Joe Redish

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
Joe Redish

Credit: Ginny Redish

Interviewed by
David Zierler
Interview date
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 Joe Redish by David Zierler on November 11, 2020,Niels Bohr Library & Archives, American Institute of Physics,College Park, MD USA,www.aip.org/history-programs/niels-bohr-library/oral-histories/XXXX

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

Abstract

In this interview, David Zierler, Oral Historian for AIP, interviews Edward F. (Joe) Redish, professor emeritus of physics at the University of Maryland. Redish reflects on the symbiotic nature of his interest in nuclear theory and physics education, and he describes his long collaboration in the latter field of Lillian McDermott. He recounts his childhood on Long Island and his developing interests in math and science. Redish describes his undergraduate education at Princeton where he was mentored by John Wheeler in studying unified electromagnetic fields from a pedagogical perspective. He discusses getting to know Charles Misner at Princeton, and he explains his decision to go to MIT for graduate school, where he conducted his thesis research under the direction of Felix Villars on nuclear reactions using quantum field theory. Redish explains the opportunities leading to his postdoctoral appointment at the Center for Theoretical Physics at Maryland and ultimately his ability to join the faculty and achieve tenure in the physics department there, in recognition of his work in three-body clustering problems. He describes the lengthy intellectual process of switching over entirely to physics education research in the early 1980s and why teaching at a large public university proved to be the ideal pedagogical proving ground for his interests. Redish discusses his entrée to the world of AAPT and what he saw as some of the orthodoxies in the field that were ripe for change, including making the field more student-centric. He describes his current project, NEXUS/Physics, which is an introductory physics for life sciences class that he developed in partnership with biologists, and he explains how this fits with his personal research interests that have delved recently into the biological realm. Redish explains the difficulty in mentoring physics education graduate students because of the expectation of their mastery of both physics and pedagogy, and at the end of the interview, he describes the Resources Framework that he is building as akin to a grand unified theory of physics education. 

Transcript

David Zierler:

Okay. This is David Zierler, oral historian for the American Institute of Physics. It is November 11th, 2020. I am so happy to be here with Professor Edward F. Redish. Joe, it's so good to see you. Thank you for joining me today.

Joe Redish:

Thank you for having me, David.

Zierler:

All right, as a first matter of business, readers might think already, why Joe? The name is Edward F. Redish, and yet you go by Joe. So let's start with that. Why do you go by Joe?

Redish:

Oh, it comes from my middle name.

Zierler:

Which is what?

Redish:

Frederick.

Zierler:

(laughs)

Redish:

Well, it couldn't come from Edward. That wouldn't make any sense.

Zierler:

(laughs) I see.

Redish:

Sorry, that's the best I've got. (both laugh)

Zierler:

Who started calling you Joe? Was it your parents?

Redish:

I have always been called Joe as long as I can remember.

Zierler:

And you've never delved into this further, for an actual explanation?

Redish:

No, I have not.

Zierler:

Okay, well--

Redish:

I have suspicions, but they're not appropriate for a conversation like this. (laughs)

Zierler:

Very good. Well, Joe, let's start. First tell me please your title and institutional affiliation.

Redish:

I am currently professor emeritus at the University of Maryland. I retired just this past summer, after 52 years at the university.

Zierler:

Well, congratulations for that.

Redish:

Yeah, I guess. I never expected to retire. And actually, I have only retired because I wasn't getting enough published¬ work done.

Zierler:

That's right. And that's what it means for physicists. Retiring means it's now time to actually get some work done.

Redish:

Yes, yes. I realized that, in the past five years I had not written a paper where I was the first draft author. I'd had a bunch of students working with me writing papers, but I hadn't written one myself. And that started to eat at me. So the past three months, I've written six papers. Small ones.

Zierler:

What's some of the stuff that you've been working on recently?

Redish:

I have been working on trying to get down for teachers what I have learned, especially from the past ten years of working with biology majors. I've been teaching physics to biology majors and doing research on the question: Why do they have so much trouble?

This has been closely tied to what my research in physics education has been all about. Namely, how do we think about physics, and what is the structure of our thinking? Actually, this ties in very well with the research I did in nuclear physics as well, and in my interest in physics topics from the first. But I really came to the point in my career, pretty early, thinking, what physics is about is not about: How does the world work? Physics is about: How can we think about how the world works? And what that does is put a critical piece of psychology into the practice of physics. That physics has to do not just with the real world, but with the interaction of the human brain with the real world.

Zierler:

There's also a philosophical dimension to this as well, where the mind is standing between us and objective reality.

Redish:

Of course it is. I've discussed this often with my dear friend, Royce Zia, who's a physicist from Virginia Tech. We go back to college and graduate school, and we've talked about these issues for many, many decades. This philosophical base comes down, for me, to three axioms.

First, there is a real world. That's the axiom I work with. That a real world exists independent of people. Some people don't take that axiom and want to work with other axioms, and that's fine.

The second axiom is, we don't live in that real world. We live in a virtual reality created by our brains from limited input data. That seems to me to be axiomatic, obvious.

But the third axiom is then, we're not getting the whole story; but we can get a better view, a better approximation to the real world, by doing the process of science, and working together communally using observations and theory-building in an interactive way.

That's my philosophical basis, and it led me in certain directions in my nuclear physics research. And in my physics education research, it's led me in some directions which I've found absolutely intellectually challenging and fascinating.

What I'm currently working on is the issue of: How do we use math in physics? That one of the problems you see when you start teaching, actually anybody, but particularly life science majors. They don't look at math the same way we do.

They view math as useful for calculating stuff, but the idea that you can use math as part of an integrated view of the world and use the math to think with is something they don't learn. They don't learn it in their calculus class; they don't learn it in their biology class; they don't learn it in their chemistry class.

We physicists have to teach it. We tend to take it for granted, and I have been analyzing and interviewing students, and interacting with them, and doing both research and teaching on this issue. It's been extremely interesting. I'm continually learning. You know, I've been teaching for 50 years, and I keep learning new stuff about the physics from my students. (laughs) It's really amazing. Which is why, even though I'm retired, I expect in a year or two, if my health holds up, to go back into the classroom.

There are so many issues that are involved in using math effectively in physics, or in any science, and the crucial element that I've identified as being missing is what I call The Blend. We blend our conceptual physical knowledge of the world with our symbolic knowledge of mathematics. And so we create quantities which have conceptual structures that are not the same as the mathematical structures that they look like.

This blend changes the way we handle the mathematics. I have a paper I wrote about five years ago, with a former student (Eric Kuo, now at University of Illinois), called "Language of Physics, Language of Math", that goes into this. We start out basically saying, it sounds like physics is a dialect of mathematics. But we make the claim, no, it's not. It's a different language. It's a deeper set of distinctions. I've been working on how we identify the components that we can teach our students to help them make the transition to see mathematics as a more useful way to parse their world.

Zierler:

Joe, I want to ask a very broad question at the outset to sort of fuse your two major research agendas. And that is, have you thought about how being a theorist influences the way you think about physics education? In other words, the counter-factual is, if you were an experimentalist, how might that color your approach to physics education?

Redish:

Absolutely. I'm one of those theorists who, you know, I went in the laboratory and the people I worked with in the laboratories liked my work in the laboratory because I was always thinking about how things worked. But I really didn't like the hands-on stuff. In fact, one of the reasons that I chose not to go to one graduate school that I had won a fellowship at is because I learned that if you didn't do well enough in the qualifier, they didn't let you be a theorist. You had to be an experimentalist. I think that's kind of unreasonable and unfair. I said, I'm not going to any place that will not let me be a theorist. (both laugh)

I'll tell you a story because I have a good answer for your question. When I finally made the decision not to try to work on both educational issues and nuclear physics issues at the same time, but to actually jump over the cliff and say, "I'm going to do physics education research full time," it was 1992. I was 50 years old, and I decided I needed something new.

I went to do a sabbatical with Lillian McDermott. Lillian just recently passed away. But at that time, she was the leading physical education researcher in the United States. She had a huge team, many, many people, a great operation. It was kind of like going to Fermilab (laughs) and learning the ropes. So I went there, and I started working with her group, and I was immensely impressed by what they were doing and what they had learned.

I started to do a lot of reading. I was reading a lot of psychology. I was reading sociolinguistics. I'd always been interested in those subjects and had talked a lot about it with my wife, Ginny, who is a PhD linguist. She introduced me to a lot of great stuff — Donald Norman, Deborah Tannen, John Bransford, and others. I started reading the semantics literature. I wound up starting to put together some coherent principles that I thought were underlying what we were doing; the theoretical foundations.

I came up with a set of basic principles about how people thought about things — physics in particular, and what a coherent view of the world meant. And I presented it to Lillian. She had a fit. She said, "This is terrible. Don't do this. Stay away from this." I said, "Lillian, you've been teaching me that in order to build new knowledge, you have to build your own knowledge. This is the way I build knowledge." She said, "Don't ever spread this around."

She had reasons, because she had started from Piagetian experiments in psychology that had been done with 11- and 12-year-olds. And she said, "How can we improve that and apply the same kind of thinking to college students?" Her papers were fantastic, and they made a big splash; but the people at her university said, "If you want to get tenure, you can't be doing psychology. And your papers look too much like psychology."

I think that burned her. She didn't have tenure when she started. What she did was very gutsy – starting physics education research in a physics department. On the other hand, I came in from a position of security. I had been tenured for 15 years. I had served as department chair. I had been on national committees for nuclear physics. And I already had a reputation.

 

Zierler:

But was the transition for you as stark as it was for Lillian, in terms of leaving a previous career behind? Or a previous research endeavor behind?

Redish:

It was. You know, one of the things was that as a nuclear theorist, in what was ranked by the Department of Energy as one of the top ten theory groups in the country, we were essentially guaranteed funding. We had funding every year. We just had to keep doing good stuff. Every five years they would come by to check us out. When I jumped over the cliff into physics education, nothing was guaranteed. The competition was incredibly fierce. I was a newbie and didn't have a track record. I did wind up getting my first grant with the first proposal that I wrote, and I didn't realize how unusual that was.

But let me finish the story. Because what then happened is, I wrote a paper describing what I was thinking. And I called it, "The Implications of Cognitive Studies for Teaching Physics." I didn't say "Cognitive Science" because Lillian had freaked me out about physicists not seeing psychology as a science. That was 30 years ago, right? That was a long time ago, and cognitive science now has come a very long way.

