Nanney Autobiographic Essays
Home | Candide in Academe | Tilting Windmills

Candide in Academe Meets Tracy Agonistes
A Memoir of the Morning of Molecular Biology :
Coming of Age in Bloomington - 1946-1951
(
Draft of March 23, 2004)

  1. Apologia
  2. The Place and Time
  3. The Shaharan Butterflies
  4. I am Launched
  5. The Landing Field
  6. Bloomington Dramatis Personae
  7. Nucleocytoplasmic Interactions
  8. The Sonneborn Research Group
  9. The Sonneborn Seminars
  10. The Sonneborn-Jennings Connection
  11. Cytoplasmic Inheritance goes International
  12. Franco American Relationships
  13. The Ephrussi Business
  14. The Conference at Gif-Sur-Yvette
  15. Ernst Mayr and the Species Problem in Protists
  16. Candide in Academe Bibliography

Apologia

When writing a scientific paper, the last section that is written tends to be the Introduction, because the writer doesn’t understand fully what the Conclusions are until the rest of the paper has been written - the Materials and Methods, the Results, the Analysis and Discussion. And the writer can’t construct the Introduction until the Conclusions are established. When writing about a life, whether one’s own or someone’s else, the same considerations hold. Of course, this essay isn’t a scientific paper, but it is concerned with some lives in science; the Conclusion hasn’t been written for all these lives yet, though the mezzo may be returning soon from her meeting at the Weight Watchers.

Part of the challenge of this essay lies in the duplicity of its subject matter. It is both autobiography and an exercise in hagiography. One of the subjects is the author, and his life is known in great detail, but his image is badly distorted by his protective ego. The other subject is seen only in part and those parts are magnified out of decent proportions; we attribute to others in our lives the characteristics we need for them to have. As fellow graduate student Dick Siegel said recently,“Tracy Sonneborn was like a God to us, wasn’t he”. Part of my reason for doing this filial reconstruction is the belief that our impressions half a hundred years ago were not totally the naive judgements of impressionable and idealistic young men. Certainly our opinions were shared by others with greater experience and emotional distance. After George Beadle was awarded his Nobel Prize in 1958 (with E. L. Tatum and J. Lederberg), one of his first acts of judgement on his discipline was to nominate Tracy Sonneborn for a Nobel Prize also. I know, because he asked me to help write the nomination papers. I must not have done a very good job; at least it didn’t work, but I keep trying.

Parts of this story I have written before. When Tracy died in 1981 I wrote and published "Tracy Sonneborn: An Interpretation" (Nanney, 1982). Later I wrote an unpublished record of his scientific conflicts, which I called "Tracy Agonistes", and which I shared with some historians of biology, particularly Jan Sapp and Richard Burian. Jan Sapp (1987) used some of this material in the construction of his book, Beyond the Gene: Cytoplasmic Inheritance and the Struggle for Authority in Genetics. Judy Johns Schloegel (1999) has subsequently discovered much more about the Sonneborn-Mayr conflict than I had ever suspected.

The autobiographical component of the present essay was anticipated in another private exercise I wrote just a couple of years ago and called "Candide in Academe". As a piece of writing it was a disaster, as attested to by both my wife Jean, and my friendliest critic - Dick Siegel. I had tried to write an autobiography in the 3rd person singular: "David was born in 1925....". This was an exercise I had used in my science writing course when trying to help students gain an "objective" perspective on a topic they were too close to. The device may have had some merits as a pedagogical ploy, but it didn’t work very well for an extended opus. Still that "portrait of the scientist as a young man" served to arrange long ago events in my head, and is a large part of this dual biography. The things omitted from this exercise are any serious look at personal matters, and any consideration of my career as an educator - which I have always considered my primary calling. I also have not attempted to tell here the story of Tetrahymena, the organism that I studied after leaving Paramecium. I did refer to some of these stories in a talk (Nanney, 1998) I gave at the University of Pisa in 1994, in exchange for an honorary degree.

Actually this present account has yet another subject, that alluded to in Johnny Preer’s (1997) recent essay, "Whatever Happened to Paramecium genetics"". The story of a science is always entangled with the story of its technology. In the case of genetics the technology, up until sequencing took over, was the technology built on the creative use of "domesticated" organisms. One cannot recount the history of genetics without considering Mendel’s peas, and Morgan’s fruit flies, fields of maize, cages of mice. Describing more modern genetics requires discussions of Neurospora crassa, Chlamydomonas reinhardii, Escherichia coli, phages T2 and T4, Saccharomyces cerevisiae, Arabidopsis thalliana , and Coenorhabditis elegans. In this story just out of focus behind the human actors is the golden slipper animalcule, Paramecium aurelia, which for a time shared major billing with the other genetic technologies, and may yet have a comeback. Genetics has provided the playing field for an organismic natural selection, with outcomes that have had significant repercussions in the careers of geneticists.

