going. So, we're now in that epoch where we're trying to get the project going. And by 1985 or—I don’t know if it’s ’85 or ’86—where we were busy getting sites developed both on the East Coast and the West Coast, I didn't finish that story. Cherryfield, Maine, was the East Coast site. And the West Coast site was a big Air Force base out there. Oh, boy. OKAY, it’s an Air Force base just east of Los Angeles. It’s one of the biggest Air Force bases in the country. I'll get the name in a minute. Sorry, I'm having trouble with names today.

Zierler:

That’s all right.

Weiss:

Well, it’s not so all right. But—yeah. Anyway, so what came of it is that the next step here was that Dick Garwin writes a letter to Marcel Bardon. It’s Edwards Air Force Base. Sorry. OKAY? That’s the name of the place. It’s important that you should know that, because it was a big question for Kip whether we should have Edwards Air Force Base be the site, because he wanted to bring his Russian friends to the United States and work on LIGO, and he was very worried that no Russian would be allowed on Edwards Air Force Base. That’s why that is important that you know that it was an Air Force base.

But anyway, JPL had a part of Edwards Air Force that was their own, because they worked on jet engines and various propulsion systems at that Air Force base. And we would have been—it’s a huge place. We would have been on land that was the Jet Propulsion land, far away from all the bombers and other crap they have that was made to bomb Russia. Anyway, so Dick writes this letter to Bardon and says, in effect—I have the letter. If that’s important to you as a piece of historical document, I'll send it to you. I have a copy of it.

Zierler:

Please, yeah.

Weiss:

But anyway, the nub of the argument is he says to Marcel—and they had been working together on something to stop the supersonic stratospheric transport, and also the beginning of Star Wars. They both had been—I mean, Garwin on his own had killed off the supersonic stratospheric transport. He had almost alone done that. But then when it came to Star Wars, he needed a lot of help. And I don’t know exactly what—they had been working together on these dramatic social-scientific issues, which is what Garwin got to be very much identified with. I mean, Garwin is responsible, I think, for killing Star Wars. He should have told you—you say you interviewed him, but you interviewed him about coronavirus?

Zierler:

Garwin, we have multiple interviews with him already, so he wanted to do one that was focused on his more recent work.

Weiss:

Okay, okay, okay. Good. At any rate, so that’s the first—the opening sentence has something to do with, “By the way, this particular thing of this guy—he agrees with us on this particular viewpoint. And by the way, I've heard that the NSF is putting a major effort into this interferometric gravitational wave detection.” And he says in a very dramatic and not polite way, “I know a great deal about that. And I wonder if it’s really a good idea. And if you persist in this, you ought to have a study with people who are really good. Really good. To see if this is the right thing for the NSF to do.”

Zierler:

What’s the implication there?

Weiss:

The implication is that, very clearly he says, “If you want to go on with this craziness, you're going to get into trouble. And if you want to get out of trouble, you better get the opinion of some people who say this is crap.” Now, it doesn't say that. Because in the end, Garwin will take credit for the fact that he saved the project. I'll get to that. He is playing a double role here. But he doesn't say it’s crap, but he effectively says, “Look, without much more reasoning and thinking about this, this may not be a smart thing to do.” All right. And so very clear—so Bardon doesn't waste any time. He gets the letter, and he looks at the structure of LIGO. And I don’t take any pride in this, but it was very clear that if there was one of the three of us that should have been doing what is about to be asked, it was me, and not Kip or Drever.

Zierler:

Why do you say that?

Weiss:

I'll say that because I was the only one who had real experience with dealing with large scientific projects and had a reputation as an experimenter with a lot of people, that things can get done. In other words, partly from COBE, partly from things I had done otherwise. You know, the ballooning. I had sat on endless government committees. You know, once you take money from NSF or from NASA like I had, you get asked for an enormous amount of advice. And you try to be very careful with that, not to try to prejudice things. So, I had a record of being—it’s in my CV—of having spent an awful lot of time on various committees, advisory committee panels, for both NSF and NASA. So, I was the right person to pull the thing together. Drever couldn't have done it. He didn't know any of the Americans. And he was too—you know, just totally incompetent, I think, for doing this. Kip could have done it, but he would have had to have—as a theorist, he didn't have quite the credentials.

So, when I think through their process, it wasn’t anything—it was by desperation they came to me, okay? (Laughter) But anyway, so I was told by Marcel in a letter that I had to get a summer study. I got the thing in April. I was supposed to get a summer study together of LIGO. And well, I know about summer studies from Jerrold Zacharias, and you don’t do that in a month. Because you have to get the people, you have to get them organized, you have to make sure they're the right people, you've got to get an awful lot of stuff written. I said, “I'll do it, but it’ll have to be a winter study.” And Bardon said, “Well, I hope it’s a fall study.” Or something like that. So, it turned out that I immediately dropped everything I was doing, and went out, on the recommendation of Victor Weisskopf, who was probably the most eligible member of our faculty who wasn’t totally against this idea of gravity waves—

Zierler:

What kind of relationship did you have with Viki?

Weiss:

Oh, I had a—well, Viki is complicated, too, as you know. I had a good one, as long as I didn't say something that criticized him. Viki was intolerant to that, even though you wouldn't have believed it.

Zierler:

But he was not one of the MIT people who wanted to disregard the project because they thought you were crazy?

Weiss:

No, absolutely not. And he, through Jerrold—see, Jerrold and Viki—Jerrold brought Viki to MIT. You have to know MIT was a punk physics department in the late forties. Who was their most famous guy? It was Van de Graaff, a guy who made these electrostatic accelerators. There was a couple of people in the department that had gotten special money from Eastman Kodak, and there’s a whole building at MIT called the Eastman Building, which is where the physics department was. Why? Because they had a lot of sort of second-rate people in physics. There was no first-rate person except for Herman Feshbach. I don’t know how he managed, but he was there. And Morse. Philip Morse and Feshbach were the kings of physics at MIT. None of the modern people in physics, none of the high-energy people.

So, what did it is there was a huge campaign. Jerrold landed a job there because he had run the Rad Lab during the war, knew MIT. And he then, because he knew everybody all over the country, made a huge effort to bring in the very best people before they got settled in other universities. So, the department got reborn in about 1946, ’47, ’48. And Viki was one of them, but a lot of people came. Bruno Rossi came at that time, Viki, Bernie Feld. I don’t know them all, but the department changed completely its character. And so, the fact that I was a product of Jerrold’s lab I think gave me a little bit of—yes, I wasn’t totally crazy, okay? Let’s put it that way. And Viki knew that (laughter). So, he suggested a guy to me who I immediately went on a plane—he suggested Andy Sessler. I don’t know if you know who he is.

Zierler:

No.

Weiss:

He was a high-energy physicist at Berkeley who had taken interest in a lot of other physics, for example plasma physics, and some of the deeper problems in nuclear physics. He was a guy who worried mostly, when he did research, about making accelerators. Making the accelerators. And so, he was a guy, in Viki’s mind, who was interested in technology, but was willing to branch out and listen to the story of gravitational waves. It was a very good—it was probably one of the best things Viki ever did for me, to suggest Andy Sessler.

So, I went out to Berkeley and made an appointment, and spent the morning with him, trying to convince him to be chair of this committee. And I said I would do all the work. Everything that was necessary. He only had to run the meetings and write the final report. But we would help with all the writing of that report. We wouldn't make it up, but anything that was needed, he would have only to say what he needed, and we would do that. And then he thought about it for a while. He said, “Okay, I'll do that, but you've got to convince a guy named Boyce McDaniel”—who was a high-energy physicist at Cornell—“and if he joins you, we will make a joint chairmanship of that committee.”

So, I went out to talk to Boyce at Cornell, almost in the same week. He was more hesitant, but I think by the time I had gotten to talk to Boyce, Andy had already called him, and said, “I'm sending you a guy. Let’s pay attention to him.” He had done already a little bit of what he would do from then on, later on. Okay, so I got Boyce McDaniel under the same kind of rules, namely “We will do the work. I will do all the soliciting of the committee, of people. I will ask you if you want them. I'll send you a list—‘Here are the people we're going to go after.’ You cut off those people you don’t want, and so forth. But I will do the soliciting and finding out if they want to.”

And sure as hell, we put together a lot of people. The names of all those people are in the thing I sent you. And that committee was called the Panel on Interferometric Gravitational Detectors or something like that. And I asked Garwin if he wanted to be on the committee, and he said OKAY, after I told him that Boyce McDaniel and others—I asked him last. Mostly by then I had the committee together. I had a couple of Nobel Prize winners. I had people who were in high energy big projects. I had people who were in gravity. There were about, I don’t know, seven or eight people, all very well-known people. Nobody in the gravity wave business, okay?

And then organized a week’s worth of presentations to the committee at the American Academy for Arts and Sciences in Cambridge here. And these were people in lasers, people in all the groups that were working on interferometry. They had to come and give testimony. People on lasers, people on vibration isolation systems, people on making servo systems. Every one of the technologies that was important for the detector. And also, theorists like Kip, and Saul Teukolsky. I think Saul was on the committee, as a matter of fact. Again, I have forgotten who was the—mostly Kip did most of the talking on sources.

Anyway, and so on the very first day—this was something which I was ready to kill—was the first day of the thing I had. We had done all this planning. I've forgotten if it was November or October of ’86. The very first day, we were all meeting at the Academy on the Monday morning, and we're all standing in the foyer, and Garwin walks over to me and says, “I have only today. You have to rearrange things so that I get the information.” And I happened to be at that moment talking to Sessler about arrangements for coffee and other crap like that. And Sessler looks at Garwin, and Boyce sees there’s something funny going on and he comes over. And they see that I'm ready to kill. I can’t possibly do what he’s asking! It wrecks the whole goddamn thing! And people are coming from all over the world, I mean, for this. And so—

Zierler:

What exactly was in jeopardy, Rai? What were you concerned about?

Weiss:

Well, I mean, if he demands that the whole thing be done in one day, many of the people who I wanted him to hear were not there that day. We were going to do this in an organized fashion.

Zierler:

He wasn’t asking for an executive summary. He was asking for literally everything to be crammed into one day.

Weiss:

The important things, all into one day, because he didn't have the time. I mean, he’s a famous man, but that was pushing me too hard. And so thank god for Boyce McDaniel and also Andy Sessler. They go over to him and they pat him on the back, and they said to him, “Look, Dick, don’t worry. We will do a good job, and you will have the report in your hands, and everybody—if you don’t like the report, tell us.” That shut him up, and he stayed for the day, and that was it. But without that, I don’t know what I would have done (laughter). Okay?

And so, all right, so the thing goes really quite remarkably smoothly. I didn't know that, being very nervous about the whole business and being sort of the—what do you call it?—the person who has to run the show, the producer. And the thing that I think—and it turns out Kip and I did not communicate on this thing, but we did this separately from each other—on one evening of that one week, I asked the committee to come eat—I don’t know if you know Cambridge well. Do you know Cambridge at all?

Zierler:

Yeah.

Weiss:

Okay. Well, you know Legal Seafood down by MIT?

Zierler:

Yeah, sure.

Weiss:

Okay. I asked them all to join me at Legal Seafood, the whole committee, and I wanted to talk to them. Some of them knew about Legal Seafood. Legal Seafood was very important, to make sure that they liked the idea. So we went down there, and I told them, “Look, I don’t know what you guys are thinking, but if you want to do something good for this project, recommend to get rid of the troika, and that there be a single director for this project.”

Zierler:

And was there really not a single director, de facto? Was it truly a troika, three co-equals?

Weiss:

No, it didn't work! Yes, there was a director. There was no director, but I was apparently the only thing that—Drever would not agree with anything. Kip would worry about what Drever felt. He would try to convince Drever to agree with this, but then Drever would—Drever was very stubborn. And so, what happened is that the only place where things got done was on these sites. That’s something Drever didn't that much care about, as long as—and we went all over, looking at many, many western sites. I mean, I spent I don’t know how many weeks. Every week I would go out with him and Frank Schutz and we’d look another western site over. He could never make up his mind about which one he wanted. Oh, it was maddening.

I can tell you endlessly about western sites and how you go out west and you see a nice, flat place; all you have to do is look up on the hills a little bit, and you'll see a water line, a flood plain line, virtually on every one of the flat places in Nevada. It does rain there, but once in fifty years. But when it rains, it’s a doozy, okay? So, it turns out that you've got to be real careful if you want to building something (laughter) on a western site in the southwest. Okay, enough.

So, there was no director, no. And we were making slow progress, not fast enough progress. And Kip felt the same way. And poor Kip was in a terrible situation, because here he was, between Drever and myself. I mean, we were the ones who were not being able to agree. Then he tried some stuff which was completely nutty. He tried to give us titles. Drever would be chief scientist for one aspect of the detector; I would be chief scientist for another aspect. I said, “Kip, forget it! That’s not the problem!”

Zierler:

That’s window dressing.

Weiss:

The poor guy. I mean, he lost a lot of weight during that time, and I really worried about him. Anyway. Okay, so what does the committee do? And that you have. And again, I'll paraphrase it for you. The committee had—to me, it was stunning, what they said. I can’t remember every bit of it, but it’s in that document you have. They said—this might not be in the right order—they said, “Don’t build another prototype. Build two detectors of full size.” That’s two recommendations. Effectively, “Don’t wait. Go full bore. Make the two big detectors. Because you’re going to have all sorts of new problems in those two big detectors than you will have in a prototype. We're convinced the idea is all right, effectively, so it’s time to spend the money.”

And then they said, “Make sure that there’s enough support money for contingencies.” This is mostly directed to the NSF. There was an instruction sheet of how to make this project succeed. And then they said, “And you should—because now the divided relativity of two sites and two—it’s best to have a single director.” I think it’s one of the recommendations. There are many more, but they're minor compared to this. So, it was the most positive thing I could have imagined. And so, what happened is I told MIT about this, and I said, “Now can we look for a director?” And all they could think of is it would be me. And I said, “No, I can’t be the director.”

Zierler:

But why not? You're kind of de facto the director already.

Weiss:

No, I'm not. It would have not worked. Drever would have gone back to Scotland immediately, okay?

Zierler:

Which suggests that even after all this time, he still never fully accepted the deal from the beginning.

Weiss:

That’s absolutely true, and even in the design that came next, which was a design which was made so he could be almost independent of everybody else, it didn't work, and he got thrown out of the project. We'll get to that in a minute. And I felt very bad about that, because—anyway, so what came of it was that Caltech made a very interesting proposal to both the NSF and to us of how about using Rochus Vogt, Robbie Vogt. He was provost at Caltech. He was just finishing being provost, had done an amazing job on cleaning up a whole set of bad problems at the radio astronomy facility. I've forgotten the name of that facility, but it will come to me in a second. And it was a big thing. A big thing at Caltech was radio astronomy. But they had a lot of chiefs and no Indians. That was one of the problems.

And so, he actually—I didn't realize how severe he had become, but he had knocked some heads, but he converted that facility into one of the best astronomy facilities in the world. And it’s still a very good facility. So, everybody who knew about the guy, or knew a little about him, said, “Yep, this is wonderful.” And I said, “How can you resist a guy like that? He'll fix this problem.”

And, well, it turns out he was a complicated man himself. I mean, there’s no—and in the beginning, he made a very small effort to try to be equal about it. Namely, here he was a Caltech person, but he had to now do a lot of things which were not able to be done by the troika. First of all, his very first thing that he had to do is we were both—Caltech and MIT—we were both due for our next proposal, our research proposals, and he decided that this would not be two different proposals; it would be one proposal from the LIGO project, and with jobs laid out across both institutions.

Well, Drever had a lot of trouble with that, but luckily for Drever, a lot of the things that—he was closer to Vogt than I was, because I would talk to him on the phone, but they intersected all the time, so a lot of the ideas that were considered the right ideas to work on came from Drever. So he lived with that for a while. People at MIT—I knew I had to make compromises. There was no way out. But I lost a significant number of people in my lab.

Zierler:

Why?

Weiss:

Because they thought the ideas that we had were the ones that should have been worked on, that they had been working on. Here are a whole bunch of ideas that had been worked on by Drever and the people at Caltech; they were being worked on. And I didn't give a damn that much about it.

Zierler:

But you probably didn't want to see people leave your lab, either.

Weiss:

No, I didn't. But there was no good way to stop it. Either I was going to quit and not go along with this whole thing—and I couldn't see—I thought this was going to be finally the necessary thing. I argued for what I wanted, and unfortunately, I think Vogt, who is not dumb, saw that I was arguing from the position of weakness. Because he was there, at that meeting, when John Deutch gave the great big zero. He was sitting off in the back, but he was sitting there, and he saw the whole thing. So, he saw that there would be no—a lot of problem with MIT. There might have been a problem with some scientists at MIT, but there was not going to be a problem for him with MIT as an institution. And I think that’s true. I'm making some of this up, when talking about it, but—

Zierler:

Well, you're not making it up; you're speculating.

Weiss:

Yeah, I'm speculating. Yeah. So anyway, some of the people in the lab, what they did is some of them actually—I said, “Look, maybe the right thing to do now is we send some of the people from MIT”—since what was taken out was that I wanted to build that five-meter prototype, and the very first thing that I wanted to do is start building things for the ten-kilometer system or the four-kilometer system. And Robbie Vogt said, “No, you're not going to finish that five-meter thing. You're going to build things that are essential to make LIGO work, but the most important thing right now we need to make LIGO work is to make the Caltech detector be sensitive enough.”

That was his priority. And he was right to make that priority. Why? Because we hadn’t really completely finished the five-meter thing, and my instinct was not to build the better detector; it was to build parts for the big system. And maybe I was wrong about that. But the Drever instrument was designed—the forty-meter was designed to try to get the sensitivity. And if we had that, we could tell people who were critics we have at least that displacement sensitivity. I wasn’t pushing for that. But they desperately needed things that we could build at MIT, put into the forty-meter system so that it could make it.