The first paragraph that I wrote in that paper was, "I had to write this paper. You're not going to publish this, but I have to write it anyway. And I figure since I wrote it, I might as well send it to you." Well, I decided I would try to send it to the American Journal of Physics, and I removed that first paragraph. It was immediately accepted and published a year later. It is one of my most-cited education papers. People still read it. It's very popular.

For the next ten years, Lillian and I became the spokespersons for the West Coast and East Coast of physics education. We both were popular speakers. I must have given 100 talks in the 90s and 2000s, and Lillian gave more. She would talk about the observations, and I would talk about: There are three legs to any scientific study. There's the observations or experiment, there's the practical applications or engineering, and there's the sense making or theory. And the most important leg of a three-legged stool is the one that's missing. Theory is our leg that's missing.

Lillian and I were continually ping-ponging back and forth. Lillian was saying, "We don't want you to talk about theory here." And I was saying, "You have to talk about theory." So yes, the theoretical perspective had a tremendous influence on the way I looked at physics education research – and on how I chose to think about what I was doing. I did work with a lot of people. I did make a lot of observations and experiments. But it was always in the service of building a theoretical understanding. So the fact that I was a theorist had a huge impact on the way I've done physics education.

Zierler:

Joe, another really broad--

Redish:

And Lillian was a nuclear physicist also.

Zierler:

I know, I know.

Redish:

But an experimentalist.

Zierler:

There's food for thought therein. Joe, another broad question. In what ways has your career at the University of Maryland been beneficial to your interests in physics education? In other words, as a large public research university, you have probably benefitted from an undergraduate student body, where on the upper end of things, students and their intellectual capacities are entirely competitive with places like MIT, Princeton, Caltech, but because it's a large public university, the range of aptitude is much broader than it would be at those kinds of institutions. So I wonder if you could answer generally how your tenure, your career at a large public university like Maryland, may have been beneficial for the way you've thought about physics education over the course of your career?

Redish:

Yeah. That's a good point, and I think that's right. As I go back to the beginning of my career, I see some of the hooks that started to pull me into physics education. When I had been teaching for a couple of years at the University of Maryland, I got invited to spend a year at Saclay, France, at the Centre d'Études Nucléaire, doing research on their dime. So Ginny, my two young children, and I picked up and went to France for a year. And from there, in the summer of 1974, we all went to England because there was a physicist at the University of Surrey doing theoretical research on issues I was interested in.

While I was in Surrey, I ran into Louis Elton, who had been a nuclear theorist the previous decade and had switched into physics education. Because he had worked on problems I was interested in in nuclear physics, I went and talked to him. He listened politely to what I was doing, and when we finished with that, he said, "And are you teaching? How's your teaching going?" I said to him, "Well, I've been enjoying it. I've been teaching physics majors, and I really seem to resonate very well with them, with the really good students who are there at the top, and they want to come in and talk to me. And that's been very exciting." He looked at me and he smiled and he said, "And they're the ones who don't need you." I kind of paused. It was like an arrow to the breast. I've never forgotten that phrase. "They're the ones who don't need you."

I remembered my own best class ever was the worst teacher ever, where I decided if he can't teach me physics, I'm going to do this myself. That's when I learned how to learn physics — as a second-year physics student. It was a fantastic experience. I didn't need him! But there were students there who did.

Being at the University of Maryland gave me the opportunity to work with a wide range of students, with a diverse population of students, and then to do some…, at least to do one big study, a couple of big studies, with a large number of students. My first two papers in physics education research were real experimental papers with 500 students at Maryland and 1000 students at other universities. And so forth.

Soon after those first big physics education research projects, David Hammer came to us, and taught me to use videos and to do qualitative research. And I said, "Oh this is where the theory is!" And I just totally switched over from big quantitative studies to qualitative in-depth research.

But you're right about having the diversity of students. It was of tremendous value to me. But it's also having... So let me see, how to say this. In the past ten years, when I got deeply involved with the course for life science students, I got to work intimately with people in other disciplines, especially biology. I worked with a lot of biologists, and that has been fantastic. Having the ability to draw on a large population of faculty as well as students at the large university has given me many opportunities that I really have appreciated. And it has influenced my career very much.

Zierler:

Joe, let's take it all the way back to the beginning. Let's start first with your parents. Tell me a little bit about them and where they're from.

Redish:

Okay. My grandparents were all born in Eastern Europe. Eastern European Jews. They came about the beginning of the twentieth century. My mother was actually born in Lithuania. She came when she was an infant. My father was born here, in New York. I grew up on Long Island. My father was a medical doctor who died when I was nine years old. We struggled along for a while. I managed to succeed well in high school and wound up getting into some prestigious universities. I went to Princeton and MIT, got my first job at the University of Maryland, and that was it. Turns out I didn't need any others. I was hired on a one-year renewable postdoc, and they just kept renewing me.

Zierler:

Joe, that's a very nice microhistory, but we're going to dial way back here. First let's start a little bit with the fact that your dad died when you were so young. Did you have any other father figures in your life after that?

Redish:

No. Well, a little bit. I had my mother's brother who served slightly as a father figure. But he wasn't around much. He was kind of the black sheep of the family. An artist, he was off in Paris for half of every year. And then traveling around the States. He had an influence on me, at least on my artistic tastes. But I didn't really have any male role models.

Zierler:

What grandparents did you have in your life?

Redish:

Well, my mother's father was a rabbi. He was very old when I was born. He was born in 1865. (laughs) And he did not speak English well, so I didn't know him well. But I did know that I had come on that side from a long scholarly tradition. He had married the rabbi's daughter, who had come from another string of rabbis, and so I was kind of the ninth generation, the first generation that didn't have a rabbi. And so there was a long scholarly tradition in my mother's family. I didn't know much about my father's parents. My father's father was a steamfitter. But he didn't interact with us much. My mother was the one who ran everything, and she made it happen.

Zierler:

Is Redish an Anglicized name?

Redish:

Yeah, I think so. I'm not sure what it originally was before they came.

Zierler:

Oh.

Redish:

I have not been deeply connected with previous generations and with the family tradition. I have cousins who are; but I rather look forward, rather than back.

Zierler:

Joe, when did you start to get interested in science? Was it early on?

Redish:

Early on. Early on. For some reason, I was interested in equations. And in my baby book, I found a place where I was, before I was writing, I was writing equations. I didn't know what they meant, but I was copying in the-- I don't know where I got that from, but I found that book. I must have been five. I learned to read about three and a half. My mother used to read me comic books, read with her finger. And I picked it up. I was actively reading by first grade, and they put me in the public school, where apparently I caused a lot of trouble. I had read all the books in the classroom by the first month and was bored. And the principle called my parents and said, "You've got to get him out of here. We don't want him."

My parents found a private school and sent me there. They skipped me up a grade. But I was interested in math and science the whole time. I was really taken by dinosaurs and astronomy early on, and when I was 11 and 12, I was reading George Gamow's The Universe and Dr. Einstein. I was fascinated by physics. As a high school student, I read Heisenberg's little book on nuclear physics. And that was interesting, because I thought I would be a doctor.

Zierler:

Like your father?

Redish:

Like my father. That's what you do, right? You're a doctor. Then I got to biology. I had general science the first year in high school, and I really liked it. I remember that one of the things that just absolutely turned me on was Archimedes' principle. Archimedes' principle says if you put something inside water, the upward buoyant force is equal to the weight of the displaced water. And I looked at that and I said, "What? How can that be? That water isn't there anymore. How does it know what to do?

The way I like to say it is, this was the first thing that I saw in school that seemed deep. Everything else was either obvious, like math, or was a little challenging and interesting, but still pretty obvious. Or it was just some fact, like the war took place in this year.

But Archimedes? Oh, how did that work? And then the fact that the math described the real world really did it for me. Now, I was interested in biology. I had a biology class, which was not very good. It was 1956, 57. There was none of the new stuff that was happening at the time. DNA was just new. We had all the phenomenology about structures of animals and we were dissecting worms and frogs. It was all experimental. There was no theory. So I really didn't like it.

But I was still interested in biology because when I was a junior in college, we had to do a junior term paper. Two actually. Two long term papers that you had to do over the whole semester. They were like 40 pages long. And you had to do a bunch of research and write a coherent paper.

I went to the library, and I said, "Is there anything on mathematical biology?" I went to the bookshelves and found, like, there were three books. And I pulled them out and I looked at them, and I said, "This is bogus." There was good mathematical biology going on at the time, in the journals. There were no books because it was pretty new. I said, "There's nothing here." I really didn't like it. It's interesting, because if I had been at Harvard instead of Princeton, at that time, E. O. Wilson was setting up to do mathematical ecology, and bringing undergraduates in. I might well have been a biologist if I had wound up at Harvard instead of Princeton. But as I went on, I stayed in physics.

Oddly enough, both my children wound up doing biology. My son, David, is a neuroscientist and my daughter, Deborah Redish Fripp, has a PhD in marine biology.

But I'm back doing biology now, right? I'm working with my botanist friend Todd Cooke on describing the physics of the forces and energies that are responsible for bringing sap up 300 feet in trees. The biologists know how this works, but they don't explain it a way that seems clear to a physicist. They have a lot of words to describe it and they know what's going on. But as I go in and build equations to disentangle it, I say, "Oh this is what's happening." I'm really enjoying it. It's absolutely fascinating. I really like biology, and I might have been a biologist. You know, I've seen lots of opportunities to do very cool work in biology. It's such a great field. I have a lot of respect for it. There's a lot of interesting stuff. (laughs)

Zierler:

Joe, what larger events might have captivated your interest in physics? Perhaps the advent of the nuclear weapons race, or the space race, when you were a kid?

Redish:

It's interesting that I was-- I'm a big science fiction reader. Loved science fiction from the time I was seven or eight or so. Reading Ray Bradbury and Asimov and Heinlein and Alfred Bester and all of those writers. I was excited about the space race, but that didn't turn me on much. I was excited about particle physics. I was excited about Einstein and the uncertainty principle and quantum mechanics. And then the weirdnesses that had been discovered in the first half of the 20th century. That the world wasn't the way it looked. That there were things that didn't match with your experience. That just really excited me.

I read a lot about relativity, special and general. As an undergraduate, I thought I was going to be a general relativist. I did my senior dissertation at Princeton with John Wheeler. That was interesting, looking back on it, because Johnny gave me three potential topics to look at in 1962. One topic was detecting gravitational waves. Another was gravitational lensing. Both of these were pivotal research areas of the last half of the twentieth century. They both really worked out.