I must, up front, acknowledge that the tales I spin are personal tales, with no serious attempt at documentation, and bear a spin that others do not share. I have been particularly troubled that my perspectives on Tracy Sonneborn were not endorsed by Ruth Sonneborn, his lifelong companion and sturdiest supporter. "You were not there, " she said, when she read my published Interpretation; "You do not understand." And I am sure that there is much I still do not understand, and never will. I am encouraged to do the best I can without disciplined skills, by the knowledge that I share substantial biases with professionals such as Jan Sapp. And I was heartened by a letter from Marcus Rhoades, the maize geneticist who moved from Urbana to Bloomington in 1959, just as I was moving from Ann Arbor to Urbana, and Salvador Luria was moving Urbana to MIT. Rhoades shared Tracy’s later years and admired him greatly. After I published T. M. Sonneborn: An Interpretation, Rhoades wrote, "I just hope that when I am gone, someone will be able to view my career with as much sympathy and understanding as you have shown Tracy."

This essay is also in part motivated by a conviction that science is primarily a communal activity and that the focus of most scientific narratives on the heroic achievements of Giants is a distortion if not a misrepresentation of its essential nature. The origin stories of scientific advancement that indoctrinate the young are often reduced to lists of names which are used as mnemonic reference points. In genetics we evoke the names of Darwin, Mendel, Pasteur, Morgan, Beadle and Tatum, Luria and Delbruck, Jacob and Monod, Watson and Crick..., along with pithy phrases that summarize their contributions, but with little context or continuity. I contend that these names are symbols of communities, and that many essential contributions have been made by communities not recognized by an icon.

The modern disciplines of metabiology - that include historians, sociologists and philosophers - are far more sophisticated in their nuanced accounts of the origins and the significance of scientific advance's, but practicing scientists often lack the time and motivation to search beyond the disciplinary myths. And biology textbooks, because of marketing strategy, are loaded with soon to be obsolete details and provide a poorly developed story line. This personal account of a particular period from the perspective of a successful but not distinguished 20th century geneticist just might help put a human face on what must be a very common kind of episode in the perennial struggles of ideas and personalities, that constitute the scientific process.

The Place and Time

My experiences in Bloomington, Indiana in 1946-51 represented an exciting departure from my earlier life. They sharpened my personal values and set my professional career on course. I have tried to understand what was happening during those years ever since. Here in another midwest university town, over half a century later, I still cannot account fully for what happened to me, and around me, but I keep returning to the exercise. I hope the circumstances and emotions, recollected in the tranquility of maturity (senescence, or at least obsolescence), may be of some slight interest to others engaged in the life of science and the history of genetics.

In October of 1946, I celebrated my 21st birthday at home in Wewoka, Oklahoma, and then travelled by train to Bloomington, Indiana. There I began graduate studies in the Zoology Department at Indiana University. Nearly five years later, in June of 1951, I received my Ph.D, and drove with Jean Kelly, my new bride, in the green 1941 Chevrolet sedan which bore well its 130,000 miles and was a wedding gift from her parents, to begin teaching genetics as an assistant professor at the University of Michigan in Ann Arbor. I focus on this particular time and place, with glances forward and backward.

One thing that I have come to realize belatedly is the significance of time and place in understanding events. 1946 was not only a distinctive time in my own personal history, but it was a momentous time in the history of the United States, and particularly in the history of higher education in America. World War II had ended and American troops were drifting home in large numbers from battle grounds in Europe and the Pacific. They were beginning to reconnect with families and friends who had been left behind, but they had been deeply affected by their experiences around the globe. The entire society, as is characteristic of war times, had been exposed to a heavy dose of “shock and awe”. Few of us were able to return comfortably and unquestioningly to familiar roles.

The education that America got through World War II was not written in books and could not be taught in schools. Even before the war ended, President Roosevelt’s last presidential campaign incorporated that changed global outlook and it was verbalized in the campaign for One World, launched by Wendell Wilkie, Roosevelt’s articulate challenger. That vision would fade back into a cold war and a new isolationism, but for a generation humanity was welded together as never before. The defeated enemy, through enlightened programs such as the Marshall Plan and the MacArthur occupation of Japan, were being treated as human beings whose dignity and prosperity were essential for the realization of a global American future. Scholars from over the world were streaming from damaged homelands and were flocking toward the beacon of opportunity in the United States, to seize opportunities and to make catalytic contributions.

The biological and social imperatives of returning veterans and immigrants, became a palpable force in the American economy. For one thing, veterans were ready to start families. In the next few years they enthusiastically generated a "baby boom", which had an enormous impact on the national economy and the system of education. The veterans’ international experiences, their exposure to human diversity and the power of technology, modified traditional values, lifted aspirations and enhanced motives for higher education. Both young men and young women were newly motivated to train for learned professions in a wider world. The "grunts" who had survived horrific battles in the air and on the beaches from Normandy to Iwo Jima, did not come home to accept happily the limited horizons and the menial assignments in the depressed economy they had known as children. The young women who had served ably and responsibly in a male-depleted home society were not content just to sit at home, nurture the men, and raise the babies. Veterans were encouraged and faciliated in their hopes of higher education by a government with remarkable wisdom. They were treated as heroes ready to participate in an explosive American economy and they were equipped by the GI Bill of Rights with resources to seek the professional levels in a new society that they hoped to achieve. An unprecedented fraction of young people enrolled in colleges and professional schools.