And I understood that argument. I tried to explain it to my people, and they didn't quite like that. And part of the reason—it could have been done better, had Robbie Vogt come to MIT and explained his strategy, instead of using me as a way of explaining it. And that was one of the asymmetries that never got fixed. And I didn't realize that—that Vogt couldn't travel well. When he traveled across the country, he would get sick for a day, and so you couldn't bring him east but once—he came once or twice. It was a big ordeal for him. I had no idea.

Anyway, nevertheless, what came of it was that it worked, and MIT started building stuff for the forty-meter. We started building stuff. We started doing stuff that Caltech hadn’t done. And it became a more communal thing at the lower levels. And Drever more and more withdrew. Now, he withdrew because he didn't want to do what Robbie Vogt asked him to do. In other words, more and more of his options were being closed. He wanted to try—instead of finishing something, which was one of the troubles—why the Caltech system had been so slow—they had a wonderful man named Stan Whitcomb, who was an assistant professor, who was hired very early when Drever came to Caltech. And that poor guy had a terrible life. Absolutely horrible life. He would work on the forty-meter system. Ron would go back to Glasgow for a couple of months. He would do that until—I've forgotten—in the early days of him being at Caltech, he was part-time still at Glasgow. So, they were building this forty-meter system and Stan would call the shots, what’s to be built. Drever would come back from Glasgow, have all sorts of new plans. He would say, “This all is stuff that’s gotta be scrapped. We gotta do it this way.” And poor Stan was stuck in a dreadful situation. And they made very little progress.

And the trouble was—this was just that trouble—Ron would continually have a new idea of a different kind. He wanted to try that out. “Oh, no, no. That—we have a better idea.” And this is one of the difficulties of dealing with the guy. And you could not do something under him. So, what happened is Stan eventually got fed up, and he left. He went to industry. He got then back into the project after—but Robbie was then stuck with the following—Robbie Vogt was stuck with the following problem. How do you deal with a guy who we're now working on a project together, and he can’t keep his piece of the project running, without interrupting it over and over again, with jumping something out of the plan?

And that eventually caused unbelievable friction between Vogt and Drever and led to something which was unnecessary but caused Drever—caused us to get into real trouble. And that was—I don’t know how many years it was into it, probably two years into it, by eighty- we all wrote a proposal together in 1989, which was the proposal that started the LIGO project. Finally, there were some people who were engineers, which had been picked out of JPL and had been picked by Robbie Vogt, who he knew. And do you know what a skunkworks is? Do you know what that is?

Zierler:

No.

Weiss:

At NASA, they have that—a very tight little group of engineers that work together, and they never work on any other project but one project. In other words, they don’t get loaned to other projects. That’s called a skunkworks. See, in many places, what you do is you take all the engineers and put them into a separate organization within the main organization, and the scientists are over here, the engineers are over there, and the scientists want to do a project, they go to the engineering directorate and say, “We need x engineers,” and they get rented from there, and go into a new thing called the project office. And the project office is the place where the scientists and engineers come together. But the engineering is then—these same engineers that came out of the engineering directorate don’t stay with that project. They go back to the engineering directorate, and go on to some other project. So, the science directorate and the engineering directorate are not the same. Okay?

And so, the engineers get effectively moved from one project to the other that the scientists invent. That’s the way Goddard works. That’s the way a lot of the NASA projects work. And the way Robbie wanted to avoid that is he wanted to have a group of engineers—since a lot of his experience was NASA experience, Robbie Vogt—he was once principal scientist at Jet Propulsion Lab, which was all NASA. Remember, Jet Propulsion Lab was a subsidiary lab of Caltech. Are you with me?

Zierler:

I'm with you. I'm with you.

Weiss:

Okay. It’s a lot of stuff, all at once.

Zierler:

It is, it is!

Weiss:

(Laughter) Yeah, we're going to get into trouble again, damn it! Anyway, what came of it was that Robbie, having had NASA experience, wanted to avoid—he didn't want to rent engineers from the engineering directorate at JPL. He wanted a cadre of engineers that were going to be his for LIGO and would see the whole project through, from beginning to end. But he could only afford so many of them. So, he had a very tight little engineering office which was the skunkworks.

And what happened was—and there were names of people that—maybe not important to you, but a thing that he invented was very good. One of the things that Robbie invented before he went off on the deep end with Drever was he invented something that he called a scientific liaison to the engineers and the contractors. And what that means is that here is this scientific project which has all these requirements—on the buildings, on the vacuum, all the things you're going to build in this infrastructure—but they're very strange requirements because of this enormous sensitivity that’s required. So, he felt very uneasy about having the engineers go alone, and he wanted scientists to go work with the engineers, and also when this got put into industry, to work with the contractor. A very good idea, I thought. And by the way, the reason why John Mather succeeded so well is because he broke the rules at NASA. I hope he told you that.

Zierler:

Well, in so many words. How do you see how he broke the rules?

Weiss:

Oh, because! He didn't have a skunkworks. He had these engineers that were rented from the directorate, and a very tricky project. They had never done cryogenics at Goddard before. They had never done infrared astronomy much. So, he was in a situation where there was very little experience in the engineering directorate, but he needed these people to do COBE. He couldn't do it without them. So, he did the following. He would not let an engineering meeting go on without him there. In other words, the engineers were not separated from the scientists. He wouldn't let that happen. And that is exactly what Robbie did, too. And that’s how it happened that COBE succeeded. The engineers had no experience with this.

Zierler:

You're saying there’s no way it could have succeeded any other way. This is the only way.

Weiss:

I don’t think it would have succeeded any other way.

Zierler:

Wow.

Weiss:

There’s nothing wrong with engineers. There’s lots wrong with scientists and engineers. When you put them together, they work pretty well (laughter). I'll tell you why. There are many, many, many examples of that, in a large project. You see, the scientists—I mean, Drever is an extreme example, but many scientists keep changing their minds. You make a plan, and they think about it—“Oh, no, here’s a better one. I could get another factor of two if we change the plan.” That’s what killed the superconducting supercollider, the big accelerator in Texas. Poor Clinton was stuck with the problem that some guy had decided, “We could make it better by a factor of two!” After having already invested $6 billion in it.

Zierler:

Right, exactly.

Weiss:

And those guys were arrogant enough to try to push this through. It was a mistake. But, okay.

Zierler:

Rai, the story just keeps getting better and better. And more improbable, too. That’s the incredible part. You would think that as you're developing this narrative, things sort of start to come together, where you can see—

Weiss:

Well, they do.

Zierler:

Right. We're not there yet, though (laughter).

Weiss:

Not yet (laughter). They will. One more generation of administrators (laughter). And then it smooths out.

Zierler:

Okay. I figure it’s got to at some point!

Weiss:

Yeah, it does. It does. But the fact that the NSF lived through all of this is sort of unbelievable. Anyway, Okay, so what Robbie Vogt does is he gets himself together a little skunkworks, and it turns out that some scientists pulled out of Drever’s group, which Drever was very jealous about and didn't want that to happen. And some out of my group came to work with the engineers. And I worked—for example, I put six years of my life into working on one of the major projects that had to do with infrastructure, namely the beam tubes, those things that—the vacuum system that runs the four kilometers. Turned out not to be a trivial job at all.

And then there was a guy named Mike Zucker who was part of the group at Caltech, and he worked with an engineer on making the vacuum system that is inside the buildings where the tanks are. The beam tube vacuum system is this very strange thing with four kilometers of just tubing. But then you have to have all these tanks which hold the inner parts of the interferometer and all those parts together. And they had some very special properties that had to be thought about. So, there was about a $60 million contract—this is later, so you can see where we're heading—for doing the beam tubes. And an engineer and I were in charge of that. There was another about $60 million project for doing the vacuum system in the buildings, and that was Mike Zucker and an engineer.

By the way, the engineers’ names—the guy who worked with me was a guy named Larry Jones , an absolutely wonderful engineer. The guy who worked with Mike Zucker was a guy named John Warden, who was a wonderful engineer. They're really first-class people. And then there was another wonderful engineer named Fred Asiri, who worked with another one of the scientists in the group called Rick Savage, who worked on the civil work, the buildings. And each one of those contracts was of the order—when we finally got to it—of about $50 million. Five-oh. This is later on, when we had a management that worked. We'll get to that. But the idea that there should be a coupling between an engineer and a scientist was something that Robbie insisted on, and that was already causing friction for Drever. It didn't cause any friction for me, because I understood it.

Zierler:

What did you understand? That it was necessary?

Weiss:

It was necessary. Oh god, yeah. And Drever did not understand that. He thought he had been robbed of somebody who was essential to him. So, at any rate, there was a lot of little things that made friction between Vogt and Ron. I cannot tell you the exact thing that caused all the trouble between them, but there was a point when they had hired a new assistant professor at Caltech named Fred Raab. And I don’t know the exact date, but ’89, he was already on the ’89 proposal. Som my suspicion is this happened in about ’91 or ’92—what I'm about to tell you—is that he effectively told Drever that he was no longer going to be involved in the project. There’s stories, which I don’t believe, that they actually locked him out of his office. I don’t believe that, but Drever contends that.

And so effectively, this genius, as he was considered by the Caltech people, was removed by this “second-rate” scientist, the project director, who they all knew, Robbie Vogt. He’s not a second-rate scientist at all, but a very good scientist. But he had no—there was no simple way for him to solve that problem that he had with that man. And he did something which if he had a little more patience, maybe a little more wisdom, he could have done without injuring himself. But he didn't do it that way. He absolutely made it so that Drever could run all over Caltech and say, “This man has thrown me out of my project, and they'll fail because they don’t have me.” That’s what he did. And the last thing anybody at Caltech and the faculty want to hear is that they're going to fail. And there was already a lot of problems between provost—him being provost—this is now Robbie Vogt—having been provost, and the faculty. And Caltech, by the way, is sort of an anarchy. I don’t know if you know that.

Zierler:

I've heard that.

Weiss:

It’s a very tiny administration. At MIT, it’s now half—well, not quite that, but a third of the place is administration. At Caltech, it never—and so they have these large fluctuations. I mean, when they have a crisis, it’s enormous, and there’s no damping to the system. So, what happened there is they had a rebellion, and it’s all over every newspaper in the country. Lots of stuff in Science magazine about this famous scientist, this brilliant man, being thrown out of his own project, by a cruel dictator.

Then they started saying—you know, because he was German, people would ascribe to him any—maybe something a little Nazi-like. But look, I knew him pretty well. He was a victim of Nazis. I didn't like—I had a lot of problem with him, too, with Robbie Vogt, but I understood what he was up to. At any rate, so then there was this really horrible thing, which Robbie had to live through. The Caltech faculty began to campaign to get him thrown out, get Vogt thrown out.

Zierler:

Are you thinking at this point that the whole project is in threat, at this point?

Weiss:

Absolutely. It was in dire threat. Absolutely dire threat! I mean, by the way, by this time, things had changed at MIT. Completely. We had a new provost We had a new president who thought that LIGO was the greatest thing that ever he could think of. This is the president. Vest. Chuck Vest.

Zierler:

So, it’s not that your stock had risen; it’s that there were new people calling the shots.

Weiss:

That’s right. MIT, the president was all for it, and new elements of the physics department thought it was a great thing to work on. They thought the science was first-class. And they began to look for faculty for this. They began to allow me to have other faculty. So, everything had changed at MIT by this time, and now Caltech was falling apart. Okay?

Zierler:

You can’t make this stuff up!

Weiss:

(Laughter) You can’t! And I would have these very complicated conversations with Vogt, who was trying to convince me that I should get the MIT administration to back him. And I could not bring myself to do that.

Zierler:

Why not?

Weiss:

Well, first of all, I didn't really know where they stood, really, in the end, when it got to being a crisis. And I didn't trust them to take the project over. I don’t think they could have done it. And if they had suggested it, I wouldn't have agreed to it. There was already so much investment made at Caltech that I thought the better deal was to try to change it at Caltech. In retrospect, that was probably what was driving me most. The last thing I want to do is lose the project.

Zierler:

So, what are you doing to keep the whole thing on life support at this point?

Weiss:

I can tell you exactly. I would have these endless conversations with Vogt, trying to calm him down. He knew that I had endless trouble with Drever as well. But neither of us knew exactly what was the trouble with Drever, until what I learned later, what I told you earlier today—that he had a real personality issue that was in his head, this thing that he had from childhood. None of us knew that.

So, what happened is that I convinced—I did a couple of things behind the scenes. One of them is I convinced Stan Whitcomb, who was the guy who couldn't work with Drever, to come back to LIGO. He had been working in industry in California. To come back to this thing, which he really enjoyed working on. And he did; he took the job. Now the reason he did it is because he got along very well with Robbie Vogt. Robbie Vogt had been one of his mentors when he was a student at Caltech, and they liked each other a lot. And poor Stan got into a terrible problem when this whole thing broke up between Drever and—I mean, he could understand why Robbie Vogt couldn't deal with Drever, because he couldn't either, but he didn't know quite what to do about how to keep the project going at Caltech. Although he had just come back. He was assistant director to the project. He was deputy director. So, he and I had lots of talks about that.

So, we decided—what had happened is that—“Let the crisis blow out.” Caltech, once they came to their senses—and I talked with many people at the level of the—well, who were the people at Caltech? I talked with the president, a lot, who was Eberhart, who wanted to know what should they do, what should they do. And I said, “Well, maybe you should get a committee from JPL together, a Caltech committee, to look at this whole problem. But stick with it, because the fact is that Caltech has done this major investment already.” So that was one side that I was working on. But not only me; other people were doing the same thing.

And little by little, as more committees were formed to calm the thing down—and people from the top committees, executive committees were formed to look at this thing—they came to the conclusion they let go of Robbie Vogt. They didn't throw him out of the project, but they took him away as a director, and they replaced him by Barry Barish. And Barry Barish was a high-energy physics guy who had been a major pusher for the superconducting supercollider. That had gone down the tubes. He was available.

Zierler:

And what kind of contact had you had with Barry up until this point?

Weiss:

None. None. None at all.

Zierler:

Did you have opinions about him? Were you aware of his reputation?

Weiss:

No, I didn't know anything about him. I had no opinions at all. Except that I knew one thing—that I had to have a heart-to-heart talk with him in part because of the contract with CB&I. We had started a contract with the Chicago Bridge & Iron Company to start building the beam tubes. In other words, we had enough money, the project had gotten through the National Science Board. We had enough money. So that contract was the very first contract that was let.

And what had happened is before that, I had gone out to Caltech a lot. We had worked on a prototype vacuum system of the kind that you would need if you wanted to build LIGO. Remember, that very long tube. And why was this such an interesting problem and complicated problem? If you had built the four-kilometer—it’s really sixteen kilometers of tube, since you have two sites, and both have the letter “L” on them, right? So, four/four, four/four. It’s a lot of money. That’s where a big pile of money was going to go. And we had decided, oh, maybe two, three years before the—Larry Jones and I and some engineers, with Vogt, had decided that we were going to try some new ideas out for making vacuum systems cheaper, big tube systems like that cheaper, by not just a factor of—something like a factor of two—than what people thought it should cost. And I had some ideas. There was an engineer there who had some ideas of how to do that. We felt that in order to prove that you could do such a thing, we would have to build a prototype of something with this new technology in it before we could actually sell it to a company to build, or even sell it to the NSF as something that we thought we wanted to build. So, we did build a prototype with this technique. I can tell you lots about that, but that, you really don’t want to hear! I mean, it’s a lot of grungy engineering details there. But what it had to do with is—I'll tell you one thing, because there is one piece of science. What we were violating was a fundamental rule about all accelerator vacuum systems and plasma tube vacuum systems, and that is—industry vacuum systems—there’s a ratio that’s in every handbook given to the engineer, that there should be—for a given surface area of stuff in a vacuum system, you had to have so much pumping capacity. And the cost of the thing skyrocketed with having so much area, in a thing that’s four kilometers long, a meter in diameter.

And so, consequently—and we had sixteen kilometers of that. So, this would be an enormous saving. And there were some tricks that we had come up with that had to do with the way this welded, had to do with the way you treat the material, so you didn't need all those pumps. And we had to prove that this could be done, and we built a prototype that succeeded in doing that. And all during that time, when the NSF sent committee after committee to Caltech—that’s what they do; once you get a lot of money, every six months, you get a review committee. The review committee had people on that thing which thought that we were out of our heads, and they would tell the NSF, “This is now going to fail.” They would write it down. And I can’t see how Rich Isaacson took that to his management. But he did. He had to. But in the end, eventually most of that stopped.

In fact, we built another prototype, twice as big, at the Chicago Bridge & Iron Company. That’s the first thing we built, out of all the materials, exactly the processes we're going to use in the field. Everything identical—the welding, the cleaning, the heating. Everything was done exactly as we were going to do in the field. And it worked. And that stopped it! So, they could then take a—they took a fixed-cost contract to build this. And they made money. Interesting. Okay?

Anyway, at one of those meetings in the early days of dealing with the Chicago Bridge & Iron Company, Robbie Vogt is still in charge of the project. He has thrown Drever out and is actually—I think he was in a state that I can only imagine was in an internal state in his head, which was such that he couldn't—he was not in a proper state to do what I'm about to tell you. He came to one of the early meetings in Chicago, to the Chicago Bridge & Iron Company. Here’s the president of the company, who I had gotten to know, because I had worked there to build the prototype. And the prototype—oh, I'm sorry, I didn't tell you the contract was divided into two pieces. One very much smaller than the other. One was about a $8 million contract to build the prototype. And then there was a $50 million contract to build the thing in the field.

And he came to the meeting after the prototype was finished, and I had spent a lot of time there—maybe most of a summer, in Chicago, in Plainfield, working with those guys on making that prototype—and he came, and there happened to be a representative from the NSF there. Now, when the NSF had never really done a big project before—I don’t know; this is a long story. If you look in the history of the NSF, they did a terrible big project at its inception. This is when it was crooked and Johnson was in power. And that was the Mohole project. I don’t know if you remember that.

Zierler:

I've heard of it.

Weiss:

It was so crooked it was pathetic.