But the third was a pedagogical project. It was: We have electric lines of force, and we have magnetic lines of force. But what Einstein teaches us is that electricity and magnetism are not separate when you view them in four dimensions, and when you look at the electric and magnetic fields in 4D, they are a unified electromagnetic field. Is there an analog of electric and magnetic field lines in four dimensions? Electromagnetic field surfaces?

I just looked at that and said, "Oh, I want to do that." So I picked the pedagogical project rather than the two Nobel Prize projects that he offered me. (both laugh) I think that says something about my excitement and interest in physics education. You know, the fact that when physics education research started to happen, I was a fish ready to take that bait.

Zierler:

Right. There was something inherent in your outlook that suggested this is where you were headed.

Redish:

It's interesting, because in high school while I was reading Heisenberg and Bartlett and Gamow and all of those about the physics, I was also trying to read semantics. I read Korzybski (Science and Sanity) and Ogden and Richards (The Meaning of Meaning). I didn't get much out of them. I kind of said, "Well, this isn't going to get me anywhere." But in my past 20 years, a lot of my work has been about semantics.

Recently I've been working on: How do we make meaning with mathematics? And what I decided is, what do I mean by making meaning? What does "making meaning" mean? If I don't know what making meaning with words means, how can I tell you what making meaning with mathematics means?

So I had to go back and read a lot of semantics literature. That really helped. And it helped me form theoretical concepts for physics education research. Blending for example, comes out of some cognitive linguistics (Fauconnier & Turner, The Way We Think). Epistemological framing and various other concepts that I've used and helped develop come out of literatures that physicists don't traditionally (Erving Goffman and Deborah Tannen). It goes back to this interest that I seem to have had from the beginning. I kept it going by marrying Ginny — a wife I could talk to about it.

Zierler:

Now, was your plan to pursue a degree in physics from the beginning at Princeton?

Redish:

Absolutely. Yes, absolutely. And in fact, it was a little scary, because I came in, and they gave me... I had been a hotshot, especially in physics and math, in high school. You know, high score on the SATs and the state Regents exam, and so forth. And they gave me a little test and put me into the honors section.

Most of the kids in that section, there were about 25 people, had had the equivalent of AP physics. I didn't have a clue as to what was going on. I was also in honors math. I wound up with a C+ and a B- in physics. And that was my chosen major. In honors calculus, I had A+ both semesters.

So that summer between my freshman and sophomore year, I did a lot of heavy thinking. What am I going to do with my life? And what I decided was, I didn't want to be a research mathematician. I loved mathematics, but I wasn't that excited about the fundamental issues in mathematics.

I decided to stay in physics rather than go into mathematics. I said, "Look, you got a 97 in the state regents exam. If you can't be a professional physicist, you can teach high school physics." My decision was to stay in physics because I would rather teach high school physics than be a professional mathematics researcher.

Zierler:

And was this more of a pure mathematics or applied mathematics that you were thinking about?

Redish:

Sure, oh, I took the honors class, the whole math major's sequence. We did topology, and we did real analysis, and we did, you know, all kinds of exciting stuff. It was not particularly practical. Though there were some things that were obscure that I looked at and said, "Non-measurable sets. Who cares about non-measurable sets? In physics, everything is smooth." And so I kind of, in the upper years in math, I got knocked down because I didn't close my proofs on the fancy sets and stuff. I said this isn't interesting.

Then of course, Mandlebrot and fractals came out, 15 years later. I said, "Oh my, that's what they're talking about. It's everywhere!" (laughs) And there were things about doing calculations, for example, simplexes, where you could break up surfaces, approximate them by little triangles. I said, "Who cared about that? That's not useful." But that's how you do every big calculation in complex realistic situations, solving partial differential equations.

We didn't have the computers when I was in college, so none of that looked relevant, and I was not interested in it in the abstract. I was interested in applications and how the math fit with the world and how the math gave you ways of thinking about the world. I really liked that from the beginning. So... I don't know if I answered your question. (laughs)

Zierler:

Where was general relativity when you were an undergraduate? Was it –it goes up and down in terms of when it was fashionable and when it was not. I'm curious what your sense of general relativity when you were an undergraduate?

Redish:

I was pretty clueless as an undergraduate. I mean, I was at Princeton and there was a lot of general relativity. Robert Dicke was there, just missing out on a Nobel prize. And I took general relativity from Dicke, and I did my thesis with Wheeler. I thought general relativity was one of the things you did in physics. In fact, it's amusing because I had had Charles Misner as a professor in junior year. And I wanted--

Zierler:

Was he visiting Princeton then? Or was he on the faculty?

Redish:

He was an assistant professor. I wanted to do my senior thesis with him. He said, "I'm sorry, I'm leaving next year. I'm going to the University of Maryland. But you should work with Johnny." And I said, "No, I couldn't." He said, "No, no, no. You do it, you do it. It's great. He's fine." And so I did and it was one of the great experiences of my life. But the amusing thing, of course, is I wound up at Maryland with Charlie Misner, and we worked together on a bunch of projects. He was a co-PI on the first project on education I worked on. We were using computers in physics education through the '80s. So that was amusing.

One of the reasons that I didn't wind up in general relativity is that I wound up at MIT for a variety of reasons, and there wasn't anybody interested in doing general relativity at MIT. There were very few places that did it, and I didn't realize that.

And for the next couple of decades, as I got to know some of Charlie Misner's students and so forth, it was really a field on the outs. I knew many of the students who were graduate students in general relativity because for some years I was the graduate advisor. I would interview Charlie's students, and I would say, "You know, there are not a lot of jobs in general relativity." And I would get things like, "I know I'm going to have to go on and do something else. But this is such an exciting opportunity, that if I get to work five years in general relativity, that's enough." So it wasn't really a hot topic. I wasn't one for following hot topics anyway. Because if I had, I would have been a particle physicist.

Zierler:

That's right. That would have been the time to go into particle theory.

Redish:

Well, I started, and I tried to work with some people in particle physics. But first of all, I didn't want to be at the cutting-edge where it was so cut-throat. I was looking for more cooperative science. My sense, and I had a lot of evidence of this later in later years, was that the people at MIT were looking to win a Nobel. And many of them were nasty about it. They were willing to use students. But there were a lot of good people at MIT – a lot of really nice people. Many of them influenced me a lot. I worked with Felix Villars, and he was a wonderful gentleman, and I met Victor Weisskopf, who was a huge influence on me, even though I only met him a little bit. And Steve Weinberg, I took a course from Weinberg. He was great. But there was a lot of, if you're not going to be absolutely first-rate, go away. I'm not interested in you. I just wanted to do my work.

But more important was that I found many-body theory. My friend Royce [Zia] and I had a course with Petros Argyres in quantum many-body theory, non-relativistic quantum many-body theory. (I did also have quantum field theory, which is relativistic many-body theory.) I saw stuff where you knew all the physics. It was just quantum mechanics, perfectly well-known. You knew the interactions. It was all electromagnetic interactions. But there were 1023 particles. So you had no idea where to begin. The calculations that you looked to try? Often, they blew up. One critical example for me was the perturbation series for the electron gas that just transformed my view of the whole physics world. Every term in the perturbation series was infinite. But what Gell-Mann and Brueckner did was they looked in each order, took the worst divergent term, threw everything else away, added the remaining terms up under the interval, and now it converged and you got a reasonable answer. I looked at that and I was just enthralled.

Zierler:

Why? What was so enthralling?

Redish:

I guess it was that I was reaching for the fact that the best insights in physics are not about what are the rules of the world, but how can we think about it. And so the fact that, yeah, I could put it into a computer maybe and get a result with 104 particles and maybe calculate a number… I felt that doesn't matter. That's not useful. Yeah, it shows you your equations are right. But if you have a way of thinking about it that can pull out of that mess and say, "Okay, here's a toy model. If you organize it so that the toy model is at the center, now everything can be organized around that and it will all make sense." To me, it was like a super-saturated solution, where all of a sudden, you had the seed that made everything coalesce into something coherent. That to me was brilliant. I loved it.

I later came to see the atomic shell model that way. Because I just fairly recently, like I don't know, 15 years ago or 20 years ago, I started to do some estimations. I was working with estimations and calculating one of the energies. If you just say, "I have a nucleus and I have 20 electrons, how would you organize that?" The first thing that you see is the strongest force is the repulsion between the electrons. Your best idea ought to be to spread those out in a net. Keep them as far apart as possible. You know, that was the original classical model, right? If you try to do that, that doesn't work at all. It's totally useless. The best thing to do is ignore all those repulsions, average them out, and take a single particle model and, woah. Now you've got the atomic shell model, now you've got chemistry. You've got all of chemistry. If you look at the history of it, it's totally bogus how we got there. It was pure accident of the various models that were created. If we had been starting from a lot more information, it would have been murder to get that model. But the whole idea of having models that you could think about that organized a whole subject for you, just really excited me.

My friend Royce started out as a particle physicist and did his thesis in particle physics, then he went to CERN. But he used to argue in favor of the fundamentals. You know, fundamental laws. And I would argue in terms of the organizational fundamentals. And so we eventually decided to call it fundamental sub-1 and fundamental sub-2. And we both agreed to disagree that they're both fundamental, just in different ways.

Zierler:

So Joe, when you entered MIT, nuclear theory was not yet a done deal? That would come over the course of your study?

Redish:

Still not a done deal.

Zierler:

(laughs) Well, you have to get a degree in something.

Redish:

Yeah. A lot of the problems that I worked on have not yet been solved. What I was trying to do and failed, but had fun doing, was to organize nuclear reaction theory in terms of the number of active particles involved, to synthesize multiple very different and apparently contradictory approaches. What happened is, it got too hard, and other topics became more interesting. It was what Steve Koonin calls "agenda-hopping." You declare victory and move on, right? You didn't solve it, but you realize that you're not going to solve it now. You're going to move on to do something else where you think you can make more progress. I was disappointed because I think we could have learned a lot more about nuclear physics in the past 30 years, but it was hot when I was starting, and I was mostly interested in mathematical techniques, and how you organized clustering and stuff like that. So it was more about theoretical and abstract issues.

Zierler:

Who were some of the professors at MIT that you grew close with?

Redish:

Well, Felix Villars, Earl Loman, Bill Bertozzi taught me nuclear physics. Bill was an experimentalist and helped me shift to a more real-world perspective. Francis Low also had a big influence on me. I had a wonderful course from him in Statistical Mechanics. Steve Weinberg gave a marvelous course in Quantum Field Theory.

Zierler:

What did you focus on for your PhD thesis?