The Saharan Butterflies

Before getting too caught up in Bloomington, I need to explain a few things about where I came from, not necessarily because I believe I can identify among the precursors any significant predictors of what happened later. I know better than to claim that particular things were caused by identifiable things that came before, though of course some of those things or some combinations of those things, had some consequences.

I was born in Abingdon, Virginia, a small town in the mountains near the Tennessee border, on October 10, 1925, the second son of Pearl Ledbetter and Thomas Grady Nanney. Both Pearl and Grady had been born in the 1890s in the Blue Ridge Mountains of North Carolina. They came from "Scotch-Irish" farming families that have been traced back, in fact, to Welsh and English immigrants who arrived in Boston in the 1630s, with some spice added by way of genetic contributions from a Chickahominy Indian woman along the way. When I was born, T. Grady was the pastor of the First Baptist Church in Abingdon, but in another year he would move with his family some 1300 miles westward to a distinctly different environment in the new state of Oklahoma.

The auspicious circumstances attending my birth were described to me in a letter written by Pearl, and given to me on my 21st birthday, before I left for Indiana. The letter was colored by my mother’s romantic imagination, and was intended as a propitious augury, and hence was suspicious in its details if not its intent. The first frost was on the Blue Ridge and Grady had built a wood fire in the fireplace in the larger apartment to which the pastor had moved in anticipation of the birth of a second child. The first child, Thomas Jr. had been left with friends in anticipation of an interesting night. The lights were out when the contractions began to come more quickly, and Pearl turned on the lights again to find the room filled with smoke, pouring from the joists beneath the floor under the fireplace. Grady leaped out of bed and grabbed the telephone, uncertain as to whether he should first call the doctor or the fire department. It didn’t seem to matter; the operator did not respond. Grady quickly pulled on a pair of pants and ran to the Model T, intent on going for help in some direction. Never being of a mechanical inclination, however, he managed to flood the carburetor, leaving him standing beside a useless automobile in a hopeless dilemma. According to my mother’s letter, Dad appealed to the Almighty, running down the street screaming "Lordy, Lordy, Lordy. Somebody help me. My wife is having a baby. My house is on fire. My phone won’t work. My car won’t start. Lordy, Lordy, Lordy. Won’t somebody help me!" According to the "offical" account, help materialized, and David appeared without obvious complications at precisely 10 minutes after 10 o’clock, on the 10th day of the 10th month of 1925.

I don’t have any memories of life in Virginia, except as a fly-by visitor in the 30s as my parents went back to the Blue Ridge every summer to keep in tenuous touch with relatives. My memory hoard begins in Oklahoma where my family moved before I was a year old. The circumstances of life in Oklahoma were radically different from those in the mountain valleys of the eastern sea board where my ancestors had dug in for nearly 300 years. Oklahoma was just opening up unexpectedly to "American" settlement, in lands that had been given to displaced indigenous Americans, because their lands in the East were coveted by white settlers, and the lands across the Mississippi were deemed pretty worthless. Indian "removal" from the farms and forests of the Appalachians began well before the Civil War, and was forcibly accomplished along the "Trail of Tears", through Georgia, Alabama, Mississippi,and Arkansas. Many of the Native Americans in the East had already adapted to the "civilized" ways of the whites, taking up farming in the style of the European settlers, including using the labor of black slaves in the South. In the far South many of the Indians and many escaped black slaves had retreated into the untamed lands of the Florida Territory, but most of them were rounded up by federal forces under General Andrew Jackson and marched westward.

Seminole Indians had lived in intimate but ambiguous association with blacks in Florida, and were moved together to the "Seminole Nation" in Oklahoma Territory. Wewoka, in the wooded eastern part of the state, was the capital of that nation. The related Creek Indians were deposited near-by. The "Black Seminoles" developed a settlement down on Little River, where their circumstances were even more strained than in Wewoka. The Creek Indians periodically made raids on the Black Seminoles, capturing them and selling them back as slaves to the southern states. A group of them, determined to find to a safer haven in Mexico, bundled up their few belongings and fled across Texas. Despite harrassment by the Texas Rangers, many of them crossed the Rio Grande and found employment as scouts in the Mexican Army.

This part of the history of Wewoka and the Seminoles was absent from public discourse when my family arrived in 1926. A few Seminoles lived in settlements near the town; a few were in the public schools. A few blacks were recruited from down on the river to be servants in the homes of white families. Virginia Lottie lived in a shack in the backyard of the Nanney home, and took primary responsibility for cleaning and cooking after the third Nanney boy, Daniel, was born in 1928.