Zierler:

And it probably made them sort of gun shy about doing—

Weiss:

Absolutely. Everybody would refer to that project—we had endless trouble with Congress about this. And that’s something we haven't talked about at all, how Congress dealt with this. That’s another whole topic. But the Mohole project—that’s all Johnson and offshore drilling. That’s really what it was. But anyway, so these—yeah. So, what NSF had done, they relied on Caltech to be the equivalent of a NASA center. In other words, they relied on Caltech to manage it. They weren’t going to manage it. But they had to have, internal to the NSF, not only a guy who worried about the discipline of gravity, which was Rich Isaacson, but now that he was inheritor of a big project that was going to cost, at least on paper, something like $150 million by that time, that there had to be a guy there who would—only that job of looking how the money was being spent, and how the engineering was going, looking for problems.

And that was a guy named David Berley. I don’t know if you know David Berley, but you might. He was a guy who had worked at Los Alamos and had now a job at the NSF. And he was the manager—the NSF manager of the LIGO project. He’s the guy who would tell Rich Isaacson if things were going well or poorly in the project. So, he was the eyes and ears for the NSF. And he came to that CB&I meeting. It was Robbie Vogt and his chief engineer, Bill Althouse, the guy who I was liaison for, Larry Jones, and me. That was the group of people from Caltech and MIT that were there. And this one guy from the NSF, and one host of people from the company, Chicago Bridge & Iron. And the president, who had given us a bargain for building the prototype. They really gave us a bargain.

And the thing that happens that should never have happened—Dave Berley asked the question—this group, we're all sitting in a room—that had something to do with the money for the—it was a question to try to get information. It wasn’t a question being one of—it was a question that was not being critical. It was a question of, “In the future, how do you intend to do this?” And Robbie Vogt—something in him, that pushed a button that made him go nuts. I mean, completely nuts. He had a fit. And effectively—and the president kept looking at me. The president of the company, damn it! Because I knew him by the time I was out there, having spent a summer there. And he couldn't believe his ears. I mean, here was Robbie Vogt chewing the living shit out of Dave Berley for interfering in the project. Totally nuts!

Zierler:

You saw this coming, or this hit you like a—?

Weiss:

No! I didn't expect any of that. I was flabbergasted. I didn't even know how to smooth it down. And the president walks out and leaves—so we go back to the airport, and we're all in the car together—Larry Jones, Bill Althouse, Robbie Vogt, and me, going to the airport. And I said, sort of halfway down the trip—and this was a betrayal—I said to Robbie, “What you did at that meeting destroys us. We can’t let that happen. And if you can find a way that you can turn this over to Stan, that would be the best thing you can do.” And not another word was spoken for about half an hour, in that car. And Vogt, as we separate at the airport, he says, “Weiss, you always see everything in such a bad way” (laughter).

Zierler:

What was your response to that?

Weiss:

There was no response. I said, “We can’t do what you did at that meeting.” I think something like that. And shortly after that, NSF started calling me up saying, “Get MIT to talk to Caltech so that Everhart realizes that they should replace Robbie Vogt.” And so forth. So there was about three months of—and what had caused all the anxiety at the NSF was that he had had a fit with them, also, and one of the review committees—I was not aware of that—and effectively refused to let them look at his budget. Now, you can’t do that. And I don’t know what went on in his head about that, but he felt that he was being manipulated by the NSF. And all of this is going on simultaneously with him having trouble with his own faculty at—that already had died down a little bit.

So okay, much of that is behind the scenes. I had tremendous empathy for Robbie, because he did a lot of good things. He got us through a wonderful proposal we wrote. He organized the engineering as he could, in the way I thought was sensible. I'm unhappy that he had this battle with Drever but I understood it. But then why did he have to have a battle with everybody else? That made it very hard. And so, I felt that the project was being jeopardized again. And so, I had said what I said.

It wasn’t in any way in animosity to the poor man. I mean, he had a terrible history. He was a 13-year-old in Germany when he was—because he’s very tall, he was drawn into the Nazi army, as a boy. His father was killed by the Nazis. He was not a Jew. His father was a newspaper editor of one of the—I think in Salzburg; I don’t know exactly the town—who was anti-Nazi. He was taken prisoner and killed by the Nazis. And by the way, he became a prisoner of the Americans during the Second World War. When the Americans came into Germany, he happened to be taken prisoner, so he spent some number of years in Arizona in a prison camp! (laughter) I mean, he had very complicated history, that man. So, there is no bad guys, no good guys; it’s just a mess! And thank god for Barry Barish. Sensible (laughter). None of this craziness. And so, what he does, he takes over, he looks at the project and decides that—two things. He comes in with all this wonderful support from MIT and Caltech and everything. So, he’s in a position so he can demand things.

Zierler:

And is your sense that Barry has read up on everything that has been happening?

Weiss:

Barry knew pretty much everything that was going on. I don’t know if he knew every detail, but he knew Drever.

Zierler:

He knew what he was getting into, though, you're saying.

Weiss:

Absolutely. Absolutely. But he brought with him a very, very successful project manager, a guy named Gary Sanders, who had worked with him on the superconducting supercollider. And Gary Sanders is a physicist. In fact, he’s a MIT graduate, as a high-energy physicist. Gary Sanders. Barry was a graduate from Berkeley in high-energy physics. And the guy who did most of the work of getting things straightened out was Gary Sanders. Barry sitting way up but effectively be the symbol for, “This is what we have to do.” And he told people what it was, but then Gary would be the one who had to exercise it. He still was very busy with high-energy physics at the time. Barry Barish was a professor. Gary Sanders was a scientist. Barry was one of the high and mighties at Caltech, the professors, a big deal there. And Gary Sanders was a guy who was hired in to be the project manager, although he was a physicist in his own right, with a PhD and everything else. So, he was very competent; let’s put it that way. He knew a lot. So, Barry Barish didn't have to make any decision. That’s what I'm trying to tell you. It was that he had somebody he trusted could make decisions at a lower level.

Zierler:

And this was really missing, up until Barry’s involvement.

Weiss:

Right. Well, that, and Barry then immediately saw that the engineering office was too small, and he hired three more engineers. He realized that there was not enough modeling going on. Modeling. You know, most of the stuff that was going on in Drever’s group was, “Oh, let’s try this, and if it doesn't work, let’s try this.” But there wasn’t that much thinking going on. Why? Because the way Drever wanted to run—this is one of the troubles that Robbie ran into also. The way that Drever liked to have things, he would have some insight, come up with a design, and everybody would build it, and then they'd go on to the next one. There was not a hell of a lot of modeling, calculation being done. At MIT, there was more modeling done. In fact, I spent a lot of time modeling things. But we didn't have what I would call experimenters who enjoyed modeling at MIT either. Caltech didn't. But there are experimenters who enjoy working on a computer and modeling things. And that was what he brought in.

And he brought in a guy named Hiro Yamamoto, who did a lot of the optics modeling, a very good guy. And he brought in more people. He brought in a whole group from University of Florida, a guy named Guenakh Mitselmakher who was a buddy of Barry’s in the high energy business. He had just gotten a job at the University of Florida in Gainesville. And he convinced Mitselmakher, “Why don’t you join LIGO?” And Mitselmakher looked for a group that he could bring in to LIGO. And so suddenly now there was another university mixed up in this thing, not just MIT and Caltech.

That’s something, by the way, that NSF had asked for, over and over and over again. And what happened during the critical times of Robbie establishing himself, he thought he just barely had enough energy to solve the problem between Caltech and MIT to furthermore add in a lot more universities. But the NSF was demanding that, and in fact I was willing to put time into that to help make collaboration, a LIGO collaboration. That didn't happen under Robbie. It happened under Barry.

And Barry was completely—I mean, he knew that’s how you run big experiments. You had people in several big universities do much of the work, but then you have to have a collaboration that does—at the time, we didn't have a plan for how we would do the data analysis. We didn't have a plan for how we would actually run the detectors. A lot of the stuff that was in the future, he was thinking—thinking ahead. And he was looking at experiments in high-energy physics as a model, see? And there you do all this planning ahead of time.

So one of the things that Barry, very shortly after he took over, did was to get Boyce McDaniel, same Boyce that helped me out, and him—Barry and he, they knew each other well because they were both high-energy physicists, and Barry had an experiment at Cornell so he knew Boyce. And Boyce made the argument, along with Barry, that there should be a LIGO Scientific Collaboration, and that they would be responsible for the scientific output of the experiment, of LIGO. And then also that there would be groups within it that did astronomy, the astrophysics, that did the data analysis, and also looked ahead to what new things you would like to have the detector have. And that included people within the labs, within both MIT and Caltech, but it also included people who were from outside, who wanted to be in the LIGO collaboration. And that’s how we got to the point where instead of just about fifty people working, we're now up to 1,200 people in the collaboration, something like that, many of them very specialized. And that has its problems, too. At any rate, so that’s the beginning of Barry. And Barry then made some very, very fast decisions.

So, let me say first what were Robbie Vogt’s decisions. Robbie Vogt’s decision was pull the two groups together, have an engineering office. He also made some decisions on the instrument which caused a lot of trouble for my people at MIT, but there were two rival designs for one part of LIGO—how you take the light and bounce it back and forth in the arms. It’s a technical detail, but it was a very important one, namely we were working on something called delay lines, and so the Germans were, also, where you take the light and actually bounce it in discrete spots, a hundred times, back and forth, the four kilometers.

Ron had a different idea which is much more difficult to make work, but eventually, all of us together made it work, was to use a Fabry-Perot cavity, which has two flat plates, slightly rounded plates, like this, and you pass the light in from the front, and the beams are all on top of each other, and they go back and forth. Very much harder to make that work, but it gets you the same effect as though you had gone back and forth one hundred times. And one was called the Fabry-Perot system and the other one was called the delay line system. And we were still working on the delay line, all the way—and because the Germans had made that work better than anybody had ever seen anything before like that. It was ready to be put into LIGO, I thought. So, I was not going to go along with the Fabry-Perot, until you could prove that a Fabry-Perot worked as well as a delay line. And that’s something that Drever didn't want to accept. He didn't want to wait until that. He wanted a decision made. So that was a complication, there.

And so, what happened is that—and this is a tough decision that Robbie made—he decided that in the proposal that was now going to be made to the NSF to build LIGO, there would be two phases of that. One phase would be the one where we go in with what we have now, and call that the initial detector. And then there would be, after that was running and satisfactorily running, either having made a detection or not made a detection, that would then be replaced by something called the advanced detector. And the facility should be made so as to accept both the initial and the advanced detector. We were all part of that decision. It was a very good one. And so, based on that, he made the decision that the first detector to be put in there would be a Fabry-Perot. That caused a lot of dissention within the MIT group, but I tried to explain it to them. Some people, that was part of the decision. But see, Robbie didn't come to MIT to explain it to people.

Zierler:

What were you trying to explain, exactly?

Weiss:

Why the decision went with the Fabry-Perot. And it should have been—according to the people at MIT, and I didn't disagree with them, it should have gone with the delay line, because at the time, it was, oh, maybe a factor of three better, and more complete than the Fabry-Perot system. But the Fabry-Perot system had certain advantages. You could use a much smaller mirror. And the other thing that was known about the delay line is it had a problem, a scattering problem, that you could solve, but it was this problem that most people thought the Fabry-Perot wouldn't have, if you ever got it working properly. So that was a very complex decision he made. I understood it. Some of my people didn't understand it.

But then he made another decision which I didn't understand. And that caused me a lot of trouble. And that is we were part of the—I didn't tell you anything about the laser. There’s so many pieces to this! The laser that started this whole thing, we used a—your experience with lasers are what? Helium-neon or—do you remember any lasers that you worked with?

Zierler:

No.

Weiss:

Okay. Well, all right. The laser we started with was the argon laser, and it had a big problem. It was dramatically inefficient. Dramatically inefficient. For one watt of laser power out, you had to put 10,000 watts of power into it. So it had an efficiency of one ten thousandths, Okay?

Zierler:

And a pretty high electric bill to boot, I assume.

Weiss:

A fantastic electric bill! On top of that, it was a miserable laser. And at the time when we started this field, it was the only laser that we knew, and we could make one at MIT—I made one—that you could get enough power out of, so you could get—make a good—make a small—make a sensitive measurement of the position. So, when the people in Germany got into it, and when the people in Glasgow got into it, they decided to use the same laser. But I knew it was a horrible laser.

And there was a wonderful suggestion made at that thing that was held at the panel on interferometry. One of the people on lasers made a really wonderful speech about how yes, the argon laser is okay, but here is a laser which will absolutely kill the argon laser dead. And it was a solid-state laser. The only problem with it was it wasn’t visible; it was an infrared laser. And it convinced me of it. So, I immediately got some people in my group to start working on that. We started building an infrared laser and stuff like that. And it worked beautifully.

And so, I was convinced that when I was going to build a five-meter system, it would be using that one-micron or invisible laser. And the decision was made by Robbie—this is going now further forward—that yes, it would be the argon laser that we would use in the first—it was an absolutely hopeless decision, as far as I was concerned. It’s the power bill, alone. And the total inability of that thing to function without service once every couple of days. I mean, you couldn't have done anything like we're doing now. So, I argued with him about that. And he told me over and over again, “If you keep arguing with me, I'm going to quit.” And so I said, “All right, I won’t argue with you about it, but I will—side issue, we'll keep working on it in case you need it.” So, when Barry came in, Barry made the decision—I told you all the decisions that were made by Vogt—Barry made the decision right then and there to go to the one-micron laser. And he didn't do it because I talked to him, but he had talked to some people who he knew were experts in lasers, and they talked him out of the argon laser. Anyway, so that made some people at Caltech upset. But then, from then on in, I think it went very smoothly.

And I think I ought to quit. So, we’ve got still, ahead of us, the whole discovery, right? And of what happened under Barry. One thing we've left out of this whole thing—okay, I'm sorry—is the battle that took place which Robbie Vogt was involved with, in Washington, in Congress, with pushing through LIGO. It was a horrible situation. That was also in the press, by the way. That happened, by the way, that whole effort that he made to try to get the astronomers to agree that this was a useful thing, and the astronomers at that time were very powerful in the Congressional committees—is the story that the astronomers effectively tried to kill LIGO. And they almost managed. So, if you're interested in the politics of that, that’s an important piece.

Zierler:

Rai, I'm looking at the time right there, and I think that’s probably a good stopping point for picking up next time. So, we'll get into the politics for round four!

Weiss:

(Laughter) Well, that’s the politics is about twenty minutes at most. Maybe fifteen.

Zierler:

Yeah, but we're not going to finish today.

Weiss:

No, we're not. Oh, no, I know that. I know that.

Zierler:

So, let’s cut it here.

Weiss:

Okay, good.

 

[End 200621_0165_D]

[Begin 200628_0175_D]

 

Zierler:

Okay, this is David Zierler, oral historian for the American Institute of Physics. It is June 28th, 2020. I am so happy to be back with Professor Rainer Weiss. Rai, thank you so much for joining me for part four of our epic conversation!

Weiss:

(Laughter) Well, thank you. Thank you for having me. We had gotten to the point where Robbie Vogt had become director of the project, which was a good thing. And he had multiple problems. Not only did he have problems organizing the two groups together—the Caltech and the MIT group—but he also had a problem with Washington, which got worse and worse during the time that he was director. And it was promoted by—and here are some names, which I will stick with—it was promoted by John Bahcall, who had run the astro decadals. We had never been asked to be judged by the decadals. We were advised by the NSF, in particular Rich Isaacson, that there was no point in being judged by these decadal surveys of astronomy. And partly because they already were quite skeptical of this whole thing. And it started with Jerry Ostriker who was—most of this whole antagonism against this was concentrated at Bell Labs and at Princeton University. It really began with Jerry Ostriker not particularly liking Kip Thorne.

Zierler:

Personally, you mean.

Weiss:

Personally. That’s my own feeling. My suspicion is they shared a girlfriend or something, at some time (laughter).

Zierler:

Okay!

Weiss:

That’s just a suspicion. But the thing is that Ostriker was against the whole idea of gravitational waves, and on a level which was right at the deepest intellectual level. He contended not that the waves weren’t there, but that they will never be detected, and there are no sources for them. And John Bahcall had become this very important figure in American astronomy because, well, he was just a very good person. But he ran this decadal study. It was a whole list of things that they had made—I don’t know how well you know these decadal studies.

Zierler:

Sure.

Weiss:

Oh, you do? Okay. Well, they had made a list of things that they thought would be worthwhile supporting. And of course, here was the NSF, he had learned, not from the astronomy division but from the physics division, pushing a project that smelled of astronomy. And the problem with—I mean, I'll be very primitive about it—I'll just say that the problem that it smelled of astronomy was my fault. I was asked early on by Kip and others—we were trying to figure out what kind of an acronym should we use for the laser interferometer gravity wave—experiment? If you made that into an acronym, it sounded terrible. LIGE You couldn't pronounce it. It sounded something like an evil spirit. So, the next obvious thing to put on the end of that was an “O.” It was the laser interferometer gravity wave observatory.

Zierler:

There you go!

Weiss:

That smelled of astronomy, and that was my big mistake. Now, it’s a little bit pulling your leg, but this is what happened, and a lot of people actually in the business blamed us for doing that. That was dumb, to use the “O,” because that raised the hackles of the astronomers, who had just been through this decadal study, and had never mentioned or even heard about gravitational waves.

Zierler:

Why would the name LIGO set off this nerve among the astronomy community?

Weiss:

Because it has the word “observatory” in it.

Zierler:

Oh, I see. That’s their domain. You say “observatory”—

Weiss:

Astronomers have observatories! (Laughter) Physicists have experiments!

Zierler:

I don’t know; are we talking about PhDs, or are we in the playground here?

Weiss:

We're talking about what went on! (laughter) And you can ask others about this. I think that was the red flag, but there were deeper problems. And the deeper problems were still, irrespective of what you called it, even though we had been through this panel which we talked about last time, which we had survived, most people didn't know that. I mean, it wasn’t made very public that Garwin had suggested this, Garwin had agreed at the end that this was a good thing. That was not part of everybody’s thinking on Capitol Hill. So, John—and here is a third person in this thing who is very complicated, and that’s Tony Tyson. Do you know of him?