Redish:

My PhD thesis was nuclear reactions using quantum field theory, non-relativistic quantum field theory. One of the problems in quantum field theory was in describing clusters. And of course, in nuclear physics, clustering is one of the biggest things. I was trying to find ways to get corrections due to the fact that the nucleus was not of infinite mass. And I found some ways to make some progress.

But what happened in the next couple of years, I decided quantum field theory was just not an appropriate structure for non-relativistic physics. Yes, it handled the symmetry perfectly, but it didn't handle the clustering which seemed to me the critical element. So I went instead to an alternative theoretical approach, which handled the clustering, but you had to put in the symmetry by hand. That was basically the Faddeev theory approach, where you looked at various coupled equations for different clusterings and interactions.

Then the question was, how do you put in the boundary conditions in general. For N-body systems there was a lot of cool mathematics. What was fun was that you were continually thinking about, well, what's the physics? How does that physics get represented in the mathematics? Well, that physics is not well-represented in the mathematics, but that's what's important, and so how do I rearrange the mathematics so that the model matches the physics. What's happening? That was fun. I had great pleasure out of it. But it wasn't enough.

I got hired as an assistant professor, promoted from my postdoc. In the first year, one of my colleagues said, "Okay, now you're going to have teaching, but your teaching doesn't matter at all. You just do the minimum you can to get through that. Emphasize your research, because that's what's going to get you tenure." I nodded politely, but proceeded over the next four years to do all kinds of creative and innovative stuff in the teaching. I put in new stuff, I made students do projects, I was designing things in new ways. I put a huge amount of time into my teaching, and I interacted a lot with my students. And so it was clear, as I look back, it's clear that there were some threads that were visible in my orientation, in my spirit that eventually played out appropriately. I look back and I say, I'm glad I did everything I did.

Zierler:

Joe, who was on your committee?

Redish:

My PhD committee?

Zierler:

Mmhmm.

Redish:

Arthur Kerman, Hermann Feschbach, let's see, who else? Probably Bill Bertozzi. And of course Felix Villars. The good guys in nuclear physics.

Zierler:

On the social side of things, being in Boston in the late 1960s, did you involve yourself at all in any of the movements on campus?

Redish:

No, actually, I did not. And I feel guilty about that. As I look back, I'm pretty sure I'm at least a touch, if not a whole lot, Asperger's. I was fine talking with people about my work, but more emotional social issues were not always comfortable for me. Ginny and I went to a few protest marches when we got to DC, but I was focusing on my physics. I now feel that I didn't do enough getting involved in the social issues of my time, and I regret that because I feel strongly about them.

Zierler:

What would you have wanted to do? What were some of the issues, looking back?

Redish:

Oh, I would have wanted to march against segregation, for civil rights, and so forth. No, those were terrible, terrible issues. And the war. The war came, and they started drafting people. I was pretty lucky because I was already in graduate school when they started drafting, when they turned off student deferments. People started going to graduate school who wouldn't have gone to graduate school if they weren't trying to get out of the war. Then they gave a test, like a GRE, which I took and I passed. And I got married and we had a child very quickly. So I was safe even when they stopped doing the test, and were going to draft grad students.

A friend who was just a year behind me wound up leaving MIT and going to teach high school. He wound up being one of the top high school teachers in the country, being an award-winning teacher, a big player in the AAPT, and doing wonderful work. That's where I could have wound up and been safe.

Zierler:

After you defended, what opportunities were available to you? What kinds of postdoc positions were you thinking about?

Redish:

There were only two in my field. In the whole country. And I was very lucky to get one. Well, actually, the funny thing is, I applied for 100 postdocs. But there were only two that matched pretty well what I was doing: one at Maryland and one at Minnesota. Ginny, who was pregnant at the time with my son, said, "I am not going to raise this baby in the bitter cold of Minnesota." That baby is now a full professor at the University of Minnesota Medical School. (both laugh) And he has been there for, I don't know, 20 years? (laughs) He and his wife have raised three of my grandchildren there.

Zierler:

It skips a generation. Becoming a medical doctor and being willing to live in the cold.

Redish:

He's not a medical doctor. He's actually trained as a computer scientist.

Zierler:

Oh wow.

Redish:

But he, like me, was interested in cognition. In how people think about things. He was a computer science major as an undergraduate, but then he got his PhD in computer science with the computer neuroscience joint program at Carnegie Mellon and Pitt that had just started up. He was one of their first PhDs, and he wound up doing computer modeling of rodent navigation using some AI techniques. He then wound up getting—this is a whole 'nother story, but he wound up getting a postdoc with a top neuroscientist and learning how to build a laboratory, and then they hired him at Minnesota to build a laboratory. Now he has his own laboratory, and he has a hundred papers and 10,000 citations. (laughs) So lots of students. But Minnesota is just the irony of ironies.

Zierler:

And so where did you end up?

Redish:

I ended up at the University of Maryland.

Zierler:

For the postdoc?

Redish:

For a postdoc. I got a postdoc at the Center for Theoretical Physics, University of Maryland, one year renewable for a second.

Zierler:

And the idea was to work with Misner?

Redish:

No. No, I was working with the nuclear people.

Zierler:

Right.

Redish:

Nuclear people. Yeah, I worked with the nuclear people. I actually went to work with Leonard Rodberg, a previous student of Felix Villars's who was in fact not on campus, because he had taken leave to work with Daniel Ellsberg on the Pentagon Papers.

Zierler:

Oh wow.

Redish:

And in fact, I never saw him. He wound up involved with the Pentagon Papers and testifying in Ellsberg's trial and so forth. But I wound up with the other nuclear theorists at Maryland. Bill MacDonald, Jim Griffin, Manoj Banerjee, Carl Levinson, Jerry Stephenson. They were the nuclear theorists in our group. I wound up working mostly with, Jerry who was another Villars student. We published a bunch of papers together.

Zierler:

Was the plan for you to take on new research for your postdoc, or were you looking to revise and expand your dissertation research, as a postdoc?

Redish:

No. I wanted to look for new stuff. But I had my own ideas of what I wanted to work on and what I thought was interesting, and I was not handed a project. Jerry, Harvey Picker (another postdoc), and I kind of negotiated a project. It was funny because in my, what, second year, a fellow from MIT, a friend from MIT, got a postdoc at Maryland. He came and talked to me and he said, "How do you come up with new projects? I'm supposed to come up with some ideas. How do I do this?" And I kind of looked at him funny because that was not one of my problems. My problem was figuring out which of my many crazy ideas were actually worth spending some time on. That was why it was good to have good people to work with. We kicked a lot of ideas around, and they helped me purge the stuff that was totally unworkable to stuff that was just a little bit crazy and might have gone somewhere. Didn't. But you know, I wound up publishing 75 papers or something in nuclear physics.

Zierler:

The postdoc was a limited two-year term?

Redish:

It was supposed to be two years, yes. And then what happened was, the university's physics department had expanded like crazy. In 1957, they had four faculty members. They then hired John Toll, a student of John Wheeler's, as Chair, and in 10 years, he expanded the department to 57 faculty. It was, you know, Sputnik, and we just exploded. But we had too many assistant professors. Not all of us were going to get tenure. When the postdoc finished, Maryland offered me a visiting assistant professorship for one year, with a strong hope that they would be able to convert it in a year or two to a regular assistant professorship. And they were able to, and the rest is history. I'm still there. (laughs) They never figured out how to get rid of me.

Zierler:

Did you ever go on the job market formally, or the postdoc sort of just morphed into a faculty line position?

Redish:

No, I did apply. I applied for a bunch of positions. I was offered a regular assistant professorship at Brooklyn College. I was offered a postdoc at Case Western Reserve, a very good postdoc. And there might have been a third offer. The tempting one was the regular assistant professorship at Brooklyn College. But I didn't want to go back to New York. They were interested in a lot of the same stuff at Brooklyn that I was, but I'd been happy at Maryland. I'm glad I stayed. It was a lot of fun. Many good people to work with.

Zierler:

And what was the position you applied to? Was it in nuclear theory?

Redish:

Yeah.

Zierler:

Was it a new line, or was this somebody who had retired?

Redish:

Oh, no. These were all everything... It wasn't the stable, replace one by one. All the fields were growing. Everybody was growing. They were adding new areas. And in fact, what happened is, the year I went up for tenure, 1974, the department sent five of us up. And the administration came back and said, "That's too many." They said, "Redish is the junior person, keep him back a year." And so they did. That was okay. (laughs) Didn't hurt one bit. But you know, it was... I mean, we had assistant professors. Some of us got tenure, some of us didn't.

One story that might be of some interest. In my second or third year assistant professor. we hired a hotshot from Stanford who had won a Sloan Fellowship. I had not. He came and gave a talk on a field that I was interested in. I went up to him afterwards and I said, "Hey, I really liked what you were saying. I have some ideas. Maybe we could collaborate." And he looked at me and he said, "You're crazy. We can't collaborate, we're competing. Both of us are not going to get tenure." And I said, "Okay, if that's what you want." He was right. Both of us did not get tenure. But he might have done better if he had worked with me. (laughs) I think one of the reasons that I did get tenure pretty smoothly is because I learned that I like working with people, talking and sharing ideas. I was talking to a lot of the experimentalists.

 

Zierler:

How quickly were you promoted from assistant?

Redish:

As I say, in my... Actually, I was recommended in my third year.

Zierler:

That's fast. That's very fast.

Redish:

Yeah. That's very fast. But I wasn't actually tenured until my fourth.

Zierler:

And the tenure was exclusively for your work in nuclear theory? Physics education was a non-factor at this point?

Redish:

Absolutely. I was one of the people who really talked to the experimentalists. I think I got a lot of support from experimentalists. (laughs)

Zierler:

What do you see as some of your main achievements in nuclear theory up to that point?

Redish:

Oh, I don't know. We had done some-- I mean, I was interested in three-body problems. And so what I was doing, because the crucial idea, and this is interesting, because this comes back to some of my recent work in education where I've been thinking about the role of toy models in physics.

Okay, a little background: One of the problems that students run into, that teachers of physics run into teaching biology students, is we use all these trivial toy models, right? Frictionless vacuum. Ignore air resistance. Treat it as a point mass. And the biology students come in and they look at this and they say, "These are not relevant. This is not the real world." And they know in biology, that if you simplify a system, it dies. You can't do that. In physics we do this all the time. Simple models are kind of a core epistemological resource for us. You find the simplest example you possibly can and you beat it to death. It illustrates the principle. Then you see how the mathematics goes with the physics. The whole issue of finding simple models is where a lot of the creative art is in physics.