The town of Wewoka to which Grady Nanney brought his family was a brand new town. None of the homes or public buildings had been built a decade early. The hanging tree was standing in the courthouse lawn, and the execution tree, where criminals were dispatched by rifle fire in earlier times, still stood in front of the bank. A few blanket Indians lounged about the courthouse. The streets in a newly imposed town grid were given Indian names - Okfuskee, Mekusukee, Ocheesee, Hitchitee - but the business of Wewoka was the exploitation of the oil reservoirs that had been discovered in the near-by Dixie Lease. The luxurious homes built on Bluff View at the edge of town housed the executives of oil companies. High school boys made spending money working on oil pipe lines in the summer. Oil money financed the building of a recreation park on Wewoka Lake, a few miles away, equipped with carousels and other mechanical rides. Wewoka was a prosperous oil boom town.

That Wewoka didn’t last very long, however. The Stock Market Crash came in 1929; bank runs cleared out the funds in the local bank; respected leaders of the community were wiped out, committed suicide, or left town. The Wewoka Lake Resort was closed with chains and locked; it was forgotten and gradually rusted away. But I don’t really remember much about the Crash and the end of the shallow oil reservoirs either. I do remember the Dust Bowl days that came later, in the early 30s, when the rains failed, and the farms dried up and blew away in clouds of grit, and many Oklahoma farmers gave up and moved to California, to tell their stories to John Steinbeck. When I became acutely ill with a bone infection acquired on school grounds in the summer of 1933, a family friend - a mortician - drove me with a high fever and without water - in his funeral van through clouds of dust to the Crippled Children’s Hospital in Oklahoma City, where I could have my infected pelvis bone ("an inch from the rectum") scraped by a specialist. We would wait to see if scraping would do with job, or if maggots would have to be applied - in these days before antibiotics. Fortunately the scraping worked, actually three scrapings, and immobilization in a body cast for nine months.

I am sure that I cannot explain which of my experiences before going to graduate school influenced me in what ways after my arrival, but I am sure that I was affected by having passed formative years in a minister’s household with Southern cultural values and practices, in an oil boom town in Oklahoma that collapsed in the Great Depression and dried up in the Dust Bowl. I am inclined to view as more consequential than most of my circumstances my illness at the age of eight. Anne Roe was the wife of George Gaylord Simpson, the distinguished paleontologist who helped establish the Modern Evolutionary Synthesis in the 1940s. Ann Roe (1953) used her husband’s connections with famous scientists to undertake a biographical analysis of the determinants of scientific careers and to write "The Making of a Scientist". One of her observations was that a significant fraction of her informants reported a memorable separation from their peer groups in early adolescence. In several cases the separation was conditioned by an illness that prevented normal socialization and peer bonding. In other cases the rupture was caused by a traumatic economic change in the family fortunes or a move into an alien environment. Roe speculated that the failure to establish strong peer linkages may have insulated these future scientists from what was later called "group think". The proto-scientists grew up to perceive experiences as outsiders, and were more critical of accepted wisdom. Roe’s study was probably not designed with proper controls, and may not merit serious modern consideration, but she may have identified a kind of early experience that is common among academics.

The peer separation effect may have been mediated by the substitution of a different source of information and stimulation from that normal for the age and place, i.e., the printed word. I grew up in a bookish household. My Dad was to a large extent a self-educated person, a person a bit embarrassed not to have had the stamina to endure the entrainment required to acquire proper credentials for his profession. He couldn’t muster the discipline to learn the Greek that normally constituted a part of the admission costs of the ministry. At another time, in a different context, his failure to complete a college degree, much less seminary training, would have sent him in another direction. But the Great Revival of the Nineteenth Century that surged up the mountain valleys of the Appalachians placed more emphasis on inspiration than on education, and welcomed an untrained clergy rising up in the midst of an uneducated people.

Grady Nanney did not complete his formal education at Wake Forest College, and never attended seminary, but he was more learned than many with framed diplomas on their walls. His approach worked for his time and place. He became a trustee of Oklahoma Baptist University, and was called to be pastor of the University Baptist Church in Shawnee. He wrote three books, one of which was entitled Dear Gene, though he knew nothing of genetics, and was awarded an honorary doctorate from OBU. His office and his home were lined with bookshelves filled with books of considerable diversity. He was a sucker for the itinerant book salesmen trudging through Oklahoma, trying to make a living selling books from house to house, instead of selling apples on street corners as they might have done in an urban setting. We had multivolume encylopedias, sets of histories of the world, the writings of Irvin S. Cobb, of Jack London, of Mark Twain, of O. Henry.......