Zierler:

I don’t know.

Weiss:

Okay. He’s a very good scientist. He’s the youngest of this group consisting of Ostriker, Bahcall, and Tyson. He’s the youngest one. And he was in the most precarious situation in this whole thing. Because Bell Labs just about that time—he was at Bell Labs at the time, as an astronomer, and had converted a lot of astronomy into electronic astronomy, which was very important.

Zierler:

I didn't even know that Bell Labs had an astronomy program.

Weiss:

Yeah, I know. Well, that’s always the thing. Bell Labs discovered the three degrees, in fact, as you know. Penzias and Wilson. Had nothing to do with Bell Labs, either, except there was a communications stunt. So, what happened is that this was near the end of Bell Labs, because anybody could smell it—the divestment of the telephone co, the Bell Telephone—you know, the breakup of the company was imminent. And one of the things that everybody who was trying to break them up wanted was to get rid of the Bell Laboratories, because it gave them a tremendous arm for research. It kept them way ahead of everybody else. And Bell Labs was the equivalent of the best university in the country, in terms of research.

So, what happened is that Tony Tyson, who, while he was at Bell Labs, maybe 15 years earlier, had been one of the people who had been able to refute Weber—now, imagine that. He was a guy at Bell Labs, and as one of the projects that he did, he built a bar detector at Bell Labs. And he and David Douglass, who was at Rochester University, ran a coincidence experiment, and they saw nothing. That was very much close to the beginning. Tyson was one of the young Turks in physics at the time. He managed to pull this off in record time. Then he got interested in the rest of astronomy that would have results, which is using CCDs for doing cosmology, and he became very successful at that.

But he was precarious in the following sense: he knew that Bell Labs was about to go under. And I know he was looking for jobs, around. So, he was vulnerable to, for example—I don’t know this for a fact—what I'm about to tell you now is a surmise on my part. But because he was in this vulnerable situation, when John Bahcall started ranting about LIGO being a big mistake, he probably did go to Tony Tyson, because Tony was one of the few who had done gravitational wave work and was now doing real astronomy. You see?

And Tony I knew pretty well. I was moderately close to him. And I got a shocking email from him saying that, “Yeah, you might not like what I have to do.” That’s effectively what he was saying. And what it was was the three of them, Tony being the lead expert, because he had been in the field, they started going over to Capitol Hill, talking to people in committees. It was mostly Bahcall, but Tony was always the guy who did the technical description. Bahcall would make the statement that the NSF was going to waste a lot of money on something which smelled of astronomy but had very low success, and in fact had—and then he would turn to Tony to explain what was true—that, namely, we were still not quite at the sensitivity to do what we were saying, but we were getting there. But Tony would say, “Maybe it should wait. Wait until the technology is finished.” He wasn’t dead against it, he wasn’t completely supporting Bahcall, but he was giving the argument for why right now was not the right time to support this. And that’s all that Congress needed to hear, and all that John really wanted to do. He wanted to avoid it from happening, to cut into the NSF budget, especially where he thought it would be counted against astronomy as an expense.

And that—there was even New York Times editorials written about this, where Bahcall got a statement about how the NSF was going to waste money if they did this. It got very outspoken. And there was a lot of public press about the fact that the NSF was about to make a terrible mistake. And this all came from Bahcall, who got helped by Tony, by an argument that wasn’t against it, but it just says it was too soon. Very hard to argue against.

And so what happened is that a lot of time was spent by Robbie Vogt going to Congress. At the time, there was a search going on over all of the United States for where you might put LIGO. Instead of having—now we were forced into a situation—I told you last time, I think, that the people—we kids (Kip, Ron and Rai) had picked two sites. We had picked Edwards Air Force Base as one site, and that Cherryfield blueberry field in Maine. We had picked those two sites as where we were going to do it, to put the two detectors. In fact, the guy who became project manager was using that as a basis for what he was working on.

It turned out there was one story which was sort of funny, which I'll just tell now, which got the NSF into trouble, and that was just a little ahead of the Bahcall thing, but it didn't make any friends. It was the following. This was a piece of trying to help the LIGO cause internal to the NSF, but with a fairly significant mischievous outcome. What it had to do with was that there was a guy named Eric Block who was head of the NSF at the time. I can’t tell you exactly when that was. It must have been in about 1988, is my guess. Eric Block. He was an engineer. He had been trying—the people below him had been trying to get LIGO started. And one of the things they thought would be very good is if, for example, they found a site—neither of the two that I just mentioned, but they found a site where there was a powerful congressman. They wanted to play that game. And they had a piece of ammunition which was wonderful. At the West Virginia—the National Radio Observatory site in West Virginia, over Christmas of ’88 or ’89—I don’t know when that—yeah—I don’t know what Christmas it was, that epoch—one of the antennas collapsed. One of their new antennas collapsed.

Zierler:

What was it, a weather-related issue?

Weiss:

I think it was a fault in the design. I don’t know for sure. Nobody got killed. There was a guy in the control room, but he got away, of this very big dish. But the thing is, it was in—West Virginia had the guy who was head of appropriations in Congress, Robert—oh, boy. The first name is Robert; I'll get his last name in a minute. Yeah. That’s a pity that I can’t remember that, right off. He was head of appropriations for many, many years in Congress. He was from West Virginia. It’s very important that his name come up. We may have to look this up. But Robert—damn it! Okay! His first name certainly was Robert, and he came from West Virginia, and he was head of the appropriations committee.

Zierler:

This is in the Senate?

Weiss:

No, in the House.

Zierler:

In the House.

Weiss:

Where real appropriations get started is in the House. They only get approved by the Senate. So no, this guy was a very powerful congressman, because he was head of appropriations. Byrd! Robert Byrd! B-Y-R-D. That’s the guy. Fshoo! (laughter) And, well, so the NSF had this somewhat cuckoo idea, due to Eric Block, that yeah, we'll give them another antenna. That’s just all that Byrd wants. And that’s all he cared about. And so, they assumed it would make no difference to Byrd if it was the gravitational wave antenna rather than a radio antenna. You get the significance of this? It’s the fact that there was a thing—it would put big money into West Virginia, but it was for the wrong purpose for the people who were there. The people who had lost the antenna were radio astronomers, real astronomers, and they had for years lobbied to get more money for their site, but they always were doing only radio, and they had put themselves behind the mountains. You know, there’s a bunch of mountains that shield them from Washington D.C. and all the radio activity in Washington. They had done that. That’s why it was built there. And the last thing they wanted was a noisy thing like LIGO which would make radio noise. But more importantly, they wanted a replacement for the radio antenna, not a gravity wave antenna.

But what happened—one of the very first things that Robbie Vogt had to do is send some people from JPL there to see if you could shoehorn LIGO into that place. And you could. But it would have been a hell of an expensive thing, because there’s nothing flat there. But you could have got it in there. So that caused one hell of a lot of trouble with all of the astronomers. And that was before the Bahcall incident. Now, I don’t know how much that triggered Bahcall. I doubt if it did, because it never had gotten to the point where it was approved. But it made them angry. And so we had this little new legacy to deal with, sort of a blunder on the part of the NSF, to try to get this thing going by using the accident of the collapse of an antenna, and in Byrd’s district.

So anyway, enough. That preceded the story I just told you about Bahcall, Ostriker, and Tyson, but mostly Bahcall and Tyson going on Capitol Hill. That was important that you knew that, because it didn't make any good blood. So eventually it never really stopped, but the National Science Board, which is the head of the NSF, and through the director of the NSF, who was Walter Massey by that time, who was a Black man from the South, a very good man, they were saying effectively, “Screw the astronomers. This has nothing to do with them. We're going to go forward.” And they began to use the site survey that had been done by the project, which was one of the very first things—after that disaster with the radio antenna site, what Vogt did is he pulled together a committee internal to LIGO which looked at all states and sent letters to all the governors of all the states asking them if they were interested in having LIGO.

And that came to fruition, and there were five or six—and I was not part of that committee, because I was tainted by having already started working on Maine. And if Maine had won, somebody would have said that was an inside job. So, I was not part of it, so I can’t tell you all about that. A person who can tell you a lot more about that is Stan Whitcomb. You'll hear more about him in a minute. But he then had become again the deputy director of LIGO under Vogt. And he was in fact very much involved with that committee. And he’s still alive. He’s now retired by Caltech, but he’s busy with us writing that book that you got one chapter of.

And so, what happened is Stan and that committee sent I don’t know how many pairs—my guess is five or six pairs of sites—where the governors had said they wanted this and sent them to the NSF. And the NSF then had the choice of about four or five site pairs. And the final decision was made by Walter Massey. LIGO did not put a preference in the thing, which one would be the most preferential of the five sites or six sites. They were all considered equally good.

And it turned out the two sites that won were the LIGO site in Louisiana, where we did build, and the LIGO site in Hanford, Washington. Those were the ones that the NSF picked. There is a lot of controversy that I don’t know enough about, about that decision. Walter Massey is thought to have made that decision for two reasons. One was that the South had complained bitterly to the NSF that there were very few scientific facilities in the South. There were, up to LIGO, not very many. There were a lot of military facilities, but big NSF facilities, there were very few. And DOE hadn’t put a lot of facilities in the South, either. So, it was a way of doing something for the South, putting it in Louisiana, so that’s what Walter Massey felt, I believe.

And the other thing is it had powerful senators. One of the senators, a guy from Louisiana, was Bennett Johnston. he was deep into American scientific policy. He was a guy who pushed for the supercollider. He had done a lot of things. And when this came as an opportunity to put a major site into Louisiana, he really went to town. So, Johnston was one of the guys. And the other guy was the guy from—I've forgotten his name—he has always been called “the senator from Boeing.” He was a Democrat. He’s not as important. The guy who was really important was Johnston. But turns out the senator from Washington was no second sister, let me tell you, but I cannot remember his name, because I never really interacted with him. But he was the senator from Boeing, and all you need to do is look up the senator at about the time of 1991 from—yeah. Do you remember? Do you remember the senators from Washington state at all?

Zierler:

I'm a young guy. I may be in sixth grade when—

Weiss:

Okay (laughter). Sorry!

Zierler:

That’s all right!

Weiss:

Okay. If that name comes to me, I'll tell it to you. So what happened is that the people who got us out of the mess with Bahcall and mostly Bahcall, and the attack, was Walter Massey, and later Neal Lane, who was the other director of the NSF, who just kept telling the astronomers, “Get the hell outta here. We're building this. We want to build it, and these are the reasons, and we're not going to have to—you can’t turn something off. And if you have a better idea, tell us, but we're not turning this off.” That’s pretty much the way they approached this. So, we were very thankful to them for doing that, and it pretty much became quiescent after that. Okay. So that’s the story of the political thing, okay? There’s nothing more to it that I know.

Zierler:

Rai, just for the context here, are you following this in real time, or is the narrative that you're presenting now your capacity of piecing these items together in retrospect?

Weiss:

Well, it’s a little of both. Look, I lived through it. The trouble—I was not part of the lobbying. I had lobbied Maine. I had done a lot of work on Maine, when it was still our decision. But I did not get mixed up in any of these site decisions. In fact, once Robbie Vogt became director of the project, I pulled back all political activity. I had no business—I didn't feel anymore that I was useful in that level, and I stepped back into doing science and engineering. That’s really what happened to me. And I think that Vogt was very happy that I did not try to horn in on his activities. He would have been I think insulted if I had tried to get mixed up with that. And he didn't ask me for advice, which bothered me a little, but that’s okay.

Okay, so the next big event that happens—I don’t know, I think we may have talked about that—is Robbie Vogt organizes the project. He organizes the two groups, gives them both things to do, which was unheard of. I lost some people because of that. Caltech, not at the time. And he made some decisions we couldn't make. Some of the important decisions were the configuration of the detector, for example. This had been not a competition, but a question, between the MIT group and the Caltech group. Should it be a particular—it’s not the idea of the detector, but a certain optical configuration in the detector. And one of them was something called an optical cavity, which is a way of storing the light, and that was called the Fabry-Perot cavity, and Ron Drever was an advocate for that. And I was an advocate for the delay line, which was a system with two mirrors again, but discrete spots. You could see the individual spots as they bounced back and forth. And the Germans had done a beautiful job of that, much better than we at MIT had done, and they demonstrated this thing on a thirty-meter prototype, and it looked like it was really working well, whereas the Fabry-Perot had not really worked well in the hands of either the Scots or the Caltech group. But it had a very clear advantage, should it work; it could use small mirrors. It didn't have to have a big mirror, which had all the spots distributed on them. All the spots were on top of each other, so the mirror could be much smaller, and cheaper thereby, and also easier to fabricate. So, it had that, and one technical advantage, which could be solved other ways, but the light in the Fabry-Perot didn't scatter as much as it did in the delay line. But you could fix that by using tricks with the laser.

But nevertheless, a decision had to be made. And that was one of the requirements from the panel on interferometry. That was one of the things they said—“You've got to make a decision on that detector.” And Robbie made that decision. He also made the decision on the laser, which I thought was a mistake, which was later corrected by Barish, but nevertheless it caused a lot of worrying in my group, because we had committed to changing the laser to a more efficient laser. I think I told you a little about that. okay, so those were the big decisions that he had made in the beginning. And now here was Robbie having to deal with this problem of the rebellion on Capitol Hill and now the sites. And then what happened is little by little, he got himself into a terrible squeeze with Ron Drever, who is a complicated person. I don’t think we've talked all that much about him.

Zierler:

No.

Weiss:

And he should be talked about, because he was a guy who contributed a lot of ideas to this, but he was also a person who was very difficult to work with. And let’s leave him go for a moment until later, because I learned only later why he was so difficult. It had to do with—I think we talked about that, maybe?—he was an Asperger’s syndrome person.

Zierler:

Right.

Weiss:

Yeah, I think we did talk a little about it.

Zierler:

Right.

Weiss:

Okay. So, well, if we come back to it, if there’s time for that. But the thing is that what happened is Ron—I had trouble with Ron, the project manager that came from JPL had trouble with Ron. Even some of the people in the group who worked at Caltech had trouble with him. And one of the things that was most troublesome—and Robbie knew this and effectively tried to solve it, but he did it in a brutal way. What happened is that I think the most serious problem is the people in the Caltech group could not work on a single idea long enough to bring it into fruition before Ron had a different idea which he thought was better. And this business of continually changing what would be the project you're working on for the people in that group was devastating. And because of that, they made not as much progress as they should have.

And here we were, all working together now. I mean, Caltech and MIT’s groups were working on the same problems, on one system. There was no longer any of this competition between the two groups because that had not been solved but it had been suppressed. And what came of it was that you couldn't keep doing that—every day coming in with a new idea—when the group had gotten so big that you were disturbing all the effort across something like forty people.

And what happened is Vogt—Ron refused to accept the idea that he would be controlled by—he felt that he was an expert, and Robbie was not an expert in this particular thing, and so his ideas should dominate, because his ideas were better than the ideas that Robbie was pushing.

Zierler:

And Rai, where are you in this? Are you running interference between the two of them? Are you mostly on the sidelines?

Weiss:

No, no. I was never running interference between the two of them, because I was not aware of all of this. See, I was busy at MIT dealing with two things—getting the jobs we had been asked to do done. And those jobs were things that Robbie asked us to do, and I had gotten extremely busy with the beam tube. And I'll get to that. Making the infrastructure that actually made LIGO. And so, I was not involved with the day-to-day activities at Caltech. I knew very little of it. I don’t think I could have solved anything anyway, because what happened is that Robbie eventually threw Drever out of the project.

And that hit every newspaper. I think I may have said a little to you about that before. It was a scandal of first order, and the faculty at Caltech rebelled against it, and the NSF got very worried, and even MIT got worried, a little bit later. Because now by this time, MIT was for this project. You know, in the early days, they were not, but now they were. Or they had no control over it. It was in the hands of Vogt. So, this very unpleasant thing happened that Vogt effectively became totally ineffective, because he had to do battle with the Caltech faculty, or so he felt he had to. On top of that, he had to keep this thing operating in some way.

And then he was also mixed up now in a primary contract. A contract had been let, because the money had been given for the project. And I don’t know; I think I had mentioned to you what the contracts were, but I'll just say what they were again, because that’s what I was involved with. One was to make the beam tube, and that was Chicago Bridge & Iron Company. The other one was to make the vacuum system you put all the equipment into, and that was a company called PSI—Process Systems…I don’t know what the last thing stood for. And then there was a contract with Parsons Engineering for the buildings and civil work at the sites. All these contracts had been let.

And the thing is, he was now dealing with a very complicated situation which he was not really good at. This is Robbie. What happened is the NSF was very worried that because they saw this terrible problem happening at Caltech, they would send teams in to make sure that their investment—the NSF now would send committees to NSF [sic] and to MIT, but mostly—and we’d all have to come to converge at NSF, making sure that the project was still in good shape, despite the craziness that was going on at the very top. And they were not satisfied with that, because they found that the project was not spending the money at the rate that they thought it should be. The committees were told that so much money would be spent every year, and that the money was not being spent at that rate. And Robbie didn't feel that was a bad thing. To him, it was—but the NSF couldn't deal with that. They had taken that money from other projects internal to the NSF, and they wanted to make sure that now that they were pushing this project, the money got spent. In other words, it’s a funny situation to be in. They didn't want to make the project last longer; that would be more expensive. But the felt the project wasn’t being managed properly, because it wasn’t spending the money. And Robbie couldn't see it that way. Almost everybody else in the world saw it that way. Because Robbie was so busy with all these fires that he had started.

So eventually what happened—there was a big move by the NSF, which MIT became part of it in the sense that they saw the danger of it—and the president of Caltech, who—I had his name the last time we talked about this— I think was his name. Everhart. Everhart, excuse me. Everhart. And they made a move now to replace Robbie. The MIT president was involved in that. The NSF was involved in that. People at the NSF were involved—

Zierler:

Were you aware of this at the time?