My issue in nuclear physics was clustering, and so you have three bodies, that's the smallest number where you have clustering – where you have two interacting clusters and a particle gets transferred from one cluster in the original arrangement to the other. That's a rearrangement reaction. You need at least three particles to have a rearrangement reaction. I spent a lot of time working on three-body problems, and the mathematical techniques to set up scattering theory with three-body was tricky. The standard stuff you learned in quantum mechanics didn't work. It didn't specify the boundary conditions correctly. So you had to do stuff, and there were some Russians who had come up with some clever ideas, but what I was doing was matching, taking those ideas, and matching them to real physical situations, and saying, "This is very interesting. What can you learn from taking these abstract techniques that have been developed in mathematical physics and applying them to real situations?"

A lot of my work today has been about taking things I have learned in other fields and trying to bring it across. Bridging. I'm still getting citations for a 1971 paper on how you treat the interaction of two particles inside the field of a third. When you're building up the scattering of protons from a nucleus or you're knocking out a particle from a nucleus. How do you treat that interaction? My postdocs and I came up with some innovative ways of doing it that got very widely accepted. We still get an occasional citation, decades later. That was the thing that turned out to be most useful.

What I had the most fun with was that I wrote a whole series of papers on how you write a general N-body scattering theory with different kinds of clustering. Organized around two-to-two cluster scattering, but how you then did breakups, and how you built a formal theory. I had learned how to derive quantum field theory from first principles. But that doesn't handle clustering. So how do you go to Schrödinger theory and derive a scattering theory with arbitrary numbers of clusters from first principles? And how do you organize and arrange them? I did that. Actually, a bunch of us did that. There were a number of competing models. Mine was set up so as to be able to match to the experimental data that told you what the organizational parameters were. That's the work I'm actually proudest of, even though it doesn't get so many citations.

Zierler:

Why? Why are you so prideful about that work?

Redish:

Because it does what fits with what I'm trying to do everywhere, which is to see how you make sense, how you blend — make sense of the physical world with the mathematical world. That's what's driven me the whole time, and it's interesting, now that I'm thinking about the teaching, how I'm still doing the same thing. (laughs)

Zierler:

Joe, let's try to get the origin story down for when... I mean, in 1992, when you made this complete switch over to teaching, right. It obviously didn't happen overnight.

Redish:

No, no.

Zierler:

So help me put together the narrative of your earliest inklings that at some point, this is where you would be spending the bulk of your energy.

Redish:

Right, well, I mean the first two years that I taught, I taught sophomore physics, electricity and magnetism and waves for physics majors. And I did a lot of innovative stuff. What I discovered was that I had learned all this stuff four or five times beforehand, but when I sat down to teach it, that's when I really was able to pull it together and begin to make sense. That was the first step.

The second step was that I really liked giving the kids projects. Because I remembered doing project work. Long-term projects where I was in charge. I remembered how useful those were for me; and I wanted the kids, my physics majors, to have that experience. So from the second semester I was teaching, my students were doing term papers that were worth a quarter of their grade. They had to pick their own problems and work it out themselves. I would help a bit and so forth. So that was a start.

But where I really then made the switch was 1980. I was assigned to teach graduate quantum mechanics. That was fun. I really enjoyed teaching graduate quantum. But as I was teaching it, the first, maybe it was the second time I taught it, I said, "You know, I'm a professional quantum mechanic. Everything in this book is all analytic. There are no computational problems, but half of what I'm doing, more than half of what I'm doing, is building code to actually do calculations. I ought to put some computer into the class."

This was 1980. We had mainframes and the students could get on terminals. They often had to get on late at night, they had to learn job control language, which you probably don't know, (Zierler: No.) and you're lucky you don't have to. You needed it to get in and to control the running of the code and, you know, there were not even debuggers. You would run it and then you would have to wait for the data to come back, and then you'd say, "Where's my error?" It was awful. It was a disaster. I said, "Okay, that's not working." (laughs) So I said, "Okay, I'm not going to do this anymore" even though I had 33 students, and when I asked, "Can you program?" 32 of them said they already knew how to program in Fortran. But December '81, the PC came out.

 

Zierler:

Yeah.

Redish:

The IBM PC.

Zierler:

Did you realize immediately the potential for education with the PC?

Redish:

Yes.

Zierler:

Why?

Redish:

I said, "That changes everything." I bought one and I gave my son the BASIC book. He was 11 years old. He taught himself BASIC, and he wound up getting his PhD in computer science. (both laugh) The sticking point for the computer was the mainframes and getting access, and nobody had terminals. We didn't have personal computers. Right? If there was a PC, that was going to make it possible. So I bought one and started to get used to it.

One year later, the chairman of the department's term was over, and somebody said that I should run for chair. They asked me to apply. I wrote a long letter saying why I'd like to, at some point it would be very interesting, but I'm 40 years old, I'm in the most productive research of my career, so I wouldn't do it now. They came back and said, "You know, that was such a good letter. We want you to come back and consider again." (laughs) And so I started to think about it and I said, "You know, I might be able to do something with these computers if I were chair." So I agreed to run for chair.

In 1982, they made me chair of the department. The first thing that happened, actually two things happened that really made a huge difference, and I wound up being right at a couple of pivot points. The first exciting thing that happened was that the four chairs in the college – physics, math, computer science, and the Institute for Physical Science and Technology – started to meet regularly. We met for lunch once a month. And the chair of math (John Osborn) and I got into some discussions. We both were excited about the use of the personal computer.

We put together a proposal for the first personal computer laboratory for students on the University of Maryland campus, to be held in a room in the basement between the physics and math buildings, and we sent it to the Provost. The Provost at the time was Brit Kirwan, who later was Chancellor of the whole University of Maryland system. But he was a mathematician originally. He wrote back to us, and he said, "This is a brilliant idea, but I'm going to put it out to the whole campus as a request for proposal."

We said, "Wait, wait!" He said, "Don't worry, you will get yours." So he sent it out, and they got back something like 81 proposals for computer laboratories. He granted five or six from campus funds (including ours), but from those 81 proposals, he pulled together a proposal to IBM called "Project Fulcrum". (I still have a plaque from it somewhere.) IBM wound up giving the University of Maryland $6 million worth of IBM PCs. So now I had a PC lab and I could think about what to do with it.

The second exciting thing that happened was that I had a colleague who had a joint appointment with physics and The College of Education named John Layman. John was a big shot in the American Association of Physics Teachers. He had been a president, and they have a four-year presidential sequence, so he was very much into the AAPT. At the time the AAPT was located on the campus at Stony Brook (SUNY). Stony Brook said to the AAPT, "We don't want you anymore. You guys have to leave. We need your space." And the AAPT just said, "Okay, we're going to go somewhere else. We're going to buy a building, and we're going to set up our own space."

John Risley from North Carolina State said, "Come to Raleigh, we'll get you a place. We'll be excited about having you there." Jack Wilson, the executive officer of the AAPT said, "That sounds like a good idea." John Layman came to me and said, "They don't want to do that. They want to come to Washington, to us. Because everything is going to be in DC. It's the center of the government. The government is expanding in lots of ways. This is where they should be." So he and I put together a proposal to bring the AAPT to College Park. I offered Jack Wilson a quarter-time appointment so he could teach, and the opportunity to collaborate on education issues.

Zierler:

Joe, when did you first realize that the AAPT was a thing? In other words, you probably started thinking about pedagogy before you realized that there was a whole world out there that you could tap into.

Redish:

That's right. And it was when John Layman brought me this.

Zierler:

Yeah.

Redish:

I didn't know much about it. But then I met Jack Wilson. And Jack-- I don't know if you know who he is. He's now retired, but he went on to become president of the University of Massachusetts system. He had started out as a physics teacher in a small Texas school using computers in the laboratory. Then he went to AAPT as the executive officer. We met and really hit it off. We started to put together ideas. We came up with this project that we called M.U.P.P.E.T, the Maryland University Project in Physics and Educational Technology. If you're not aware (and you should be!) Jim Henson is a graduate of the University of Maryland. Kermit the Frog was born on the College Park campus.

Zierler:

I did actually know that. (laughs)

Redish:

And in fact, we now have a bronze statue (Zierler: Right, right.) of Kermit and Jim Henson. So we named our project that, and started pushing the idea that there could be real value in using computers in undergraduate physics education. We started writing papers. We wrote something called The MUPPET Manifesto (Computers in Physics 2, 23 (July/August 1988; http://www2.physics.umd.edu/~redish/Papers/MUPPETManifesto.pdf) about how computers were crucial for what was happening in physics at the time. How non-linear dynamics was important. How there were all kinds of things happening in physics that you needed a computer for, and the personal computer made it possible to start including these ideas in intro physics.

We put together a proposal with myself, Jack Wilson, Bill MacDonald, Charlie Misner, and Jordan Goodman. We got money from FIPSE, the Fund for Improvement of Post-Secondary Education. We got a grant and created the first environment designed specifically for undergraduate physics majors, freshman, to do programming in their classroom. We chose Pascal, which was the most common language taught in high school. BASIC wasn't quite useful enough. It wasn't close enough to Fortran, which we thought was where everything was going to be forever.

We wrote a shell in Pascal so the kids didn't have to learn to build all the user interface tools from scratch. We wrote subroutines to draw graphs easily, to make menu bars easily, to get data input easily with simple calls. Then we gave the students a frame program, and the students had to write the code for the physics.

And then I resigned my chairmanship early, before the end of my 5-year term. In '85-86 I did a sabbatical, and I came back and for three years taught the intro course for physics majors using computers. I taught the physics majors for four semesters in a row with these materials and had the students all do projects where they had to do computations. They had to come up with their own projects. It was fantastic, just wonderful. I would tell the students that they couldn't get an A on their project unless they taught me something I didn't know. So they had to pick something that I wasn't an expert in. I farmed them out to the rest of the department. I said, "Oh, you're working on that, go talk to Professor Misner. See if he'll advise you on this."

Zierler:

Did your colleagues appreciate this?

Redish:

They did. Nobody ever turned us down. And the number of students doing research in the junior and senior years doubled. Of course, it had started small, and now we have other ways of getting them involved in research, but it was a fun project at the time.

We started writing papers. Jack Wilson and I wrote a Physics Today article (Using Computers in Teaching Physics, J. M. Wilson, and E. F. Redish, Physics Today, 42, 34 (January 1989)) and a paper for the American Journal of Physics (Student Programming in the Introductory Physics Course: M.U.P.P.E.T,, E. F. Redish, and J. M. Wilson, American Journal of Physics, 61, 222 (1993)).