Pearl Nanney was also a reader, though she - like Grady - had only earned what would be called today an "associate’s degree" from the Round Hill Academy. She was qualified to teach in the public schools of North Carolina, and was in fact involved in "special education" in Raleigh when she was being courted by Grady. She was also involved in a project to bring some educational polish to poorly educated women aspiring to higher culture in Wewoka, a project promoted by an organization called the Delphian Society. The Society distributed printed materials from classical sources for reading and discussion by local groups. She was also familiar with the Wewoka Public Library and, when I had exhausted the shelves at home, and had consumed the books contributed by parishioners, she went to the Library and searched for something suitable for a preadolescent. She learned that a book about paleontology or archaeology would keep me occupied longer than a story book, even though it was over my head.

By the time I got back to school my head was stuffed. And I had acquired a reading habit that needed to be fed. During my illness I processed a book a day, and I continued to read too much to have time for socialization. The continued restraints on active participation in sports reinforced my autonomy. By the time I reached high school I was some kind of classroom menace for an inexperienced teacher. I didn’t understand at the time why Mr. Simmons, the high school principal, moved me out of Clara Zobisch’s History of the World and into Mrs. Porterfield’s Art class. It’s true I told Miss Zobisch that I thought Aristophanes was pronounced AriSTOphanes and not AristoPHANES, and that the French builder of the Panama Canal was not L’Enfans, as she claimed, but DeLesseps. L’Enfans was the one that had designed the city of Washington, D.C. Miss Zobisch kept getting red in the face, and after a week Mr Simmons took me away.

I suppose I really have to say something more about the family in which I was raised. Frank Sulloway’s (1999) book "Born to Rebel" supported my prejudices about the powerful effect of the ecology of the nuclear family in shaping the expression of genetic capabilities in humans. I am sure that my mother’s position in the middle of eleven children shaped her values and priorities, I suspect that the death of my father’s mother when he was a child (at the birth of his sister) engendered in him needs that focussed his gifts and guided his development. I am sure that growing up in a minister’s home in a small town in Oklahoma had effects that blurred any simple genetic determinants of personality. But I can’t very well analyze precisely how growing up in a bubble, identified as "one of the Nanney boys" may have affected my professional career.

Certainly having a mother who was a "professional" pastor’s wife, and who took her position in the community seriously, resulted in more regulated and monitored behavior than might have been good for us. Dad’s need to be loved meant that the burden of discipline and direction fell heavily on our mom. Her well-intended efforts to direct the boys into appropriate professions were perhaps too heavy-handed, and not very successful. She wanted Tom to become a minister like his father. He conformed to the extent of starting to OBU as a ministerial student, with a 4D classification in the Selective Service. He soon dropped this, however, and volunteered for the Marine Corps. He spent the war years in the South Pacific, making numerous beachhead landings, in a unit that suffered 150% casualties. He never recovered fully from the experience. The jungle rot that still eroded his body until his death at the age of 80 was the lesser of the traumas he endured. For the family, sitting at home, listening to the news of the war, hoping day by day to hear from Tom, sometimes for months at a time without a clue, brought a silent but enduring trauma to the home front as well. He was buried with honors at Arlington National Cemetary in 2004. Roger Aldridge, my one classmate in high school who consistently bested me academically, joined the army early in his senior year and died before graduation, leaving me the valedictorian cup. When I came of age I volunteered, but was rejected. When Dan came of age, he volunteered for the Marines like his older brother and was accepted, late enough in the war, fortunately, that he saw no active duty.

Tom’s interests in the Christian ministry had evaporated by the time he got home. He thought maybe he would write, but decided that writing for a living was a hazardous way to escape penury. He finished an English degree at OBU and then decided to build on his home experience as our chief gardener. He went to Texas A & M for a master’s degree in horticulture, thinking maybe to start a nursery to support himself while he wrote incomprehensible poetry. His lack of start-up money for a nursery, combined with his writing skills, brought him to the staff of a nursery trade magazine - The Southern Florist and Nurseryman. From editing that magazine he moved to Iowa to become a garden editor for Better Homes and Gardens for several years. His lack of diplomatic skills and the small-town mentality of Des Moines sent him on with his family to New York where he edited trade magazines before joining the American Petroleum Institute as grants manager and propagandizer.

Pearl decided that with my experience of childhood illness, I should become a doctor. Besides, doctors are honored professionals in the community, and they make good money, not an unimportant consideration with a family that had endured the Depression. My contrariness (or her over organization) resulted in my decision NOT to become a doctor. Dr. Trent, my biology teacher at OBU, insisted that I take the medical aptitude test when I was there, and I scored in the 99th percentile, but I refused to apply to a medical school. What I did instead is recounted further on.