Weiss:

Oh yeah, I was fully aware of this. And in fact, I ran into a problem of my own, which made me convinced that—and this was one of the hardest decisions I ever had to make—is—I was—I think I may have told you this, but I'm not sure—I'll get to myself in a minute. But everybody at the top level knew this. My own experience was we were having a meeting at Chicago Bridge & Iron, and Robbie came to the meeting. It was one of these biggest contracts, $50 million contract. And Larry Jones, who was an engineer, and I were the guys who were in charge of the project.

But Robbie Vogt came to one of the first meetings to decide—they had just built a model of the beam tube, which is that big tube that runs from one end of the apparatus to the other. You know, four-kilometer beam tube, which was a very big, tricky item. Many of the experts at the NSF who had been asked to get opinions from had decided that the way we were approaching that was too risky. It came out of a prototype we had built at Caltech. I was out there a lot at that time. This was once Robbie became director, I spent a lot of time at Caltech. But not with the politics; with actually this prototype for the big beam tube.

And we built it, and we had all sorts of new things—I don’t think I want to go into those. There were all sorts of money, economical things you could do to make vacuum systems cheaper than previous vacuum systems. Because if we had to do it just the way people who had built large accelerator vacuum systems, it would have used up all the money in the project. We couldn't afford that. We had to make some changes. And we made significant changes, and it was not something that the people who had been vacuum experts had done themselves. There was no experience with it. So, we felt we had to build a prototype. And we did. And then when we got industry involved in it, we felt again that it had to be built all over again, another prototype, to see if industry could make it work.

And Robbie Vogt showed up at a meeting at CB&I, after we had spent the whole summer with—I had spent a lot of the summer in Chicago, or in Plainfield, near Chicago, getting this prototype working. And this was a report that they were going to give now to the LIGO project on what they learned to make the prototype and what that was all—for them was—if they did a good showing, they were going to get the contract to build the beam tube. If it didn't look good, they were not going to get the contract to build the beam tube. That was the way the contract was written. The contract with CB&I was written, “Build the prototype. Then we will decide—” They got some $10 million for that, or whatever, and they would get the rest of the money or the $50 million to build after successful manufacture of that prototype. They were very happy with that, and it turned out it was a fixed price contract that they were forced into. And they were very, very worried about it. The industry was. So, they were very happy that they could show that they could do it, and within the money that they thought they needed.

So, an NSF representative named David Berley—and he’s still around, if you want to hear—he’s a guy you may want to talk to. He can tell you reams about this. But he was Rich Isaacson’s parallel on the side of the project direction in the NSF. You can call him the manager of LIGO in the NSF. Where Robbie Vogt was the manager of LIGO, this was the guy who observed what was going on for Rich Isaacson in the NSF. Rich had so many other things to do, he couldn't spend that much time looking only at LIGO.

So anyway, he came, and then a terrible thing happened—I think I may have said that, but then we'll quickly get it done with—is that we're all sitting around a table. The president of the company. Everybody has presented. And Dave Berley asks a question. Had to do with some financial thing—“How would you do this particular thing in the future?” And Robbie Vogt had an absolute fit. He had a fit. I mean, a real fit like he was nuts. In front of all these people. Effectively accusing Dave of interfering with his running of the project, and how dare he ask a question like that, and so forth and so on. And everybody in the room including myself was absolutely flabbergasted. Here’s the guy in charge of the money (laughter) at the NSF, asked what was a perfectly reasonable question, and Robbie had an absolute fit.

Zierler:

Do you think this was justified, how Robbie reacted?

Weiss:

No. Absolutely not. It’s all coming back to me; I think we discussed this before. But what came of it was that—in the end, when I thought about it, I could excuse what was going on, but I could no longer have faith that Robbie could run that project. Why? Because you shouldn't have a fit like that, even though you are under tremendous pressure. He was under pressure from the Caltech faculty. He was under pressure from the MIT administration. From the Caltech administration. All this madness about him firing Drever. It happened all at the same time. And now this meeting at Chicago Bridge & Iron. So, it turned out on the way—and this was fairly late, and before Barry Barish took over—what happened was the four of us were driving from Chicago Bridge & Iron to the airport in Chicago—O’Hare—and there was not much conversation in the car. But I finally had to say something. And in the car was the chief engineer, a guy they were training to become the chief engineer, Robbie Vogt who was in the front, and me in the back. And I told Robbie that what had happened at that Chicago Bridge & Iron meeting was not something that I could understand, and I don’t think anybody in that room could understand. It was not—having—and I told him point blank that the best thing he could do right now is to get Stan Whitcomb, the deputy director, to become the director, and if he wanted to stay on the project, working on something else. That didn't take any—that was not something that—well, I had to do it, because I was so disturbed by the whole thing, and because I could not see him succeeding, and I didn't want the project to go under. And then nobody talked about another thing, and nobody said a word until we got to the airport. And I think now it comes back to me; I think I told you this. Robbie told me that in his experience with me was I always saw everything in the most negative way. Okay? So.

That was probably a month before he actually had to leave the project. Leaving the direction—of being director of the project. He stayed on the project. And that’s when Barry Barish came in. And that was a breath of fresh air. And what Barry did is his experience had been—he’s a high-energy physicist. He had been working on the supercollider, the one in Texas. When that got cancelled by Clinton, he and another scientist that worked with him became free. I don’t know how his name got into his hopper, but he was asked by the president of Caltech to take over the direction of LIGO, and he took it. He also brought with him, then, a very good guy named Gary Sanders, who was a physicist in his own right but acted as project manager. Barry was too busy with everything else in the world, but he spent enough time on this so he could reorganize it. But much of the day-to-day problems were handled by Gary Sanders.

And then he increased the project office, made it bigger, brought more engineers in. He made a decision on the laser. That bad decision that was made by Vogt was undone, and then now we're going to use a solid-state laser. Thank god for that. And then he began—other things he saw were deficient in LIGO, as it was at that time. He started a lot more modeling and systems engineering, which is important for a project like that. So, all the things he did were positive, but they made the project a little more expensive. And from then on in, it went absolutely—I mean, most of it was smooth as silk. People have asked, “Why didn't Barry get the Nobel Prize?” Barry should have gotten the Nobel Prize. He saved that project. So, what then happened next? Okay. What would be next would be a little bit about description of how we built the detector.

Zierler:

Do you think, Rai, that Barry is an unsung hero here, in terms of how important he was relative to the appreciation that people have about his contributions?

Weiss:

No. I think he has been well acknowledged. I do. I think some people—I'll tell you, the Nobel Prize thing was—it would have been very complicated. This is immediately after the discovery, OKAY? And Ron Drever, who was still alive at the time—and Ron Drever, Kip Thorne, and I got a lot of awards from a lot of other people. You know, other big famous things. And I could see that Barry was very upset by that. I in fact started a whole—because I didn't know quite what to do about it. I couldn't tell people who to give prizes to. But the thing is I felt that that was overlooking just—this story, you just asked, had he been acknowledged—I felt he had not, at that time, been acknowledged.

Zierler:

At that time, right.

Weiss:

Yeah. But he became acknowledged. And the thing is that—and much of us, because we made a big fuss about it, both Kip and I, and people at Caltech made a fuss about it, I think. For example, some of the money that I was given in these other prizes, I turned into a fellowship for graduate students at MIT, and it’s called the Barish-Weiss Fellowship. I just wanted to point out—I don’t know if he ever felt that that was compensation, but it made him feel a little better, okay? (laughter) So, anyway. But what happened is Ron Drever died before the Nobel Prize was given, and that avoided the problem. Because he probably would have gotten the Nobel Prize, also. Anyways, it’s an unhappy—but I think in the end, that was a thing that solved itself, that problem, okay? (laughter)

Zierler:

In a very dark way, I guess he did everyone else a favor.

Weiss:

In a very dark way, that’s right. It’s not one of the better moments in this whole thing. And it’s part of the craziness about the Nobel Prize. Barry should have had it right from the start. But, OKAY. And Drever deserves some of it. It’s too many people, you know? But then when you look at it—and I always felt very embarrassed about this—if you look at it in a straightforward way, you find out that there are many others that are equivalent to the three who got this thing, in the project. And the only way I could live with this is by, whenever I go around, I say, “Look, I'm a symbol of that project. I'm not the one who got awarded.” I think there are at least ten other people who could have been me. I hate to say it. I don’t hate to say it. I mean, I think it’s not just modesty. I think it’s true.

Zierler:

There’s always a very difficult dividing line in these matters, in terms of who gets it and who doesn't.

Weiss:

I know. And it’s a destructive thing, and I've never been happy about that. And when COBE—I don’t know; that’s a long story. That’s a separate story. So anyway, so once everything calmed down—and I would say things calmed down by 1994; that’s when Barry took over—I turned entirely to that beam tube project. My group at MIT was assigned all sorts of interesting things to work on by Barry and by Gary Sanders. I mean, they discussed it with me, but at that time, it no longer mattered if a thing came from MIT or it came from Caltech. In fact, the thing that Barry did, which I better say right off, because it changed things again, is that he started something called the LIGO Scientific Collaboration, which is something that the NSF had wanted for years, but Robbie Vogt couldn't bring himself to do it, because it would make life even more complicated, having a whole group of people outside of MIT and Caltech also now wanting to do research and doing things and being part of the planning. And it was more than he could handle, I felt.

And Barry, on the other hand, had done that over and over again. He had enough self-assurance that he would—he in fact saw that we didn't have enough people within both MIT and Caltech, that he brought more people in on his own. For example, he introduced the University of Florida in Gainesville. They got brought in to build a particular part of the LIGO detector. Then later on, it turned out that when you look at the planning that was done by LIGO, there wasn’t real planning for the data analysis, or how the data would be analyzed, or how it would be stored, or what kind of searches you would do. What kind of science would you find? You know, people are too busy making things. And he saw that this would be a disaster if you didn't bring people in who would actually do the science that’s going to come from this. I was completely on board with him about that. I had one discussion with Robbie once, about that; he just didn't want to hear about it.

Zierler:

Rai, isn’t this a little late in the game for these kinds of questions to be coming up?

Weiss:

I agree with you. I agree with you.

Zierler:

What do you think accounts for that? Is it just there’s too many cooks in the kitchen?

Weiss:

It was disorganized. Just plain disorganized. If I had been director—not that I could have—we would have had that right away. Because I had run COBE. And the only way we got away with it is by putting more people on it than just the very people who started it. Well, John Mather told you that, probably.

Zierler:

Right, sure.

Weiss:

No, it was quite clear to me that we had to do something, but it was inconceivable to put Robbie into a situation where he barely was able to hold it all together with all the problems he was having already with it. And that’s another reason why I think it was so critical that Barry came in. So, what Barry did—he didn't know enough about the technology, although he learned it pretty quickly. He didn't know enough about the science. He learned that fairly fast. But he asked me to be the—once he got the LIGO Scientific Collaboration organized—and it was organized by—with one of the guys who was on the panel, Boyce McDaniel.

Boyce McDaniel was a guy who was—Andy Sessler and Boyce McDaniel ran that panel back in 1986 that I talked to you about. But it turns out Boyce, who was at Cornell, was a close friend of Barry’s. And they together went to the NSF to push for what’s called the LIGO Scientific Collaboration. And what does the LIGO Scientific Collaboration do? It does all the stuff that people realize in the project. They write the papers. That means the people on LIGO at Caltech and MIT, they are part of the LIGO Scientific Collaboration. It’s not that they're outside of it. They're part of it. So the papers are written by them. The data is analyzed by them. Those are the two big things—papers written, the analysis. And then committees to look at what are the interesting science. Four different committees were formed, internal to the collaboration, so you could think ahead how you would change things. I was the first spokesperson for that, and I in fact set up much of that committee structure and all of that.

Zierler:

And what were your goals at that point in terms of communicating more broadly what was happening?

Weiss:

Well, no, my goal was very simple. I said we were undermanned. The Caltech and MIT people, we didn't have enough people to do the whole job. And I said that to me that was essential, that it was organized so there would be groups of people who would do the data analysis. That was the first group. And then we’d divide it up into different sources—analysis for things like black hole pairs, analysis for things that are just continuously putting out gravitational waves, analysis for things that were making noise, gravitational wave noise. Another group was organized to look ahead how you would improve the detector. What science would benefit if you improved the detector in certain frequencies and certain ways. In other words, the LIGO Scientific Collaboration was used as a gigantic advisory committee. And the scientists who actually would present the data to the public. And it solved the problem which we hadn’t solved. So, as you say, it was a little ass backwards, but it was necessary. It had to be done. And Barry deserves the credit for having seen that that should be.

And then that has continued on. I was appointed the first one of them, and then I think it has now gone through five spokespeople. They have been elected rather than appointed. So, the thing has grown from maybe 150 people to 1,500 people. And they keep—the ways that people can get into the collaboration by saying they will do a certain kind of thing. They get voted on by the group that’s there, and if the group votes them positively, they can get admitted. But then they have to do what they said they were going to do. You don’t come into it just as being a player. You have to come in as a doer. You do something, you say you're going to do it, and then you get considered again every once in a while. You have a memorandum of understanding. Your group has a memorandum of understanding with the LIGO Scientific Collaboration. And every year, you get judged—have you done those things that you said you would? And, we've had problems. Look, no group that size hasn’t got problems. There are people who would like now to leave but because they feel they're being controlled. I mean, this is something now. But they can’t get at the fresh data without being in the group. So that’s the compensation. Anyway, that’s a whole other story, the LIGO Scientific Collaboration.

Then, the making of the detector—I think that I'll leave you—I mean, I spent a lot of time, but only after the beam tubes were built. We had to build the beam tubes both at Louisiana and Hanford, and I could finally get free of that by about 2000. Yeah. And at that point, the detectors were beginning to take shape at the sites. They weren’t yet by any means finished. And Stan Whitcomb became what’s called the commissioner or in charge of getting the detector working at Hanford. He spent probably two years solid at the site doing that. And then there were people hired specifically who were the commissioners. I'll give you two names that are important. One is Daniel Sigg, S-I-G-G, and he became the head of commissioning, and still is, at the Hanford site. And then a guy named Valera Frolov is in the same situation at Louisiana. Now the problem we had when I got off the beam tube and started working on the detector directly was that we had trouble getting people to work in the southern site. It was much easier to get people to work in Hanford, a lot harder to get people to work at Louisiana.

Zierler:

What, it’s just a less attractive place to be?

Weiss:

Well, a lot of things. Nobody seemed to care about the radioactivity. That didn't seem to bother anybody (laughter). At Hanford. And it still—I know people think about it, but it’s not something you think about every day. No, the problem was that it’s a—it’s a more interesting site. The weather is better. There’s more going on up there. It’s a better place to bring up kids. All in all, it’s—the South is—look, it’s a backward part of the country. And we were in the styx. The biggest thing around us is Baton Rouge. And the people who—and New Orleans. Now, New Orleans is a very interesting place, but you can’t live in New Orleans and work at the site. That’s too big a commute. You could do it, but it’s a two-hour commute. On the other hand, you can live in Baton Rouge and work at the site. That’s about a forty-five-minute commute. And, well, I don’t know—the people who were good wanted to work at Hanford (laughter). That’s what happened!

And so, we had ourselves a bit of a disaster. Barry recognized the disaster, and that is that the detector at Hanford was moving much more quickly than the detector in Louisiana. I won’t go into all the bits, but it was very clear. You could see it from reports that were coming from the sites. One reason was that we finished the beam tubes at Hanford before we finished them at Livingston. You know, Louisiana. So consequently, there was a six-month difference, maybe even a three quarter of a year difference between the beginnings of things at the two sites. But even compensating for that, it looked like Louisiana was going very much more slowly. And it turned out there was a lot of squabbling going on at that site, between people who were not being well advised.

I was there at the time—because I knew—we had just finished putting the beam tube in there, and I began to realize that there was not a very good spirit at that site, or a very different kind of spirit than at the site in Washington state. So, Barry and I had a long conversation. He asked me to take over being director of commissioning at that site. And I did it. But I still had a job, a professor job (laughter) and I still had the family here in Boston. It was tough for a while.

Zierler:

There was no extended leave of absence that was in the cards for you.

Weiss:

Well, I took whatever I could. I took my sabbatical during that time. But you don’t get more sabbaticals. Anyway, what happened is it was—it had to be done. So, what I would do is spend sort of half a month but not all continuously in Louisiana, making sure things got started and well done. And I took a very good student that I had at MIT with me and planted him there. And he didn't mind. And that’s a guy who now is a full professor at Caltech. His name is Rana Adhikari, and he’s an interesting guy. Well, between him and me, although I think more to him than me, we got that place shaped up. I can tell you reams of stories about Rana’s ability to do that. But what we were dealing with was a southern mindset. You know, the buccaneer mindset. I don’t know how much you know about the South, but—have you ever worked there?

Zierler:

I haven't, but you know, it’s the South. It has a reputation.

Weiss:

(Laughter) Ain’t that true. And it sure had! And we happened to have some guys, and they were cliquish as hell. You bring a person in there who they didn't like, and it turned out that person got nowhere. I mean, I'm talking about a—you bring a student down there, and they didn't like the student; the student just couldn't flourish. There was a cabal. They decided, “We don’t like that guy. The hell with him.” And well, he didn't—you know, he wouldn't talk about guns. He wouldn't talk about—I don’t think racism was yet on top of things. We had some problems with that, but they weren’t overt.

Anyway, so eventually the guy who made things really stabilized there was this guy Valera Frolov, who has been there ever since. And he was a guy who Barry had found in the high-energy business and turned out that he talked him into working on gravitational waves. A lot of the very good people we have, top scientists we have, didn't start in gravitational wave physics because it didn't exist. They were either laser people, or they were spectroscopists, or atomic physicists, or nuclear physicists. But there’s no gravity wave physicists. They didn't exist. And so, we had to generate those ourselves. At any rate, so OKAY, let me get to the end of initial LIGO. We worked on initial LIGO for about seven, eight years. And I'll say it to you in the most direct way: we got a very good nothing.

Zierler:

A very good nothing.