In 1988 Jack convinced me to go to an AAPT meeting. This was my first AAPT meeting, but I've hardly missed one since then. Jack had been telling me about the importance of what he called his three C's: cognitive, communication, and computer. I started going to the meetings, and I met Lillian McDermott, David Hestenes, Dewey Dykstra, Bob Fuller, …. And I said, "Holy smokes. This is really interesting." That was just kind of it. I was ready for it, and so then, as you know, I went to Lillian to learn the ropes.

I made the decision to jump over the cliff, stop doing nuclear physics, and I said I'm not even going to go to any more nuclear seminars. I'm going to do this physics education research full-time. Then, I looked around and I said, "Wait a minute, I thought this was a research field. There are no conferences. Yeah, there are the AAPT conferences, but our stuff is all stuck and scattered around. There's…"

Zierler:

You mean among subfields, it's scattered around?

Redish:

Yeah, I mean it was just... you would have two research PER talks that were very similar, and one would be in one session, and one be in another. Or they would be parallel to each other. Even in the AAPT, a lot of people didn't really like physics education research. It wasn't seen as a coherent field – an important part of physics education. So I said, we need to make this a field.

A couple of events happened that led into how things developed. The first began with Len Jossem. Len was an AAPTer, chair at Ohio State for many years, and a big shot in physics education. He had been on the International Commission for Physics Education (ICPE). This is committee C-14 of IUPAP, the International Union of Pure and Applied Physics. Much of what it does is to identify conferences as legitimate conferences, so people from small countries can know that it's worth sending somebody at high expense to them.

Len had been on the ICPE for 12 years or so and he was getting off. He nominated me as his replacement. I said, "Okay, that's cool" and got to go to Hungary for my first meeting. At that meeting in Budapest in 1993, the committee turned to me and said, "You know, we haven't had a conference in the United States in a long time." I said, "Oh, okay. I can run one." (laughs) So I agreed to create a conference.

In the summer of '96, the AAPT was meeting in College Park. So we piggy-backed the International Conference in Undergraduate Physics Education meeting just before it. I worked with John Rigden of AIP, and we organized and created a conference in physics education, a big chunk of which was about physics education research. We had 325 international attendees from 26 countries, and we published two big, fat volumes of proceedings. In addition to talks, we had workshops and demonstrations of innovative teaching methods. It was a wonderful conference. A lot of people said that that conference made a difference in their taking physics education research more seriously.

Zierler:

Joe, I want to ask. By 1992, when you're fully committed and you're familiar enough with AAPT, right?

Redish:

Now wait, because this-- I need to finish this story, because this is an important story.

Zierler:

Please.

Redish:

And I've built it up, right? The conclusion is coming. Because what I did next turned out to be important. We had our meeting one week and the AAPT meeting was the next week. I wrote to about 100 people I knew were interested in physics education research, and said, "On the Sunday between these two meetings, come to the physics department in College Park and we'll have an "Interval Day" (between the meetings). We'll discuss what PER needs to go forward as a field."

We got 75 people to come to that Interval Day. Three things were set up there. First, we set up The Physics Education Research Conference (the PERC.). Dean Zollman and Bob Fuller agreed to run the first one in 1997, the next year. That was the first PERC and we had about 65 people attend. It has been run every year since then. It now draws 400 people. There are huge peer-reviewed proceedings published from each meeting. It's a fantastic meeting. And it has spawned similar meetings in chemistry and biology.

Next we said, "We need a journal." There was no room to publish physics education research in a place read by physicists (some PER was published in education journals) because the American Journal of Physics said, "We don't publish research. We publish pedagogy." But when you're doing physics education research into pedagogy, you can't always wait to publish until you're done, and you have messages for the teachers. You have to be able to exchange information with other researchers.

I took that on. It took me two years, but I eventually convinced AAPT to let us start a section of the American Journal of Physics for physics education research. I was editor for the first five years. Our first issue appeared in 1999. For the first three years, it was a separate issue in a different color. After three years, it got integrated into the Journal. Until 10 or 15 years later, when we came up with the Physical Review (started by Bob Beichner), AJP was the place to publish physics education research.

So we now had a journal, we had a conference, and the graduate students got together and made a network, an interaction network, which spawned the multiple networks that the intermediate level faculty now have in physics education research. I feel that the Interval Day in 1996 was a turning point, a really important meeting that produced a lot of shifts. So now we can go back, but I wanted to get that story down.

Zierler:

No, that's good. That's good. To dial back to 1992, my question is, and I'm hearing elements of the answer already: By the time you're fully ensconced in physics education and you understand the MO [modus operandi] of AAPT, what were some of the orthodoxies in the field? And what were some of the ways that perhaps you were looking to change those orthodoxies?

Redish:

Okay, yes. (laughs) I have been very much about that.

Zierler:

Yeah.

Redish:

One of the orthodoxies were misconceptions. The idea was that students came in with misconceptions that were sometimes very robust, and those misconceptions needed to be erased and replaced. One of the things I had been reading that really influenced me in the year I was on sabbatical was a book on epistemologies that Len Jossem gave me. (W. F. Perry, Forms of Intellectual and Ethical Development in the College Years, Holt, Rinehart, & Wilson, 1970) Perry's book is about his study of the epistemological development of college students. He talks about how first year students often saw the world as binary with either a right or wrong answer. They would ask "What's the right answer?" After a couple of years, they would evolve and say, "Well, there's no right answer, wrong answer, any answer is equally good." Finally, they would get to a point of saying, "Well, you don't know the exact answer, but given what I know, this is the answer that seems best now." They evolved from a binary stance to a relativist one and finally to a constructivist one. That's my simplification of that work. It really impressed me, because at one footnote in the middle of the book, Perry said, "Of course this is only in fields like history and English. Not like science, where they do have answers."

I looked at that and said, "No! That's wrong. This is just what science is like. If even this expert on epistemologies doesn't understand that, here's a big, wide-open field that I can go into." I found that there were people in education working on these issues. David Hammer had recently done a PhD thesis on epistemologies in introductory physics. I read it and it had a huge influence on me. I said, "Hammer's right. That's what we should be looking at."

So I began thinking about a more dynamic theory of student thinking. I've been part of this process of developing what we call "The Resources Framework", where student thinking is thought of as more context-dependent, more creative, more putting together pieces; where there's more framing, where, by deciding "what's going on here", students will look at something and won't use a lot of things they know. The fact that a student hasn't done something in this problem doesn't mean that they don't know it – or that they have some "wrong thinking" – some misconception.

Zierler:

Right.

Redish:

One of the possibilities is that the students know something but didn't think to use it. The way I have begun to think about this is, not just student thinking, everybody's thinking, is much more dynamic, much more context dependent. My particular take on it is that there are control layers. That is, small bits of data are taken in and used to filter and transform how the rest of the processing is done.

And so an important part for the learning of physics is what I call epistemological framing. You look at a situation – like a class -- and say, "What's the nature of the knowledge I'm learning? What's going on here? What knowledge can I use? What's legitimate for me to use here?" Many of the observed misconceptions students have are epistemological in nature, not wrong physical knowledge.

This has been the non-orthodox perspective I've been trying to push for a number of decades now. There are a lot of people who buy into it. Lillian objected. She thought we should be a-theoretical. Not talk about theory at all. Of course, because she had this problem about psychology.

I had an interesting postdoc who had worked with Lillian's group, who got her degree with that group. I hired her because when I put up my first epistemology stuff at an AAPT meeting, all of the Washington people ignored me except Rachel Scherr, who came up and said, "This looks like a lot of garbage. What are you talking about here?" We wound up having a great conversation. When I had a new grant with a position for a postdoc, I called her up and said, "I want you to apply." She came, she was terrific. The first thing she said was this was all nonsense, and I said, "Well, okay, let's look at this." After many discussions that she and Michael Wittmann and I had, she came up and said, "Well I wrote two papers for my thesis. Let's take that and analyze that as if it were by somebody else, and let's see what their epistemological assumptions are that are inherent because they're not talking about it." She finished up saying, "You know, we were inconsistent." She wound up writing a paper analyzing the epistemologies in those two papers. She bought in and has since then made many contributions to the framework.

You were talking about orthodoxy. One of the papers I'm proudest of in our physics education work is our paper on ontologies. In standard education research, I mean not specifically physics, there was a very famous paper in which the authors (Chi and Slotta) said, "A lot of the misconceptions in science arise from ontological misconceptions. That they're thinking about something as the wrong kind of thing. They think that things are process-like when they're matter-like and they think they're matter-like when they're process-like."

They had an experiment which showed this to be the case. This paper got huge numbers of citations. I heard people lecture on this, and I heard the examples they were giving, and I said, "No, that's wrong. That that's not the way I think about this. I'm a physicist, I don't think about this example this way, and you're saying physicists do this." I was very irritated. (laughs) When their paper came out, Andy diSessa wrote a commentary on it criticizing it, but nobody ever cited Andy's paper or paid attention to it. I don't know why. But they cited Chi and Slotta hundreds of times.

In 2006, a graduate student who was finishing his PhD in plasma theory, Ayush Gupta, decided he didn't want to spend his life doing that. He was interested in people. He wanted to do physics education, and he'd been doing some reading. I talked to him. I thought he was pretty interesting. We actually had a postdoc available, so I hired him.

After a year or so, he was looking for a project, and I pointed out the problem with Chi and Slotta. And so he started doing work. We wound up with some absolutely smashing results. We wrote a paper on dynamic ontologies.( Gupta, Hammer, and Redish, Journal of the Learning Sciences, 19:3, 285-321 (2010).) We got three different kinds of data that showed that their result was only true in the local context, and if you shifted the context or the way the problem was stated, you got different results.

It took us a year and a half to get it published, because, of course, they were reviewing it. First thing that they said was, "That's not what we meant." So the next round of the paper, we had ten quotes from other published papers citing them, showing that this is how people were interpreting that as being what they meant. Anyway, we had a whole series of exchanges, and they wrote a commentary and we wrote a second rebuttal to their commentary, and both were published. I love that paper. (laughs)

You know, my research group at Maryland has been pushing this idea that our theoretical framework could help you understand what you saw for a number of years now. A lot of the people who were purely phenomenological are beginning to come around and say, "Well, maybe you do have to think about the theory. Maybe you do have to think how the students are thinking." The University of Washington just hired a new faculty member, an assistant professor. She just got promoted to tenure. And she's working in the Resources Framework. So yes, I've been one of the group that has been unorthodox, but we're becoming more orthodox. (I hope.)