Daniel was the last of the Nanney boys, and was plagued by the diminutive “Danny Nanney” most of his life. Though Sulloway’s study of scientists suggests that later children, in trying to find a comfortable niche, often resort to unconventional tactics, in our family the forerunners were sufficiently rebellious in our own ways to leave an open niche for Dan to assume something of a first-child role. He was the one responsible for taking care of things around the house, for dealing with the parents in their later years, to holding firm to traditional values. He was never quite able to play the “surrogate parent” role because his brothers were too far out to be roped in. Dan had a slow start educationally because of marginal dyslexia, but recovered. He resisted Pearl’s call for him to be a “business man” who could support his parents in their old age. Instead he went to college and became an educator, got married, moved to Colorado as a school teacher, became a school administrator in Littleton, a photographer, and father of three loving children. He is the only one of the Nanney boys who still professes to be a Baptist.

The Nanney boys were never very close while we were growing up. I suspect that my disproportionate share of attention forced both of my brothers to create lives of their own in corners of the bubble. We are probably closer together in our last years than we ever had time to be as children.

I Am Launched

Now, back to my story. My own presence on the Bloomington campus was a consequence of the post-war national crisis in education, particularly because of the shortage of trained teachers. I had been excluded from military service because of childhood illnesses (asthma and osteomyelitis), so after high school I had gone to college. I chose Oklahoma Baptist University in Shawnee, Oklahoma, primarily because it was close to home, and because my father’s profession as a Baptist minister helped provide a partial scholarship in a comfortable environment. As an educational institution during the war, the college was marginal. The student body had shrunken to some 300 students, most of whom were women. The men were mainly ministerial students deferred from military duty by their 4D classifications, and a scattering of 4F individuals, like me. The only way the school could maintain a regular schedule of classes was through a contract to use its facilities and staff to provide basic academic instruction to an ASTP contingent of several hundred young men. The military presence was constantly with us in marching platoons and, in a more relaxed mode, in the campus post exchange, where I was employed sometimes.

I got my initiation in teaching as an assistant in the elementary physics class for ASTP students, which was in fact taught by Johnny Guy Hill, who had been my high school physics teacher back in Wewoka and was a family friend. Later I was asked by Dr. Trent, the biology teacher, to serve as an assistant in his courses in Zoology and Comparative Anatomy. My deployment in science courses did not reflect a particular penchant for science; indeed I didn’t seriously consider a scientific career until other careers seemed blocked. At the time I was a student at OBU, the most popular program for non-ministerial students was that directed by then young Warren Angell, Dean of Fine Arts, lately come from intensive engagement with Fred Waring’s Pennsylvanians. I was devoid of significant talent in music, though I had taken piano lessons for several years as a child, played in the marching band and sang in the Men’s Glee Club at Wewoka High, and in the Bison Glee Club during the years at OBU. I had also played a tuba in the Bison Band when it was reborn in my senior year. I had no hopes or desire for a career in music, nor in the Christian ministry.

My greatest intellectual excitement came from the sparse offerings of courses in literature and philosophy, for which the instruction provided manna in a desert place. Miss Annie Juliette Earle, my prim Victorian philosophy teacher, whetted my ambition to be a scholar, but warned me that a philosopher must first of all be a scientist. My English teachers were great. The quality of instruction was marred chiefly by the narrow concerns of some of the ministerial students. “Professor Snuggs, do you think William Shakespeare was saved"”. A young man named Oral Roberts, who would soon make his mark as a televangelist and as founder of a conservative university in Tulsa, was in one of my English classes. What happens in a classroom is not entirely under the control of the teacher.

As I approached graduation from OBU, I began applying to graduate schools, hoping for a career in the teaching of literature or philosophy. My application packet included a large number of credits in a wide range of courses, but with little depth in anything. My grades ranged from excellent in subjects I liked to miserable in courses in which I rebelled. I quit attending Professor Bailey’s Chemistry course at mid-term, after he demonstrated the Plan of Salvation by pouring a test tube of a red liquid into a flask of black liquid, producing a flocculent white precipitate. With Christian charity he gave me a C for the course, by averaging the A I had before midterm and the F I earned aferwards. I had problems with Dean Solomon in my one psychology course. He ranted about Sigmund Freud’s inappropriate preoccupations, until I pointed out to him that one of the chapters in our text had been reprinted from one of Freud’s books. He simply dismissed the observation and informed the class that they need not read that chapter and would not be held accountable for it. Though I could "explain" my erratic scholastic record, I could not make it attractive to the universities I hoped to enter with a scholarship or assistantship.

As letters denying financial support began to pile up from graduate schools, Dr. Trent, my biology professor, informed me that universities around the country were short of staff to teach preprofessional students, and were offering generous teaching assistantships for qualified students. Very few graduate students had been in training during the war and the bulge of undergraduate veterans was too large to be handled by the few faculty members still in place. Although I hadn’t had much scientific training, I had experience as an undergraduate assistant in a couple of critical courses. Trent assured me that, if I were willing to do graduate work in Zoology, a teaching assistantship could be had, even though I had a spotty academic record from a school with little national visibility.