Weiss:

Let me explain to you what a very good nothing means, okay? And you'll appreciate why it’s the right words. The sensitivity of the detector had improved by at least eight orders of magnitude from the beginning to the point where we decided, okay, we can’t do any more with this detector. Now, we were at a level—and that was true at both Hanford and Livingston. Now, we had been working—and this is one of the decisions that was made early on; it was already true when Robbie Vogt took over, long before Barry did—that this was going to be a two-step process. Well, you ought to know this. Anyway, there would be a detector that’s built first, and then there would be a second detector that’s built, and that would be called the advanced detector. And that was already in process. The research for that was in process pretty much the day that we started building the initial detector, in the labs at MIT and Caltech, and also now in Florida. And in Britain, and the Scots. A lot of people around the world were beginning to do things for LIGO, and they were working on things for the advanced detector. The initial detector was all our doing. But so that there would be progress, we kept that research active as we were totally committed to working on the initial detector.

And so we were able to make a decision when we thought we had gotten to the point where we could no longer see how to make improvements in the initial detector that we declared victory and we went on and wanted to go to the advanced detector and install it. That required two things. That required that we understood why we were stuck at the level we were in the initial detector. We had to be able to explain that to ourselves. What was the reason we couldn't do improvement anymore? And the other one is that the improvements that we had been thinking about had actually been made in the laboratories. And they converged in around 2008.

And why do I call it a very good nothing? We could say without anybody refuting it that we had measured no gravitational waves at a level of a certain number. We understood the detector well enough and understood the noise in the detector well enough, and the reason why you couldn't go any further, and nothing in that noise we could say looked like a gravitational wave. And we could say to people with a very high assurance, sort of ninety-nine percent, that we saw—had seen nothing, at a level where people thought maybe you should see something.

Zierler:

OKAY, so explain the significance then. Why is it significance that you know you didn't see anything?

Weiss:

Well, because what it says is there aren’t any—if we had seen a lot of things we never could explain, and that sort of bubbled up over the noise, then you would say, “Ah, that could have been a gravitational wave. Why didn't you chase it?” We didn't have that situation. Every little bump we had, we could say, “That came from a particular cause in the detector.” And we could prove it. So yes, there might have been one chance in a million—I don’t know what—but we had enough evidence that everything we saw that was anomalous could be attributed to a particular cause, and we could make that cause come and go at our will. But we couldn't improve things a lot. See?

So consequently, we had a very different situation than Weber had. We had a situation where he couldn't say if what he saw was a gravitational wave. He saw a lot of anomalous signals, but they happened to be coincident. We saw no coincidences. That was number one. Very important. And we saw none of these anomalous things happening together at the two sites. And then even then we could explain what the anomalous things were. So, we felt that we could make a good case for why we should go forward and build the new detector. And we set a limit, which was I would say by the time 2008 had come, many people in astronomy had already decided that you had to do better than what we were doing to make a detector. But we didn't know that in 2000. But by 2008, we knew that.

Zierler:

But in terms of keeping momentum going, this is actually quite useful.

Weiss:

What is quite useful?

Zierler:

The very useful nothing. The fact that—

Weiss:

Oh, yeah! No, it’s damn useful. It’s very important. It’s very different than getting up on a bar, which is what people did, and having a lot of noise in the bar, and saying, “Well, much of what we have could be gravitational waves, but we don’t know.”

Zierler:

And are you starting, Rai, at this point—some of the naysayers who have been naysaying from the beginning, are some of them starting to come around at this point?

Weiss:

The fact that we could make the detector work at all, at a level which nobody believed, was enough to shut up almost everybody. There was still grumbling about should you put more money into it. Because the advanced detector cost about equal to what the initial detector cost, in terms of hardware. It’s about $100 million. Look, by the time LIGO is over, getting to now, we've spent one billion—$1 billion—on LIGO. OKAY? Now that includes operating it and all the—operating it is very expensive. It’s about $50 million a year, is operating it. And we've been in operation for about twenty years now. So, a good fraction of that billion is operations, almost all of it. I hope I did that right. Yeah, twenty. So, the actual chunks of improvement are hundreds of millions at a time. So, we've only had—so probably it’s a little more than a billion. It’s probably 1.2 billion, something like that.

So yes, at least half the argument that you could not—we were already at the stage where we were measuring the motion of the mirrors to a part in thousandths of a proton diameter when we quit working on it, on the initial detector. What we had to do is we had to change—what made the difference, and we'll get to that in a minute—was we got that sensitivity in a different frequency band. See, we got that sensitivity in a place where the sources weren’t evidently in the initial detector. And in the advanced detector, we got that sensitivity improvement and lower frequencies, where it turned out there were sources. So, we had already shown the technology. People didn't believe the technology. How could you ever measure such a small thing with a big device like this? It’s inconceivable. And that was always on the minds of people like Bahcall. Not so much Tony Tyson. Tony felt we would eventually get there, but he didn't think we would get there that fast. I think that was his worry.

Okay, but we hadn’t detected anything, okay? So, you can’t yet go and crow about it. That’s too subtle for most people to understand that. So okay, what were the things that were changed? We improved the suspensions, which is the things that hold the mirrors. And we put in completely new ground noise isolation systems, systems that take out the seismic noise that shake the mirrors. And those were the major improvements that were made. In other words, we improved—whereas the initial detector was sort of the very best, it was sort of around maybe 200 hertz was sort of the best place, by now, the best place was around seventy hertz. And if you look at what we discovered, a lot of the signals we see are the ones that start down at about fifty hertz and wind up at about 200 hertz. That’s where a lot of the best signals we have, have been discovered. So, improving the low-frequency noise, improving the laser power a little bit, and now, adding to it various ideas like squeezing light, which is using a quantum tailored light coming in one of the ports. And that’s a long story; that’s a technical story that’s very interesting, but I think at this moment, let’s forget about it for a moment, okay?

But the biggest improvements were the low-frequency suspensions were better, ground noise isolation was better, more power was applied. That works everywhere. And the last thing was using quantum tailored light. That just happened this year or two years ago. A lot of people have been working on that for years and years in advance. But applying it only happened about a year and a half ago.

Zierler:

Rai, can you talk a little bit about, as you're doing these technical tweaks, as you're improving everything, how do you know what you're improving it for? In other words, if your big goal up to this point is the detection of nothing, this valuable nothing, how do you know how to improve sensitivity and technical capability if you haven't yet seen something that you know to see better?

Weiss:

Well, you ask a very, very good question. That’s a first-class question. And it’s going to haunt us even now, okay?

Zierler:

Okay, good (laughter).

Weiss:

But let me tell you where it comes from. The thing that was—by the time we quit on the initial detector, we knew that of all the sources that would be the most likely to be seen would be something called compact binary coalescences. In other words, these are these nugget little objects that are almost incompressible in the universe, two of them going around each other. Now, two black holes are a good example of that. We didn't know how many black holes there were. That was the reason. So, we made a much bigger fuss about neutron stars. These are stars that are the weight of the sun, and the density of nuclear matter. For example, a neutron star is about the size of the city of Washington, D.C. in diameter, a tiny object, but has an enormous density, because it has the weight of the sun. So, it has a density—if you put a spoon into the thing, you couldn't lift your spoon. Because one spoon worth would weigh about ten to the fifteen grams, okay?

Zierler:

That’s pretty dense.

Weiss:

(Laughter) You couldn't move it, okay? It’s unbelievable stuff. And of course, you can’t put a spoon into a black hole, because you can’t get it back (laughter). And it would swallow you up along with it. That’s the extreme case. But that’s a geometric problem. The black hole is a thing unto itself. It’s a piece of geometry. On the other hand, the neutron star is still part of astrophysics. So, it turns out that the guess was that what we were going to be seeing is two neutron stars colliding with each other, and there had already been evidence of that. Astronomers had seen such a situation, and they had won the Nobel Prize for it, back in about 1993. I think it was ’93. Let me see. Boy, how can I forget his name? I'll get it in a second. Hulse and Taylor. A graduate student, Hulse, who—it’s an interesting story. He was a graduate student at the University of Massachusetts in Amherst here. And his professor who is Richard Taylor. Yeah, is it Richard Taylor, or—? It’s Taylor. I think it’s Richard Taylor. There are two Taylors in my life. Could be John Taylor or Richard Taylor. I don’t remember. But this Taylor is at Princeton.

Zierler:

Joe Taylor.

Weiss:

Joe Taylor. You got it. Thank you. Joe Taylor. Absolutely. Yeah, Richard Taylor is a high-energy physicist who tried to bring LIGO to Stanford. That’s a long story. Leave it go. It’s not an important part. Never happened. Yeah. Joe Taylor. They had discovered a system which consisted of—by using radio astronomy when they were still at the University of Massachusetts, where they saw a pulsar, which is a neutron star that’s spinning. And you hear the pulses because every time the star goes around, there’s a hot spot on the star—that’s a plasma—and it shoots out photons of all sizes, at the Earth or wherever it’s pointing.

And you can get the rate of the pulsar, how quickly it’s turning, by just listening on a radio—in a radio antenna how often you hear a pulse per second. And these things spin from either—oh, the slow ones go once every second, and then the very fastest ones go about 600 times a second. Imagine that. Higher than middle C on the piano. Higher than the C above middle C. That’s 500 hertz, okay? They're really booking along. So, it turns out they run into a pulsar which had a thirty-pulses-a-second period. Hold on. Is that right? No, no. That was the Crab Pulsar. It was about ten per second. And that’s how they kept track of it. They kept hearing—but they noticed something interesting about it. They noticed that instead of being a steady ten times per second, it would sometimes go a little faster, and sometimes go a little slower. And they saw that the change in frequency itself had a frequency, and that showed them that it was going around another object.

So, you had a pulsar that you could measure its motion by looking at the frequency of its spin, as detected by the people on the ground. Very beautiful system. And this happened to be in our galaxy. And they watched it for something like fifteen years. A long time. But they began to realize that the—they could solve for all the dynamics of this problem, because they had all this wonderful information about how the velocity of the pulsar changed from the Doppler shift of the pulsar. And so they made a dynamic model of that and found out that the only way they could explain that it was speeding up, these two stars, whatever—they couldn't see anything, but they only could hear the radio signal—but that the two stars were going around each other, and the period—that period of about—it took them about eight hours to go around each other. There several times involved the pulsar itself they were tracking had a spin frequency of ten Hz or spin period of a tenth of a second. That system is going around another one you can’t see, and you don’t know anything about its pulses, but once every eight hours, they come back to themselves. And they were tracking that eight-hour period and the deviations in that period, and they noticed that the period was getting shorter and shorter, and that was most unusual. And they then attributed that to gravitational radiation. And they were the very first people to measure gravitational waves. So, we had one example of a pulsar system, two pulsars, two neutron stars, going around each other. And that then became the source model. The thing we didn't know is how many of these existed in our own galaxy.

Zierler:

When you say source model, Rai, what does that mean?

Weiss:

That means—you asked, how do we know how to improve the detector? So, that’s something you know right away if you can do the theory for it. These two things would be going—they would go around each other very slowly, at eight hours, and stay there for centuries, maybe even millennia. But as they lost energy, they’d come closer and closer and closer, and they go faster and faster. And when they get to very high velocities, short periods, the rate of change of the period is very fast, because it’s all due to letting gravitational waves go. In other words, gravitational waves are making the system lose energy—makes these things fall toward each other, and they go faster and faster and faster around each other. Eventually, they collide. And we knew that we could detect the last—well, in our frequency band, from let’s say 40 hertz up to a kilohertz if it was close enough. If they got that close and they were going that fast around each other, we could detect that signal. If the two stars were close enough to us.

And so consequently we knew—we had a frequency band which we knew we were sensitive to. We knew that there were stars that would do that, namely these neutron stars. And if they come in pairs, eventually they would make a signal that would go through our sensitive band. And the only thing we really didn't know is how many stars like that there were within our sensitivity distance be close enough to us to make a signal larger than our detector noise that would do this in a year. That, we had no idea about. And that means how many binary neutron star pairs are there in a certain volume of space. That was the question. Okay, and that’s what we planned—we planned everything around that. In fact, all our distances—and we say that—we talk about what’s the distance to which we could measure such a system—we talk about—when we may chart how sensitive we are, we can see one of those out to a distance of one hundred mega parsecs about 300 million light years.

Zierler:

And who is “we,” Rai, at this point?

Weiss:

That means us. No, LIGO. LIGO.

Zierler:

But is the circle, is it narrowing or broadening at this point? Who’s the “we”? Who are the cast of characters that are—?

Weiss:

The cast of characters that are doing this are all the people working in the collaboration.

Zierler:

Okay, so this is a question that everybody is discussing, you're saying?

Weiss:

Oh, god, yeah! And a lot of people in the LIGO Collaboration did calculations. And there was one very important scientist named Vicky Kalogera who spent a lot of time trying to guess how many neutron star binaries were there within the range that we could detect. And she kept saying, “Well, you ought to see one a year. You ought to see one a year.” Well, eventually we did, but that’s not the first thing we saw (laughter). Turns out after we made these improvements—and that was a surprise—when we made all these improvements—this was sort of at the end of the engineering runs, of making the improvements of advanced LIGO—there are cute stories all along the way, but for example I'll tell you one on myself, since this is an interview about me, in a way.

I was sent by the guys running the advanced LIGO program, one of my students in fact—Peter Fritschel, very important guy in this whole story—he was a student back in the late eighties, early nineties, and stuck with LIGO, and he was in charge, effectively of installing and designing advanced LIGO. And another student of mine named David Shoemaker. These are names that you will find if you look up. These two names—Peter Fritschel, David Shoemaker—were actually the principals in—the sort of top of the ladder getting advanced LIGO going. It’s not that they did everything; it’s just that they were in charge of organizing it and deciding and making decisions on it.

So, what happens is that sort of just before everybody thought everything was ready, you have an engineering run. Just to see, does everything work? And the people at Livingston told Peter that they were worried about some RF interference—radio frequency interference. And so, Peter sent me down to Livingston to find out what was going on down there. And since my expertise is in electronics and—well, the thing I do best is electronics. That’s my sort of—if you want to call—what is the—skill I have, it’s electronics, okay? So, I went down there. I had made a diagnosis of it, and found out, to my absolute horror, that the place was singing away in radio frequency interference that in fact would interfere with the detection. And so, we sent somebody to—while I was still at Livingston, sent somebody to Hanford, and the situation was just as bad there. So, this was—I'll tell you what it was. It was about two weeks before the discovery. And so, I told Peter and Barry, at the time, that—not Barry. Who was head then? No, David Reitze was head. We had gone through other heads by that time. Do you want to know all the heads of LIGO?

Zierler:

Yeah, sure.

Weiss:

Okay, well, it was Barry Barish. That was followed by—oh, god. Mm! Well, the one I—there was one in between. The last director, the current director, is David Reitze. Barry Barish, and then there’s a guy between the two, who wasn’t there long, but I'll think of his name in a minute. Bothers me I can’t get that—I can’t dig it up. He made some important decisions, but not—yeah. God, that bothers me that I can’t come up with it. But, let it go. There were the three heads of the project, and the one I'm missing is the middle one. Hmm! Okay. Not okay it was Jay Marx who made a critical decision to move the ½ length advanced LIGO detector to India to allow better localization of the sources (once we began detecting them).

Yeah, there was a discussion between Dave Reitze, Peter, and myself, about what to do. And we were about to run. In other words—I can give you the dates. They were sort of in the middle of December of 2015. What happened is that I went down there probably the first of September. Yeah. Made the discovery of the RF interference by the second or third. And now the big discussion is taking place on the fourth of September—“What should be done?” And by this time, a lot of people in the LIGO scientific collaboration had already bought airplane tickets and decided where they would be—because when the run started, they all had to be in different places. Some people—a lot of the LSC people would go to the sites to work at the sites when you start the run. So, there was a lot of already investment been made in people organizing their lives, expecting the run to start on the fifteenth or sixteenth.

And so, we decided that what would the RF interference cause? It would cause trouble for a search for a continuous gravitational wave, a source that was putting out continuous waves, not bursts. And the other thing it would disturb is—probably not terrible disturbance; there would be some disturbance—the thing that got disturbed more than anything else would be a search for a stochastic background of gravitational wave noise. And we decided that would probably not be one of the things we would first be sensitive to, and it would give us more time to fix things if we waited a little longer, but we were not going to stop the run. That sounds very convoluted, but there was some logic in it. The logic was there was already a lot of money spent on making that run gonna happen, and this didn't look like it would stop searching for a black hole pair or stop looking for a neutron star. So that’s why the decision was made. And what happened—so I said, “okay, good.” Because it would have taken something like three weeks to fix it, was my thought. And so, I went on vacation. To Maine! (laughter).

Zierler:

Probably couldn't have come soon enough, right?

Weiss:

Well, no, not that, but (laughter)-. And anyway, so on the fifteenth, the morning of the fifteenth of September, that’s when we saw the first big signal. We were not yet running. The run was supposed to start the next day. And there were still people futzing around with the apparatus, and it turned out that we had to carefully establish that nobody was futzing around with it at four in the morning, at either Livingston or in Hanford. And it turned out it was true. There was nobody messing with things. But had it been a few hours earlier, we would have been disturbed by things that we would have been doing to the apparatus. Because we were not yet running. This is a little ahead of the run. All right, so the story then gets—and that, you can read about in many places. I can send you a—I will send you something about that. Or does it have to come from me?

Zierler:

It’s your call. I mean, I would love to hear your perspective.

Weiss:

Well, I got it mostly from what I read. I can only tell you—I'll send you this. It’s a LIGO magazine. You'll get a kick out of it, and it describes pretty well all the activity that was going on that almost made it so we didn't detect this. But what happened to me was that—I was in Maine, and just like you are in the Poconos, there was a reasonable connection there, and I would go to the log. You know, the log. And the morning of the—I think it was a Monday morning, or it was Tuesday—Tuesday morning—yeah. Tuesday morning, I noticed two things that were very strange. Normally, Tuesdays would have been the time when people made the decision to repair things. It turns out Tuesday morning, all through the initial LIGO and as has continued on through advanced LIGO is the—so you don’t lose a lot of coincidence time, you are repairing the apparatus. If there’s something wrong, like, I don’t know, an electronics thing busted, or the liquid nitrogen got exhausted or something—some consumable—or people just need to go in and clean up something—you did that once a week on Tuesday morning. And I noticed that fix-it morning had been cancelled. I said, “That’s very strange.”