Zierler:

Joe, at what point did the physics education community start to think in a sustained way about how students from different socioeconomic backgrounds learned differently? How women learned, how minorities, you know, Black students? When did those questions come front and center in AAPT and beyond?

Redish:

Those questions were often discussed in the regular education community, and those of us who went to AAPT, and NSTA, the conference of the National Science Teaching Association, which is more high school-oriented, saw a lot of that. Even through the '90s and 2000s. It was there. There was a lot of feminist research -- looking at the differences between scores based on male versus female. The socioeconomic, the SES, stayed mostly in the education schools. It's only recently, — the last five years, the growth of paying attention to those issues has been spectacular. My research group focuses a lot on this now. But it's fairly recent. Of course, you're asking me to put dates on things, and one of the things I teach is, the brain's memories are not time-stamped.

Zierler:

Not down to the month or year, but you know, in the way that--

Redish:

Not at all.

Zierler:

The community, the physics community in some ways is not isolated from larger national trends.

Redish:

Yeah, yeah. No, right. No, what I'm saying is that you're asking me, "When did this happen?" and I'm saying, well, it happened sometime back in there in my memory, but I don't-- If I don't tie it to a national event with a date, I have no idea when it happened.

Zierler:

Right, right. I guess what I'm asking, it's a very fuzzy question, but the default when you think about physics education, is that there's a default student, right? Who is this student? That student, by default, up at a certain period of time was of course a white male student. And that's problematic because physics in general wants to attract more diversity, but then even when you get that diversity, if you don't know how to teach to different students from different backgrounds, you're not going to get very far in improving physics education.

Redish:

Oh man, you're already so deep into the development. When I started this, there were no students. There wasn't an ideal student. There was, "This is the way you teach physics."

Zierler:

Right. But who are you teaching to if there's no student? (laughs)

Redish:

The thinking back then was that if the student doesn't get it, that's their problem.

Zierler:

(laughs) Yeah, yeah.

Redish:

Okay? The concluding point in my 1994 paper was, "You've got to understand what the students are going through and what they know and therefore you have to ask them and listen to them." That was a radical statement in 1994. There was no ideal student. There was only the physics. The fact that we could come to this – considering the students -- is big progress, right?

I really like this issue of diversity and seeing individual students because what I've been working on recently is the issue of students in other disciplines. Socio-economic status and gender are very important in all of these things, very important, but almost all, 96% of the introductory students we teach at Maryland, no sorry, 94%, I just did this calculation for a talk. 94% of the students are not physics majors. Doesn't that matter? About a third are biology majors and healthcare majors. Another third are engineers. Can we treat engineers just as if they're physicists? It's not quite right. The biologists certainly not. Their assumptions are totally different. Part of this message that I have been working on is that you must know your population.

This project that I've been working on for the past ten years, NEXUS/Physics, is an introductory physics for life sciences class that was redesigned after a year of conversation with the biologists. We spent two hours every week talking about what physics could do for biologists. What are the biologists bringing in? How does this class articulate with their other classes?

We then spent two years giving and watching small classes, videotaping everything. We gave the students lots of opportunity to respond. It gave us a totally new way of looking at this course, because we were actually looking and trying to teach to the population of students we had.

It was a challenge because you don't want to just give them what they want. You want to give them what they need. But how do we figure that out? It's an intellectually really challenging issue, and it's been a huge amount of fun working with how biologists look at physics, and what the difference is.

How does understanding that difference help? I would love to see people doing the same thing, to be working with these different cultures, to understanding how culture brings in different ways of thinking. This is something I've been trying to push, which is a little bit beyond where the diversity researchers are at right now. They're interested very strongly in the issues of power, because it's hugely important. The perception of power, and the perception of expectations, about who you are and where you go and what your identity is. These are immensely important. But there are some interesting general questions, just about the nature of culture — what culture means in the context of communication. I've been trying to do some reading on that, and one of the things is that there are some hypothesis in sociolinguistics about how the language one speaks affects the view one has of the world. It's called the Sapir–Whorf hypothesis. I don't know if you've ever heard of this.

Zierler:

Sure.

Redish:

Yeah? And I've been reading some evolution of these ideas, where people are saying, you know, it's not just the language. The language is about communication, and that occurs in a context. That context involves lots of things that you know and expectations that you bring, your whole culture. I've been reading works on translation. Because in some sense, when you're trying to communicate with anybody, you're translating. And you're translating not just across language, but across deeper issues that are cultural. For example, when I'm trying to teach physics toy models to biology students, I have to be aware that their expectation is that simplification is dangerous. Simplify an organism and it dies. If I want them to accept a physics approach with simplification, it's got to be well-motivated, it's got to be explicit, and it's got to be understood how you're going to use it. This changes everything in the course — how I present everything.

Okay, that's about biology. I also know that 15% of the students in my class are form Africa. And I have lots of students from Southeast Asia. Many are recent immigrants. First generation college students. So, yes, there must be cultural issues there, major differences comparable to those between biology and physics. What does that do? I have no idea. I don't think people have been helping us figure that out. And we need to.

Zierler:

Joe, I wonder if you can characterize generally. Once your colleagues got wind of your shifting interests and that you were fully jumping into physics education research, I wonder if there were maybe two general reactions. You know, as a nuclear theorist, you're a physicist's physicist, right? And so I wonder if you got pushback to some degree that you were abandoning the field? That would be on one side of the spectrum. And perhaps on the other side of the spectrum was, you were taking your deep understanding inside physics, and by applying it to physics education, perhaps some of your colleagues would be appreciative, because who doesn't want better students with better outcomes?

Redish:

That's right. And that is exactly what happened. I could tell you a story that might be amusing that I think is going to be published in Lillian McDermott's memoir.

I came back from that year with Lillian and her people and I said, "Okay, I'm going to do this. I'm really excited." I wrote up a brochure about research and physics education. I had just gotten a grant, I could hire some people. Now I've got a student so I was off and running. Right? I hired a postdoc. We had a departmental review about three years in. On that review board were two people I actually knew pretty well: Lillian McDermott and Ernie Moniz. Ernie Moniz was a nuclear theorist whom I knew from nuclear theory, and he happened to be chair of the department at MIT at the time. Also, the future Secretary of Energy under Obama.

For the review, each group in the department would come up and present what their work was and what they were going to be working on. I was nearly the last because we were new, we were low on the priority chain. So I stood up and started to talk about it. Said, "We're going to do this, we've got this idea. We're going to be doing research into physics learning." And Ernie said, "Wait, you can't do that. You'll degrade the degree." I prepared to defend myself, but then Lillian McDermott turned around and took him apart. I didn't have to say anything. Lillian McDermott was our Ruth Bader Ginsberg. She was small but fierce, and she persisted. She defended me and the whole effort of PER in a Physics Department. When the report came out, she had won. It strongly supported my group.

There were people in my department who after that said, "Well maybe we shouldn't do this. Maybe we should take a vote. Maybe--" I responded, "Look, first of all, you're imagining what my theses are going to look like. Let's wait until you've seen a couple. In physics, we have people who do research on economics, we have people who do pure engineering, building an apparatus. You know, we do things that are not just physics, but that use physics thinking. And this is in physics. So give me a chance."

They let that slide. Partly because of this, I kind of isolated myself from the department. The department chairs said they would let me teach the classes and do research in those classes, and do development in those classes. Sometimes I would test in other people's classes. But I decided I wasn't going to try to change the way the department did anything. That was, in a sense, unfortunate, because it left me as somewhat marginal in the department throughout the entire rest of my career.

But in a sense it was the right decision, because I was really interested in the theory, and because I wasn't interested in becoming the administrator for the large classes. But I was interested in what you could learn about the students from those classes. That was actually better for me.

The McDermott group runs the big classes and that takes a huge fraction of their effort. The group at University of Illinois also does that. I chose not to do that. I'm a little sad because what it means is a lot of the innovations we created and that we published and that other people use, after I moved on to other classes, got dropped and were not used at Maryland anymore.

But you know, I was not in it to transform the teaching at the University of Maryland. I was in it to transform the way we thought about teaching all throughout the community. The way that I wanted to do that was reaching the agents of change in AAPT, not every faculty member at College Park. I did have influence on a number of instructors at College Park, faculty members who said that I inspired them to change their ways of teaching, and I was able to be helpful to some of them. But I didn't majorly transform the way things were taught at College Park.

Zierler:

Joe, on feedback mechanisms, to know that you're on the right track, that you're doing things that are actually on a nuts and bolts level actually improving student outcomes, right? There's the micro where you might be able to immediately get that satisfaction in the classroom, that all of these efforts are paying off. How do you translate that to the macro, to the national level where more and more faculties are aware of your efforts and those of your like-minded colleagues, and that students are learning physics more effectively? What are the feedback mechanisms to know that you're on the right track?

Redish:

Well, one of the feedback mechanisms was that I got invited to give a lot of colloquia. And I wrote papers for Physics Today, The Physics Teacher, and the American Journal of Physics, and people are reading them, and they get cited a lot. One year, one of my chairs in a faculty meeting responded to a question as to whether the department should be doing PER by saying, "Look, I go out to departments all over the country, and they don't ask me, 'What's Bill Phillips (our Nobel Laureate) doing now?' They ask me, 'What's Joe Redish doing?'"

Zierler:

Yeah, yeah.

Redish:

Well, you know, so, eh....

Zierler:

Which is coming from where? What place is this coming from, this interest?

Redish:

Well, in physics, there are people who care about teaching. We're physicists. We like to do a good job and we love our subject. So we're frustrated when we teach the students and it doesn't work. Unfortunately, one way out of that is to blame the victim, to say, "They're not good enough. They can't do this. This is not who I'm trying to teach." But the other is to say, "Something interesting is going on. Let's understand it and see if we can improve it." There are people in both groups. If you want to improve the teaching in your department, the first thing that you have to do is shift the responsibility for learning from purely on the student's shoulders, to at least partly on the faculty's shoulders. You have to accept the idea that we're trying to add value to all students. That it's not just the future physicists that we care about.

I talk to my students in my pre-med class, my biology students, a lot of them are pre-meds, about 40%. And I say, "You're my future doctors. Some day, maybe 10 years from now, I'm going to be rolled into an emergency room, and one of you is going to be there. I'm going to do everything I can now to help you learn to think and to do scientific diagnosis and not to treat me like a memorized fact." So I think it matters. And I think it matters to all of us.

Zierler:

Joe, at what point did the field develop to the extent that you could do graduate work and get a PhD in physics education specifically?