By this time I had decided that I definitely wanted to become a college professor. I enjoyed reading books and telling other people things they didn’t know. I particularly enjoyed entertaining and irritating the ministerial students in the dormitory with learned discourses on Freudian psychology, Darwinian evolution, and racial discrimination. I decided that going to graduate school in Zoology was better than not going to graduate school at all. So I sent off my applications to two additional universities, universities that I knew little about, except that I had skipped them in my initial application effort. Despite the skimpy list of biology courses I offered (General Zoology, Comparative Anatomy, Spring Flowers of Oklahoma), within a week I was offered a teaching assistantship in Zoology at the University of Iowa. The stipend offered was $600, sufficient - I was assured - to pay for room, board and tuition for year. I immediately responded, accepting what seemed like a good deal. The next week I got a letter from Indiana University, making a similar offer, but raising the ante to $650. Recognizing the hand of Providence in this human’s affairs, I wrote Iowa to explain that circumstances beyond my control made it necessary for me to change my plans. Whether that decision was being realistic or accepting prostitution is subject to debate. I began planning to go to Bloomington in the fall. To refresh my biology background, Dr. Trent had me teach the Zoology course for summer students in Shawnee, while I took a course in geometry.

The Landing Field

Bloomington, Indiana in 1946 was a focal point of a societal revolution - along with dozens of other institutions of higher learning. Universities that had struggled through the Great Depression with low enrollments, failing facilities and poorly paid faculties, were suddenly inundated with hordes of students that they were ill-prepared to house and too poorly staffed to teach. The flood of returning veterans, empowered by veterans’ benefits, and motivated to become doctors, lawyers and engineers was transforming the American system of higher education.

Indiana University was located in a sleepy little town nestled in the limestone hills of southern Indiana, in a quaint local culture that was reflected by the native poet, James Whitcomb Riley, whose popular poems included "Li’l Orphan Annie", "When the frost is on the pumpkin", and "The goblins will get you if you if you don’t watch out". Bloomington’s most notable contribution to popular culture was Hoagy Carmichel, whose presence was still evoked at the Stardust Bar in downtown Bloomington, and who was represented nationally by songs such as "When Smoke Gets in Your Eyes". The hills and valleys were notable for the colors of the hardwood forests in the autumn, for nests of bohemian artists, and for the limestone quarries that became "passion pits" for the undergraduates in the spring. One of the few nationally notable academic specialties was the program in folklore led by Professor Stith Thompson.

The Indiana campus was being transformed by the post-war educational challenge. The pre-war physical facilities consisted primarily of a cluster of small picturesque classroom buildings and dormitories constructed from local limestone, and totally inadequate to meet the new educational demands. The administration responded energetically to the task. A small city of quonset huts mushroomed on one edge of the campus to provide housing for returning veterans with families. A functional complex of concrete block buildings was erected on another side of the campus for single graduate and undergraduate students. A spacious new union building arose in the center of the old campus to serve as the focus for the social activities of the burgeoning student body.

The housing crisis was sufficiently critical to distort the academic calendar in 1946. In order to defer occupation of the student dormitories still under construction, classes began only at the end of October instead of at the beginning of September. As a graduate assistant arriving the evening before I was to report to the Zoology office, I was denied access to the dormitory because the paint wasn’t dry. No sleeping accomodations were available elsewhere in town, so I spent my first night on campus walking around the wooded hills with other early arrivals, watching for falling stars, and napping on campus benches. Even with the delay in starting school, time had not been sufficient to install the partitions inside “Rogers N”, the dormitory for graduate men. Six to eight men were assigned to each of the communal barracks, which would be segmented the following summer into rooms for paired students.

Dominating my sextet was Colonel David Benjamin; three of my other roomates were veterans with diverse backgrounds and extensive worldly experiences. Colonel Benjamin had seniority on the use of "David", so I acquired a new appellation as "Junior", which became my dormitory designation for the next couple of years. Among other things, I remember the pleasant surprise the first morning when starting toward the bathroom in my robe. There was another student in his robe, a black student stopping to drink at the water fountain. There had been no black students at OBU, and no water fountains in Bloomington were marked "For Colored Only". I found the dormitory lounge life particularly fascinating, indeed much more interesting than the undergraduate science courses I was required to take at the University before being qualified for graduate classes. I became something of a bridge addict, and could usually find a foursome willing to play a hand late into the evening. Back at OBU, I was adept at "Shoot the Moon" , played with dominoes, and sometimes referred to as Baptist Bridge. Cards and liquor were not allowed at OBU. The dormitory experience in Bloomington opened windows into the national culture that were not open at Shawnee in a dorm filled with ministerial students.