And within about half an hour after that, I got a phone call—no, I got an email from one of the guys, saying, “Go look at this particular website.” And sure as hell, I went there, and by god, there was this signal seen in both detectors. It was enormously strong. There was no problem in seeing it. You didn't have to do any complicated analysis. I mean, it stuck out like a sore thumb.

Zierler:

It was staring you right in the face.

Weiss:

It sure as hell was. And it looked like it was—it looked the same in both places. You couldn't have gotten—you couldn't ask for anything better.

Zierler:

Did you know, Rai, exactly what it was telling you, or it was just obvious that it was something?

Weiss:

No, I knew pretty well that if it was true—if it were true—it was a pair of black holes.

Zierler:

And what would explain if it was not true? What would that mean?

Weiss:

Well, that’s what I get to. I'll get to that. Because that’s what we spent a lot of time on. I mean, this you will not find in that thing I'm going to send you, okay? Because this is at high levels, trying to decide what to do, okay? So, what happened was that the very first thought all of us had was that it was a trial injection. In other words, all through initial LIGO, what we were doing is we were injecting signals of our own into the two interferometers simultaneously. And there were multiple reasons for doing that. One of them was to just test the instrument, but more importantly, also it was to test the data analysis software, and finally to test the people. Could they find this? And we did that all through the initial LIGO to assure ourselves that everybody was paying attention, and everything was working right.

So, the obvious thing was, of course “It’s a blind injection.” And that immediately—I mean, this became a cause celebre. We all got on the internet. And the very first thing is that Reitze—Reitze knew about it right away, because he was sort of the center point, and we found out right away that it was not—there was no chance that it was a blind injection. Because there was nobody ready to do it yet! The software hadn’t even been written yet for this particular thing. And there was nobody doing it! So that made it a little more mysterious and it took away the most easy hypothesis. And then it got more and more sinister. And the most sinister one took us almost six weeks to get rid of it, and that was, had we been hacked? You know what hacking is?

Zierler:

Yeah.

Weiss:

Oh, you do, okay.

Zierler:

This is a legitimate concern?

Weiss:

Well, it was considered a very legitimate concern.

Zierler:

Who would do such a thing?

Weiss:

Well, that’s the point. Either disgruntled employees who know enough about the system, if it’s done internally. Or are there enough people on the outside who know enough about the data system to get in? So, it turns out that we stopped everything. That’s why fix-it day was stopped. Nothing was allowed to be touched on that apparatus. Nobody was allowed to put their finger on it, because we had to go around on each site and investigate how the signals looked at all different locations in the instrument. Trace the signal through the apparatus. Was the signal only in last point of the apparatus the digital storage or was it also at the input to the digital system? But before that, it had been an analog system, trace that analog signal back, all the way back to the detector. Is it there already in the photodetector that detects the light from the interferometer? You look at every point from beginning to end of the data system. And that took a while. It turned out, yeah, it was in the detector, also.

Well, that didn't stop the hacking hypothesis, because how do you know these people are so—they could have been real smart. They could have gotten in there and put it on the detector. So, people opened up all the electronics at every place, to see were there circuits put in? And it got worse and worse. And there was a whole report written by a guy named Matt Evans, who had a whole team of people within the project at both sites and they looked at all the craziest hypotheses that you could think of. And the only thing you could come up with is that every day, the hackers got smarter every day. In other words, you could cut out every simple hypothesis. But how do you know it wasn’t some more complicated hypothesis, and the hackers knew how to do that?

Eventually, it was a thing you couldn't completely reject the idea that it was a hacker. But it looked about as simple as you could do—the simplest explanation that satisfied everything you knew was that nature had done it.

Zierler:

It was almost too perfect, was the concern.

Weiss:

That was the problem. Yeah. All right, so there’s a big group meeting, and we decide—this is all on the phone—and we decide that the detection committee, which is a subgroup of the LIGO Scientific Collaboration, should now, on its own, check all the hacking information, look at the signals with the people who had the software that could do it, again, and see, is there something they could find that was wrong with this? And if not, there was one new thing that had to be done. And what they recommended is that we wait, with something to—with what was a whole very clever idea—was there were multiple pieces of software that could have detected this.

And what the idea was—this was one of the things that was suggested fairly early by the detection committee—was that why don’t we analyze the data with all the different software schemes we have to look for wave forms like this, including the very best detections we have, which are ones where you do what you probably—well, where you have modeled what the source would be. I mean, once you know that the source is a black hole or looks like a black hole, you take and model on a computer, using Einstein’s field equations, which was called numerical relativity, you model all different masses of black holes and make wave forms for them. And that was done, had been already started, but was done now in enormous urgency. And that wave form was used as a template against the template that had been measured. And sure as hell, we found out that it was a perfect match. I mean, a one percent match, to a pair of black holes, one weighing thirty solar masses, and the other one weighing twenty-nine. And then looking at that and the data analysis of those templates, we saw that it was a signal-to-noise of twenty to one. These things had a signal-to-noise—they were twenty times bigger than the equivalent noise in those templates.

So, we had no choice. We had to publish it. At least we felt that way. So a group was pulled together to write the paper that was in the LSC. The man who was in charge of that group was Peter Fritschel, again. And Peter did a very good job, I thought. We were at that for about—well, let’s see, the discovery was made in September 15th of 2015. The paper came out February something—tenth maybe—of 2016. So yeah, it’s about five months. Something like that. And that five months—of that five months, I would say three months was involved with writing the paper and sending it through the collaboration, making sure everybody agreed on it.

Zierler:

Rai, can you describe how you're feeling at this time? All of this work, something that’s almost too good to be true. Are you apprehensive? Are you brimming with excitement? Do you realize how big this is?

Weiss:

Well, I can tell you. What happened is a (laughter) very cute story. I happened to be a singularity in this, but there’s another—what happened—remember, I'm on vacation when this all started. A good friend of mine who was also in the LIGO collaboration came to visit, in Maine, named Peter Saulson. A good person to interview if you ever—he was the second spokesperson of the LIGO collaboration and ran a group at Syracuse University. And he came with his wife—we used to go hike—we went kayaking, used to. That’s what we do up in Maine. And, well, I showed him what I saw on the screen. And of all things, it turned out that Rich Isaacson was going to come two days later, to Maine. That had been planned months in advance. And so what happens is that Peter and I look at this thing—we worry about the hacker, but it looks—we do a little calculations—it looks to us like a black hole. That it was not an injection. So, we decided that—we haven't yet solved the hacker problem, but we weren’t deeply worried—we were worried, but not that deeply. And along comes Rich Isaacson, the guy who is the man who pulled this all together at the NSF. And I asked Peter to show him this thing. (laughter). You know, it’s on the screen. And Isaacson looks at it, and he says, “It’s got to be crap” That’s what—pretty much. And I said, “Yeah, well, we don’t know what else it is right now, but it’s not an injection. It could be a hacker.” But anyway, we actually go to a restaurant and drink on it being a real possibility of being a gravitational wave.

And what happens is we begin to write. The thing that finally convinced me wasn’t any of that. What it was in the middle of writing the paper on the day after Christmas of 2015, there is another event. Not as strong as the first one, but sure as hell, it sticks out. And you have to use all the tricks, the data analysis tricks that have been developed, to really see it. It’s nowhere near as big. The signal-to-noise is—instead of twenty to one was sort of ten to one. So that makes a big difference. Maybe nine to one. It’s somewhere around there, nine or ten to one. But I felt after that that that’s all I needed. I needed another one. And since then, we've had probably 40 of them? (laughter) Maybe even more. But they were not what we expected to start with. They were black holes.

And the reason why—Kip had already predicted that years and years before. Because Kip said that the—see, the trouble with the neutron stars is they're very light. They weigh only about a solar mass to two solar masses. And the amount of gravitational amplitude of the waves varies as the mass. And he says, “You're going to find a lot heavier black holes.” And that’s what we did find.

Zierler:

What did Kip know when he said you're going to find heavier black holes? What did he know?

Weiss:

Well, he guessed at it. He guessed at it. Look, it was already known that black holes existed. The way you could separate out a black hole from a neutron star is that neutron star can’t be bigger than about two and a half solar masses. Why? Because then it falls into itself and becomes a black hole. And on top of that, the velocity of sound on the surface of the neutron star gets faster than the velocity of light if it gets a mass bigger than about two and a half.

There are many, many arguments that say that—right now, in fact—I'll tell you a tale out of school —we are looying at an event that challenges this a little bit. We have an event where we have a twenty solar mass object living together with a 2.6 solar mass object. And that’s pushing this whole black hole business. I mean, pushing the neutron star about as hard as you can. But it probably is a black hole. We don’t know what that small mass object is. It could be a black hole. It could be a neutron star. But it’s right at the edge of where a neutron star shouldn't exist anymore. So, you'll hear about that in the next couple of weeks. Somehow the boys in the astronomy part think it’s very exciting.

So, at any rate, Kip surmised—I'll give you a thing about that. Kip wrote a book—Black Holes, Wormholes, & Einstein’s Outrageous Legacy [sic]. It’s a popular book. He wrote it in 1980. And the very first chapter of the book, there is a spaceship, and he has named the pilot of the spaceship his daughter, and they're out in space. Suddenly they come upon a place where she sees two thirty-solar-mass black holes going around each other. And she says, “Hey, we better back off. We better go.” And they put on whatever they can—all the engines—and they get the hell outta there as fast as they can. And they watch from a goodly distance, and sure as hell, they see this whole thing happening. Okay? 1980. In a make-believe story, Okay? So that’s—Kip was way ahead of his time in this whole business with black holes. He believed in black holes. Charlie Misner at—

Zierler:

Maryland.

Weiss:

—Maryland believed in black holes. John Wheeler believed in black holes. And you've got about all the people in the world who believed in black holes. At my place, at MIT, they all said they were complete fiction! Okay? (Laughter) So anyway, it’s sort of a wonderful story about the field, that became truth, and the very first thing we saw is sort of a kick in the ass for all the people who were skeptics (laughter). So, what came after that? The next thing that happened was Virgo came on the air. This is now—2015 is this thing, the big event. The second—by that time, by the time Virgo came on the air, we had been seeing a black hole every month. A new black hole pair. So, we had seen about ten—something like ten per year. But Virgo comes on the air, and that is August of 2017. That’s almost two years later. And they happened to be on the air just as another black hole is seen. And all three detectors see the black holes. That was very important, because now it would allow you to determine where the black hole is in the sky a lot better. See, with two black holes, the only thing we have is the information of—the very first one we saw, we saw it in Louisiana first. In about seven thousandth of a second, it hit—it went through the Earth and hit Hanford. So, we know it’s in the southern sky. That’s all we know. And you know a little better than that. You can do a little arithmetic on that .What you get is a great big circle in the sky, and you can say—we can place this source on a circle in the sky. One circle. As soon as you have another detector, a third detector, you have another circle on the sky you can draw. And now you have two great circles, and they can intersect at two points. Can you visualize that?

Zierler:

Sure.

Weiss:

So, you have then much better position information, where this source is. Well, okay, that immediately—as soon as Virgo saw this, so we had three detections, a lot of astronomers immediately go and try to find—because they now know approximately where it is, they try to find this black hole electromagnetically, meaning by methods that ordinary astronomers use. Well, they didn't see it. And a lot of theorists think that probably there isn’t an electromagnetic display when two black holes collide because the black holes eat up everything. They eat each other and then they eat everything around them. I don’t know if that’s true. I hope it isn’t. But that’s current thinking. So, people did not expect what’s called an electromagnetic counterpart to that triple coincident of the black hole.

But then two days later, one of the most exciting things in the whole field happened, which is that (laughter)- it’s funny. I mean, it sort of got us all by surprise. I think it’s the fifteenth or sixteenth of August. No, it was the fourteenth of August for the triple, I think, and it’s the seventeenth of August for this thing I'm about to tell you. I mean, it’s too close. It’s just amazing how close it is. And what was seen was by LIGO—was seen—a very much different kind of chirp signal, which starts at very low frequencies, it goes through our whole frequency range, and stops at about 400 hertz. So, it goes from the lowest note on the—well, about the lowest note on the piano, up to two above middle C. Just run a glissando with your hand up along that thing, and you will have this signal, okay? And that’s what we saw in this data. It was unbelievable.

And, okay, well that was—I mean, we knew what that was, because we had been—that was the first signal we were ever hoping to see. Those were two neutron stars that had collided with each other. But that wasn’t the end of it. It turns out that same collision was seen by a gamma ray telescope, in space. And just at where we would have had the end point of the chirp, which is at the high-end frequency, when they crash into each other like that—that’s when they saw a gamma ray burst, about a second and a half later than what we think should have been when they collided. And it turned out the uncertainty in the position of the gamma ray telescope was such that we couldn't use it to determine where the source was.

But here’s the funny story. Virgo did not see it. They didn't see it. They should have seen it, because the strength of the signal is big enough. They should have seen it. So, the only conclusion we came to is it must have been in one of the nulls—zero points—of the antenna pattern of the gravitational wave detector. There are null points, certain places in the sky which you cannot observe from, and it had to be one of those. So using that information, using the position that we could get from the great circle that LIGO alone saw with the two detectors, using Virgo’s non seeing it when they should have seen it, and using finally the gamma ray uncertainty—that circle didn't help a lot, but it helped a little—we were able to pin the thing down to something where about of the order of twenty square degrees on the sky. That’s still very big, but a lot of astronomers were willing to take the gamble to set their telescopes and look for that. And sure as hell, almost immediately, they found it. So, the optical telescope found it. Within the next day, infrared telescopes, UV telescopes, optical—a thousand telescopes had found this object. And on top of that, within about another week after that, x-ray telescopes and radio telescopes found this thing.

And what it was a thing which is called a kilonova. Two neutron stars come together, they smash into each other, they're heading to become a black hole—that’s what they're heading to. But before they become that, they spray all these neutrons and heavy nuclei all around and make a hell of a mess. But the little kernels of these two things eventually make a black hole. And it turned out we make a lot of discoveries when that happened. These people, not that we did that—the people with their ordinary telescopes—followed the light curve that came from that, and they did some modeling, and they found that that the nuclear decays that were going on there showed them that all the heavy elements, everything that was being made there, were heavy elements, elements larger than iron, more heavy than iron. So gold, platinum, tungsten, a lot of things we cherish, are made in neutron star collisions. And it solved the problem that people hadn’t—they guessed at this, but never really been sure of it—that most of the heavy elements in the universe are made in collisions of neutron stars. So, we solved an interesting problem. All of science got something out of this.

And the other thing we learned—we learned that gravitational waves travel at exactly the same speed as electromagnetic ones. Why? This source was a hundred and—yeah, 160 million light years away from us. A hundred and sixty million. And forget about the fact that there was a two-second delay between the gamma ray telescope and ours. That probably is due to the formation time. But if you just leave that two seconds in, and here are these two waves, electromagnetic wave and the gravitation wave, and they’re leaving this source, spending the same time together for 160 million years, and getting to the Earth within two seconds of each other, tells you that the gravitational wave is traveling at the same speed as the electromagnetic wave to a part in ten to the minus fifteen, enormous precision . So that one observation led to all this science. And that’s the last thing big that happened (laughter). I mean, of that kind.

Zierler:

So, Rai, when you say that all kinds of science benefited from this, can you explain a little bit in more detail what you mean by that?

Weiss:

Oh yeah. That’s easy. In fact, it has gotten a name now. It’s called multi messenger astronomy.

Zierler:

(Laughter) That’s great.

Weiss:

(Laughter) And so what it is this. I mean, this has always been the history of astrophysics is that you might detect something with a radio telescope, okay? I mean, with your eye, if you look out at the universe, nothing much is happening. Nothing! Your eye doesn't tell you much. I mean, even with a telescope, everything looked pretty placid. That’s all going to hell in the universe if you look at x-rays and you look at radio. And in fact, there’s so much going on that you can see the formation of the black hole. The black hole as it gets formed from the two neutron stars, you can see that first as a jet that forms of the plasma that forms—you can see that in the radio telescope. You can also see the emission by these high-energy electrons in the x-rays. So we know the thing turned into a black hole.

People had guessed at that for years, also. So people who had been worrying about how do two neutron stars collide, what is the explanation of jets, a whole bunch of things that others had been thinking about, they had got another piece of evidence for. And the nuclear physicists got stuff for seeing that what’s called the r-process, the process where a heavy nucleus keeps attaching more neutrons to it, more and more neutrons get attached, and they make the heavier and heavier elements. That had been guessed at for years, and lots of calculations had been done on that. We now know that’s for a fact. And people had done that already in accelerators. They had taken and sent heavy nuclei against other heavy nuclei and see them coalesce. That has been done. But that nature does it by making neutron stars was not known. Had been guessed at. People had calculated that.

So what I mean by many pieces of physics, there’s all these different things on that has nothing to do with gravitational waves. It has to do with nuclear physics, it has to do with x-rays, it has to do with formation of jets by black holes. That may be a little bit of gravitational physics. All of that stuff got information. I mean, a lot of people—it’s not just all crazies working in the gravity wave business. That’s what I mean, that more and more science got affected by that. And that’s something now that has made it so that the astronomers have finally embraced us. They think we're—you know, they're just waiting for us to tell them, where’s the next event? And we do that. The only trouble right now is that black holes don’t do a lot of that. This happened to be a very lucky coincidence—a neutron star pair that close that you could see so much of it happens probably once a year, just as Vicky Kalogera calculated. That was very close to us. See, 160 million light years, it sounds like far, but all the black holes we're seeing are out at a billion—one billion—light years. So it’s almost at the edge of the visible universe.