Redish:

Maybe '82, '83? Lillian McDermott started that group at the University of Washington. They started producing those PhDs in physics at a pretty decent rate. UMass Amherst started probably in the late '80s. Bob Fuller at Nebraska was doing some in the late '80s. And then in the '90s, a lot of us added on.

Zierler:

And so for your chronology, at what point were your graduate students, when did you reach that critical mass where more of your graduate students who were doing more education than nuclear theory?

Redish:

Oh, I never had them both at the same time.

Zierler:

It was a clean break?

Redish:

It was a clean switch. My last PhD student in nuclear physics graduated in 1988. And my first PhD student in PER didn't graduate until 1999. So I had a 10-year break because I was starting again. I had to get money and if you first get a graduate student in 1994, they don't graduate until 1999. But I had four students who graduated in '99 and 2000. Three of them are faculty members now. Full professors. One of them has been president of the AAPT. (laughs) So that first crew was exceptionally good. And we've had lots of terrific students since then.

Zierler:

Joe, maybe it's a silly question, but as a mentor to a graduate student in nuclear physics, nuclear theory, obviously there's an expectation on you to have a mastery of the literature, right? In education, what are your expectations that they know the physics, but that they know, you know, first principles and pedagogy as well? How do you create that expectation for what it means to achieve a scholarly mastery in physics education?

Redish:

That's right. I tell students when they come in, I say, "Look, this is going to be harder, because you've--"

Zierler:

It's almost a dual curriculum, right?

Redish:

That's right. "I'm going to make you read the education literature. I'm going to make you read psychology. I'm going to make you read sociolinguistics. I'm going to make you read cognitive linguistics. I may make you read some neuroscience." And they have to take the regular physics grad courses and pass the same qualifiers as all other grad students in the Physics Department.

Zierler:

Joe, to work the narrative up to the present, of course, we connected through Brad (Conrad), and so I wonder if you can talk a little bit about your partnership with him and AIP generally?

Redish:

I haven't had a partnership with Brad. What happened was, and this has happened about a fair number of students, whom I now see at AAPT meetings, is that they were my TAs when I was transforming one of my classes.

I remember the first year that I put in Lillian's tutorials, and we were transforming to an interactive class where the point was to get the students working. The chairman of the department hand picked for me the guy he said was the best TA who had the best reviews from the students. In our recitations that he was leading, we had lab tables and the students were supposed to be working in groups with large sheets of white butcher paper, working stuff out. When I came in to our first class and checked him out, he was leaning over the paper with six students around him watching him show them stuff. I grabbed him by the collar, gently pulled him up, took the pen out of his hand and said, "You're not doing it, they're doing it." (both laugh)

Over the years, we have had many TAs, and we've always had a TA training session where the TAs are taught that you've got to make the students do it. You can't just explain it to them. You've got to hear what they're saying. You know, so I've taught a lot of TAs, mentored a lot of TAs, and many of them have come back to me and said, "That experience changed my way of thinking about teaching." And Brad is one of them.

Zierler:

Joe, to take the narrative right up to the present, you know, we started where you said of course that it was time to retire so you could get some work done. So give me a sense of, what your goals are now. What do you want to accomplish in the here and now?

Redish:

Oh wow. There's a lot. (laughs) In the past ten years, 20 years, 25 years, I have built collections of materials that I've shared on the web with people. One collection I call "Thinking Problems in Physics". These are kind of tricky. You have to really puzzle them out. Students' expectations are often to just ask, "what's the answer?" With these questions it's not a straight shot to the answer. Many of them, it turns out, are about building the blend between physical concepts and math; learning to think about physics with math.

I have 600 problems, and I am putting them in a place where they can be reused, recombined, used in various different ways. I want them to be available after I'm gone. Because I'm 78 years old, right?

I also have 400 pages of text that's basically a textbook on a Wiki for this Introductory Physics for Life Science class. I put all the Wiki on Compadre (https://www.compadre.org/nexusph/), and we're putting the problems there too, and the solutions into the Living Physics Portal (https://www.livingphysicsportal.org), both managed by AAPT.

What I'm trying to do is help the community shift their view of the introductory physics class to focus not just on the content, but on skill development. And to develop an awareness of the different populations we are teaching – biologists, engineers. Not, you know, the more general cultural issues. I haven't figured that out yet.

But I've developed a set of tools for learning to use mathematics in physics. I've gotten elements of your toolbox that I call "epistemic games" in the research literature. I've written six papers about those that I hope will be published in The Physics Teacher, I'm posting them into the Learning Physics Portal and each one will come with a set of problems, activities, and test questions that cut across the content of the course.

So, for example, you want to teach your students how to read the physics in a graph. Well, here's a problem in kinematics. Here's a problem in electric fields. Here's a problem in oscillations. All of these build the same skill. And here are the problems the students have with building this skill.

As a second arrow in my quiver, I'm also looking at changing the content for the Introductory Physics for Life Science (IPLS) students. Typically what we did before the NEXUS/Physics project was just give the standard Physics 1 table of contents and say to the biologists, "Pick which topics you want us to do."

What we did differently for NEXUS/Physics was to say, "What's not in here that we should be doing for this population?" One of the things is fluid dynamics: wet water, viscous flow, internal cohesion, negative pressure. All kinds of fluid stuff that's really important in biology that we never touch in intro physics.

I'm working with Dawn Meredith (UNH) and Todd Cook (UMD), my biologist friend, to write a series of four papers, I hope. Two for the physicists, two for the biologists, about how to do fluid dynamics with physics for biologists more realistically.

I also want to write a paper on chemistry in a physics class. All of the IPLS students have had chemistry. Bruce Sherwood and Ruth Chabay have shown us that for engineers, you don't have to do physics purely historically. (Sherwood & Chabay, Matter & Interactions) You can assume your new engineers know about atoms and molecules. The same is true for the biologists. They also know about cells. They know about DNA. They know about proteins. So we can use that in our physics course.

So we asked the question, "what is there about chemistry that we can support from the physics side?" The chemists have to do everything, and therefore they don't have time to do a careful analysis of a few toy models which might give the students better insight into what's really going on. The NEXUS/Physics team has written five papers on that topic, but I haven't written the summary paper on that for the general instructor. It would have to talk about what the biology is and what the physics is, because most of us physicists don't really know much biology.

Where I started on this was a complaint Todd Cook made to me in 2004. I met him at a campus teaching discussion group. He said, "You physicists, you're supposed to be helping us and you don't teach the right stuff." I said, "What do you mean?" He said, "You don't even teach the Hagen-Poiseuille equation." I said, "Teach it? I don't even know what it is!" (laughs) And he explained to me and I said, "Oh, that's just Ohm's law in the pipe. Is that important?" And yes, it's important. So I learned about it, and I had already been teaching it somewhat, because I was using it as an example to get students to develop intuition about electric currents. Turns out, for biologists it's way more important than Ohm's law.

So that's what I'm working on. I don't know how much time I have left. I'd love to have a full ten years, because I have new things I want to learn. I want to learn about culture and how culture fits into the resources framework that I think about. I've learned a bit about semantics. I've learned a bit about embodied cognition and cognitive blending. Those are immensely powerful ideas. People are now beginning to use it in PER. I think it's very important. But there are comparable ideas in culture. We need a way to get from that immense literature that's anthropology and culture and psychology, and to get something we can understand and use in a simple toy model way, the way we physicists like to think, and not have to dig through this huge mass of stuff.

Somebody needs to be digging through it, and I'd like to have time to do some of it, and pull out what can help us make sense of some of the complex things. Because thinking and learning is one of the most complex things that happens in the physical world. To try to treat that as if it's trivial or obvious is a big mistake. We have to figure out what it is we need to bring to our teaching to help us make sense of what's happening in our classrooms.

Zierler:

Joe, let me ask you a cute question as we round out our conversation. Is there a grand unified theory for physics education? Is there something that you and your colleagues are working toward that even if the larger physics community is not into the weeds as you are, there's a single, unified thing that everybody should know to improve the field and to improve student outcomes?

Redish:

That in fact is what the Resources Framework is trying to build. And there are multiple pieces. The mind is a many-body system. There's not a single thing. If there were a Lagrangian, it would be like I said about the electron gas. The fact that you have "the basic physics" doesn't help you. You have to know how to organize it and what the principles are. This is what I was trying to get started in my 1994 paper. (Implications of Cognitive Studies for Teaching Physics, Am. J. Phys., 62, 796-803 (1994).)

I have a series of papers which are trying to do this in more and more refined and complex ways. So I have the paper in '94. Then I did one on a theoretical framework for physics education research at a Fermi summer school lecture in 2003. It's published in a big $200 book, but the preprint is available on archive (http://arxiv.org/abs/physics/0411149), and that's gotten nearly 400 citations on Google Scholar.

I wrote a paper with Carl Smith on the role of the neuroscience in Physics Education for the Journal of Engineering Education (J. Eng. Educ. 97, 295-307 (July 2008)). And so we're building towards this. There are a few basic principles about thinking: the dynamic response, the idea of chunking or compiling, that you bring up bits and pieces and organize them together, and they activate in…. You know, I give an hour lecture on this stuff. My Oersted lecture also tries to do this; how does the theoretical frame look at this stage, and what can we expect of it? What is it that we're learning from psychology, from sociology, from sociolinguistics? (Oersted Lecture 2013: How should we think about how our students think? Am. J. Phys., 82 (2014) 537-551)

That's the most refined version I've got right now that's coherent and tries to be comprehensive. I could sit down and write a book on this, but I'm not sure I want to do that. Let other people do it. But I think there is one. There definitely is one, and my son, who's a neuroscientist and is very interested in the whole theory of decision-making and judgements, said, "This is a theory. It's coming, it's coming. We're getting there." (A. David Redish, The Mind in the Brain (Oxford U. Press, 2013)) There are many threads that are coming together, and there are many evolutions in this development of our understanding of thinking and learning, that are occurring all across the scientific fields, and so something is happening, and some point there will be a phase transition.

Zierler:

So as a final comment, then, it's undeniable that you're optimistic about the future of the field.

Redish:

I am, and I would love to be around for another 30 years to watch it develop and help do things. I don't think that's likely, unfortunately.

Zierler:

Well, I hope you're around for as long of it as you possibly can and want to be.

Redish:

Thank you so much.

Zierler:

Joe, on that note, it's been such a delight spending this time with you. I'm so happy that we connected, and your recollections and insights over the course of your career really will be so important for our collection. So I'm so glad we were able to do this. Thank you so much.

Redish:

Happy to be able to do it.

[End]