And what does the future look like? Well, the future is bright as hell, because now we have a whole tool that we can use to look at what’s happening way, way—I mean, the beginning—we don’t even know where these black holes come from. They could be primordial. They could be primeval. Because as I told you earlier, you don’t have to have a star to make a black hole. Although you could; that’s one way to do it. But you can get it from some gumrumption that happened in the early universe. I mean, geometry had so much energy in it. The distortion of space and time that’s associated with making these gravitational waves has an unbelievable amount of energy in it.

I'll give you a number so that you get an idea, then I'll quit for the day. It’s a number that’s just worth contemplating. Suppose—we'll go back to Kip’s little story, when he wrote that science fiction thing with his daughter running the spacecraft, okay? Suppose you actually had put the very first thing we had ever detected next to the sun, right where the sun is, and we had watched it from the earth. Well, not a hell of a lot would have happened. In other words, the stretching that you would have experienced, the stretching of you—you as a body being stretched—would be about one micron. That’s one millionth of a meter. You would never have felt it. You might have heard it in your ear, as a little thump. I mean, you could have, because it’s a little—one micron is a little bit bigger than the thermal noise in your ear. But the amount of energy that went through you is unbelievable. And the number—when you look at the sun and you go get a tan, the amount of watts per meter—that’s a good way of talking about it—the sun puts out one kilowatt per meter squared. Did you know that, or no?

Zierler:

No.

Weiss:

That’s an important piece of information, especially with global warming and everything else. Keep that in mind. It’s important to know. An important number. Kilowatt per meter squared. And by the way, you, as an individual, use up one hundred watts, just to keep your body going. So you could get nourished by the sun just to make you hot. But leave that go; that’s not important. But the amount of gravitational wave energy that went through you is ten to the twenty-five. Ten to the power of twenty-five, okay? Not ten to the power of three like it is for a kilowatt, but ten to the twenty-five. Unbelievable. And that tells you something which is really uncanny—that gravitational waves are the strangest thing that ever happened to man. Because it tells you that space—you know, here’s this distortion of space that’s made by the collision of these two stars—space is stiffer than steel. Space itself. You can’t compress space, easily, and in fact, it’s stiffer than steel by a factor of about ten to the twenty.

Zierler:

So a lot stiffer than steel (laughter).

Weiss:

One hell of a lot stiffer than steel. And by the way, that’s a frequency-dependent statement. It gets stiffer and stiffer the higher the frequency. But that’s at about one hundred hertz. So it turns out that space is a very strange object. And Einstein was absolutely right when he wrote in 1916—when he did his calculations about gravitational waves—that yeah, even though they carried all this energy, they would never be able to be measured. He said that himself. Because he thought that—you know, when he looked and made the numbers that now you know a little about—and he said that it’s never going to play a role in physics. It’s just too small.

And, well, now we know it does, but space is also enormously—it’s tough to distort. So you need something like a neutron star collision or a black hole collision to have any kind of evidence. And that’s something that all the astronomers who were skeptical of this thing, probably many of them knew that. I know Eddington must have known that. I don’t know; maybe even Bahcall knew that, and he figured “Well, that’s never going to happen.” And I don’t think I would have bet the way Bahcall would have, because I think that—my answer to Bahcall—and I have talked to him; he’s dead now unfortunately—is that—but other people who were very skeptical, I've been on committees—my answer to them always is, “If some crazy man comes along or crazy woman comes along with a complete nutty idea, but it’s not totally nutty—so you worry. Hmm, that sounds interesting. But it’s nutty! It’s off the mass shell. It’s something you don’t think of. And so the thing is that instead of trying to kill it, what you should do is come up with an idea of your own.”

I don’t know; that’s not a universal answer, but that’s a better answer than trying to kill something. In other words, the best defense you have, if you're worried about your field going under—come and think of something very interesting it can do. And I've told that to endless committees where somebody says that this—we don’t have the money for that, and we don’t think it’s going to work anyway. Now, if it does work, would it be interesting? Oh, yeah. So do something interesting with what you know. Look, it doesn't cover every case, but—yeah.

Zierler:

But this exemplifies it.

Weiss:

Yeah. So I don’t know if we need another one or not. What’s your opinion?

Zierler:

Rai, I think that I want to ask just a few sort of broadly retrospective questions, and I think we can wrap this up. I think we can do this.

Weiss:

Okay, so you want one more session?

Zierler:

No, no. I think we can do this now. Do you want to take an intermission? Should we take a five-minute break? Do you want to freshen up?

Weiss:

No, unless you want one. You want to go pee or something?

Zierler:

No, I'm good.

Weiss:

All right, let’s do it. Let’s get it over with.

Zierler:

(Laughter) I don’t want to get it over with. I'm going to miss you! Next Sunday, I'm going to go into Rai withdrawal syndrome. But I think we can do this now. So the first thing—it’s such an obvious question, but as you have put together this incredible narrative, what are some of the things you've learned about how science works as a community, that this narrative shows you? What has been most effective, as you worked to keep this project alive over all of the years? How does science work as a community and what are the lessons that can be drawn from this phenomenally successful story that didn't have to be, at any number of junctures, as you've told it?

Weiss:

Well, you're going to be very disappointed in my answer. I'm going to forewarn you of that right away.

Zierler:

(Laughter) Okay.

Weiss:

Look, I've been asked that in various ways before, as you can—and I'm sorry about this. This is very pedestrian when it gets down to—but what it is is this. I'm trying to say it so it doesn't sound totally trite. But the thing that drives you as an individual I think is the thing that’s most important in this. And it turns out that what drives me, and why I probably wouldn't have done anything else in any other way, is if it was fun to do, if it was really fun to do—and I mean fun, a pleasure to do it, you would do that rather than eat, or you'd rather have that—you have to have something like that. The thing had to be fun on every level.

Now, but what made it fun was I happen to enjoy tinkering with things. I get a tremendous kick out of—if I design a device or I design a circuit or I design something, and I go to the shop and build it, and it comes alive because it has now been built, and it does what I thought it was going to do, that is enough to keep me happy for many, many days. I hate to be so absolutely blunt about it (laughter). It’s very childish. But that is the thing that drives it.

And now, the thing is if it happens also that you can convince other people to work on something and they get fun out of it, that makes it even better. Because then you have something to talk about. You can go out and have a beer and—“We just made—oh my god, you got the detector working three times better? Let’s go get a drink on it!” And then you get a tremendous charge out of—it’s something that has been done by you and people that like you, and you know them. And it’s fun and accomplishment.

Zierler:

And not only that, but it probably—this fun probably is what has sustained you through all of the frustrations.

Weiss:

Absolutely.

Zierler:

The interpersonal frustrations, the bureaucratic frustrations, the political frustrations.

Weiss:

Yeah, yeah. It’s the only answer I can give you. And I hated the politics. And that’s why as soon as it was no longer necessary, I got out of it as fast as possible. Look, I started working on the beam tube, which was a great pleasure. I've got to tell you that. Even though most people think, “My god, what’s a scientist doing working on an engineering project?” it was a just wonderful thing. I worked with people I loved, and we get something done that was really great. I don’t have a better answer for you. I'm sorry. That’s it.

Zierler:

When did the Nobel rumblings become so loud that you couldn't ignore them? Because I feel like I've gotten to know you pretty well at this point, and I would guess that your first reaction to the Nobel rumblings was probably dismissive.

Weiss:

It was more than dismissive. I was resistant of it. I mean, I'll tell you, I was involved with—the trouble with the Nobel Prize is that I know a lot of people who aspire to it so that it consumes them.

Zierler:

It has hurt people, too.

Weiss:

Yes! It hurts a lot of people. Because it’s an anachronism. It’s a crazy idea. It’s an idea that might have worked when it was invented, but it no longer works. Because you can’t—and this is something I've said to everybody who will listen to me. In fact, I even said it to the Nobel people. I mean, “Can you change the way it was written?” They claim they can’t. Now, you take some other of these big awards, like the Breakthrough Prize—I don’t know if you know what that is?

Zierler:

Sure. Yeah, yeah.

Weiss:

Yuri Milner, I had a long talk with him when we got that, and his attitude is very much like mine. He says, “Look.” And we had a different attitude in one regard. He wanted to make scientists look like football players and movie stars. That’s what’s in his mind, is that in order to get respect for science in our society, and true in Russian society as well, they've got to be somebody people can adopt without having to have all this training. I understand it. And you play along with it.

But the thing is, his attitude has been—and he did the right thing—when we won the Breakthrough Prize, he gave every author of that first paper, all thousand-one-hundred-people got something—I don’t know if you know that—from the Breakthrough Prize. They did the same thing, with doing the picture of the black hole in M87. That was done fairly recently. This synthesis of a picture that was put together of what the region around the black hole in M87, which is a galaxy—so it’s a one-billion-mass black hole in a galaxy that’s about a thousand light years away. Maybe ten thousand. No, no, no, no, no. A million light years away. And that was in every newspaper in the country. I don’t know, did you see it? Maybe you did.

Zierler:

Sure. I remember it. Sure.

Weiss:

They wanted to give a prize—and what they did, the same thing there. There was, I don’t know, 500 people involved in that. Now, the thing is, that is the only way it makes any sense. The Nobel Prize has this problem that it—physics is not—science is not done, even by theorists anymore, as an individual enterprise. And so, what I had to do to accept it in my head—look, you don’t turn down a Nobel Prize. That would be an insult, and it would make you look like a jackass. But it was certainly on my mind. And the only way I could live with it is if I could take the Nobel money, and I tried the best I could before I had it, and ran one hell of a party in Stockholm, for everybody who I knew who was involved with this project. And we brought something like 200 people to Stockholm, to all share in the celebration of it. It’s not the same as winning it, but it’s the best thing that I could think of.

Zierler:

Rai, maybe to quantify this in very rough terms, if you can ascribe a portion of the effort that you and Barry and Kip had in the overall collaboration, what would that number be out of a hundred?

Weiss:

Wait a minute; that one, I don’t understand. You mean compared to somebody else?

Zierler:

No. I'm saying, if you think about all of the people that were involved that made this collaboration successful and that was recognized by the Nobel committee—the Nobel committee had to pick three people. How much credit, as a matter of percentage—?

Weiss:

(Laughter) Oh, god.

Zierler:

I mean, you don’t have to give me an exact number. It’s not an exact question. But just to give an idea of how strongly you feel about what you and Kip and Barry represent as a proportion of the entire labor effort, of all of the people that went into making this happen, what’s that percentage?

Weiss:

I'm not going to answer you on that one, because it’s hard. [pause] I mean, I can answer it piecewise. That’s the only way I can do it.

Zierler:

Good.

Weiss:

If you take the fraction of everybody’s life, of all the people including your own, that has gone into this thing, and maybe just in terms of time—but in terms of the fraction of their lives, my guess is that, of my life [pause]—well, I've done other things, but most of it went—it has been fifty years of my life, and it got more and more as the life went on. I think there are at least twenty people who have put that fraction of their life into this, also. At least. If not more. That’s the best way I can answer you.

Zierler:

That’s very powerful. That’s a very powerful way to understand what your role is in all of this, relative to all of the other people. And I'm sure it’s extraordinary generous. If I track down those twenty people, I'm sure to a letter, all of them would disagree. But still, it’s important –

Weiss:

Well, I'm not sure that’s true. But the thing is that—the way you have to think about it is another way. What it does do, if you're still alive by the time you get done with it all, it gives you a platform to push science. And boy, do we badly need that right now. In other words, I think one of the most—we are living right now through probably the worst thing of my life. I lived through Hitler a little, but I was too young to know what was going on. But what we're going through right now as a country is unbelievable. And it’s on Barry and Kip and everybody who has a public—myself as well—has a public—

Zierler:

Platform. A megaphone.

Weiss:

Yeah, a little bit of extra—that’s right, a little bit of extra, to make sure that you fight this off. And I'm trying my best to do that. I don’t know how long this will last, but I mean, the idea that irrationality has become a mode of operation is—I mean, just look what’s going on right now. Right in this whole pandemic, what we're doing in this country is irrational (laughter).

Zierler:

Hopefully these things occur as pendulum shifts, so maybe things have to get bad before they get good.

Weiss:

Yeah, I realize—I mean, look, I'm just overwhelmed by it right now.

Zierler:

Rai, for my last question—I kind of feel sad asking this, because I kind of don’t ever want these conversations to end, but they have to at some point—what do you see as the long-term implications of this research? In other words, there are immediate things that it answers in terms of how the universe works, but if we look ahead, using our powers of extrapolation, what do you think long-term are the new questions, the new experiments, the new theories, that will develop with the next generation of physicists who are looking at what you've devoted your life to? What can you imagine, looking to the future, your research helping to uncover long term?

Weiss:

That’s a hard question, but I'll try to answer it. But you have to know that this is ephemeral, also. The question you're asking, the thing I'm going to answer may not work. In twenty years, it may not mean a thing. That’s the trouble. I mean, I can tell you exactly what we're pushing right now. It has a lot to do with our future planning. Right now, we would like to take—and this is all of us, who are in this business—the Europeans, the Americans, even the Japanese—the next step for this field is to try to bring it to the level of sensitivity where we're able to look at the universe as a whole. In other words, to try to do something called cosmology. I think it’s going to continue for a very long time, this field. It’s going to become a branch, or already if not it’s a branch of astrophysics, it’s a branch of science. And there’s no reason that it should stop. I don’t see any reason why it should stop.

But one of the things—and you're asking me what are the things that lay ahead of us. There are some wonderful questions that now have occurred in our study of the universe as a whole. One of them is that we don’t know—well these are code words for lots of ignorance—but I don’t know, you probably know that we don’t really understand what the bulk of the mass of the universe is.

Zierler:

Of course.

Weiss:

That’s a huge unknown. And I would say there are, at this moment, at least 300 experiments going on in the world trying to figure out what that stuff is. Not only astronomers, but actually real experiments on the ground, trying to understand what that is. I think we're going to learn a little bit about that, even by this other way of looking at the universe. But that isn’t the most important one. The next most important thing is this discovery that was made—again, this cosmology that is so overpowering in this—is that there is a piece of gravity that pushes things apart. You're aware of that, I hope.

Zierler:

Mmhmm.

Weiss:

In other words, it was known by Einstein that he made it up because it solved a problem for him, namely he couldn't make a static universe out of his theory, so he had to have gravity push as well as pull. But then he found out that the universe is expanding, and so he said, “This was a terrible mistake.” But it turns out nature does have this. And it turns out that it’s going to make us all very lonely in, let’s say, ten billion years. Everybody is going to be very far away from everybody else. In other words, the universe is accelerating in its expansion. And why? God knows why. But it looks like gravity has a term that pushes, not just pulls. And that, again—why does that happen? And why does that couple—it’ll couple to the rest of the other ideas in physics. I'm giving you grand ideas, and not yet answerable, but I think that, again, this field will have something to say about that.

Zierler:

And so to stay on this topic of grand, do you see your research at some future point as contributing to a grand unified theory?

Weiss:

It might. It might. It depends what we find out from understanding the black holes better. That has been a direction in which they're unified, which is interesting. But the most interesting one is not in my realm, but in the realm of your children. And that is that if they get into science—is that they may very well get to the point where they would love to know, with every technique that you can think of, but gravity waves are probably the only way to do it, of what at time T equals zero. In other words, what happened at the moment when the universe got created. It’s to some people a very arrogant question to ask. But it’s a question that I think is within our capability to answer.

And I don’t want to go into all the crazy ideas that exist now, but there are many of them. But we could very well by this technique—not by the technique we have now, but we have to build this to be very much more sensitive and do it in space—that’s one way. There are ideas for how to do this. But they're not derivatives directly from what we're doing now. You have to effectively take LIGO and stick it into space. That’s what you have to do. And it’s not what LISA does. LISA is not going to do this either. But really put LIGO into space. Very much more difficult than LISA. You could get an answer to that. And the reason why is because gravity is so penetrating. In other words, with light, we can only look to the point where John told you. You know, you can look to that skin of the universe’s plasma, which is about fourteen billion years ago. That’s 300,000 years after the explosion. If you want to see what happened at the explosion, you're going to have to penetrate the densest stuff that ever was, and gravitational waves penetrate that. And that would be the final most beautiful—I mean, the most singular thing that gravitational waves could do for the world, is to find that out. It’s not in my cards. It’s not in my life. Maybe not in yours. But it’s in your kids’ life. So, yes, there’s a very, very grand thing we could answer (laughter). But it ain’t easy!

Zierler:

Well, that could almost be a motto of your entire career—“It ain’t easy.” Right? Rai, I want to say, there’s a lot of people in physics who are well regarded by their peers, and you're certainly one of them. But I just want to state on record—you know, this is what I do for a living. I talk to your peers. I talk to all kinds of people. And your peers love you. And I know why now.

Weiss:

Really? Oh my god.

Zierler:

I know why. You represent so much goodness in every sense of the word. And maybe if you refuse full credit from the Nobel committee because of this particular work, you can get some satisfaction in understanding just what a positive force you have been in the physics community for all of these decades. And it has been a pleasure and an honor getting to know you over the course of these Sundays that we've spent together. And I really just want to thank you for so generously spending this time with me. And just to state the obvious, there’s a lot of history that needs to be written about your work in particular, and about the issues that you raise about how science gets done, that are enormously important, and as you rightly point out, are more important than ever because of the situation that we currently find ourselves in. So, for a million and one reasons, I want to thank you for spending this time with me.

Weiss:

I'm very glad. Thank you for that, by the way. Thank you very much. On the other hand, if and when this crisis is over, and you travel with—I would like to meet those kids, okay?

Zierler:

You got it.

Weiss:

And sometime, you're going to start looking at colleges for them. Isn’t that true?

Zierler:

There you go.

Weiss:

And I would be tremendously pleased if I could meet some of them.

Zierler:

(Laughter) It’s a deal.

Weiss:

It’s a deal, okay?

Zierler:

All right, Rai. Thank you so much. We'll be in touch.

Weiss:

Okay, thanks a lot. Good.

Zierler:

All right. Take care.

Weiss:

All right. Bye.

Zierler:

Bye-bye.

Weiss:

Thank you. Yeah, thanks a lot.

[End of Recording]