The demands for augmented instructional facilities and staff at the university were as urgent as those for housing. As a student participant in the frenzied activities I was aware of only a few of the local perturbations. My instructor in Organic Chemistry, for example, was Professor May, who had been recalled from retirement in his 70’s to stagger into the lecture hall and to supervise laboratories for the hundreds of pre-med students who had enrolled. As a teaching assistant in the Comparative Anatomy course, I had to supervise the dissection of phenol-preserved rats, instead of the cats usually used in such courses, and on which I had been trained, because the rats took up less space than cats in the steel bins where they were stored between classes. Classes in the small laboratories had to be scheduled from early morning until late at night, and into the weekends. My first lab partner was Francis Gambell, an attractive young woman from South Carolina, who was as poorly prepared in biology as I was, but who had, like me, been recruited from a pool of inadequate applicants. She was a forerunner of the surge of women into the biological sciences.

Bloomington Dramatis Personae

Though Indiana University was responding to a national challenge to provide augmented educational opportunities, it responded in distinctive ways. Much of its success was attributable to the skills of Herman B. Wells who had just been elevated to the Presidency from his office as Dean of the Business School. Hermy would direct the University response with skill and imagination through the immediate crisis and would steer and personalize the University for the next half century. Wells was a rotund bachelor with remarkable ability in public relations. He was able to pull the campus together, to garner the confidence of the state legislature, and to find the resources to meet the challenges.

Though Wells was not a scholar himself, he was able to choose visionary advisors and to support their visions with adequate funds and public support. Wilfred C. Bain was the ambitious Dean of the Music School, who wanted to develop Bloomington into a Fine Arts Center notable not just in Indiana but throughout the entire midwest. Bain recruited academic musicians from an impoverished and demoralized Europe, and collected professional musicians - particularly aging opera singers- to become artistic coaches. And he was a talent scout for students, providing scholarships to promising young musicians around the country. His brochures reached even to Oklahoma Baptist University, and among several others brought Jean Kelly to Bloomington on a summer scholarship in 1949. Her pianistic abilities caught the attention of Anis Fuleihan, a gifted exponent of a romantic pianistic tradition, and an expatriot from Cyprus. Anis recommended a fellowship which enabled Jean to complete her senior year in Bloomington and to graduate at Indiana. She was then awarded a graduate assistantship and remained an additional year to leave with a 70 minute recital, an MA degree in piano, and a MRS in David Nanney.

Another recipient of Wells’ wholehearted support was Fernandus Payne (1881-1977), Dean of the Graduate College when I arrived in Bloomington, as well as Head of the Zoology Department. He had previously served as Dean of the College of Liberal Arts. Fernandus Payne had a personal connection with important roots of American biological research. He had been a graduate student in Biology at Columbia University at the turn of the century, a student in the biology department distinguished by the leadership of cell biologist Edmund B. Wilson and geneticist Thomas Hunt Morgan. E.B. Wilson and his students consolidated the Cell Doctrine and presided over the integration of Mendelism and cytology into the Chromosome Theory of Inheritance. Morgan founded the Drosophila breeding program and established the fruit fly as the first organism domesticated primarily as a scientific instrument. Morgan received the first Nobel Prize awarded in genetics, in 1937. Payne’s scientific contributions were not so dramatic though, in fact, he had conducted one of the earliest experiments reported with fruit flies; he grew flies for dozens of generations in absolute darkness in a basement at Columbia University to determine whether they would lose their eyes in the absence of light. (They didn’t.) However, he developed techniques for maintaining them in the laboratory and recommended Drosophila to Morgan as experimental animals, in anticipation of the development of the famous "Fly Room".

Payne returned to to his home state of Indiana to teach, but he had abandoned his experimental career during the Depression as funds for faculty and administrative salaries at indiana stretched thin. He was reported to have tossed a coin with Ichthyologist colleague Will Scott over who would have to take over the administrative tasks in biology, and who would be allowed to continue research. Payne lost the toss, but he never lost his faith in the centrality of research programs in the mission of the university. He was a small, bald, nut-brown man with a soft voice, but he made a big personal impression on me. Some time during my second semester he came into my little office in Science Hall, which happened to be next to the Zoology Office, and commanded me to follow him into his private lair. He had just received my class grades from the first semester - near the C-level intolerable for a graduate student - and also the results of my GRE (Graduate Record Examination) scores - which were in the high 90s percentiles. He set me in front of his desk and told me in unequivocal terms to send home those philosophy and poetry books that littered my desk. I must find myself a research sponsor, and give science a chance. Otherwise there was no point in my returning to Bloomington for another year.

Fernandus Payne was a committed talent scout, who considered it his primary function as an administrator to select prime candidates both as faculty and as students. He claimed at one time to have met every zoology graduate student in a significant research laboratory in America before the student finished a thesis. Though that claim seems preposterous given the scale of universities today, it provides a commentary on the sparseness of the research establishment prior to WWII, and the scale of academic research before NSF and NIH began supporting university research.

An example of the commitment of the Indiana administration to disinterested academic research is provided by the story of Alfred C. Kinsey. Kinsey had been a faculty member of the Zoology Department for a number of years when I arrived. He was a popular teacher, and an ardent student of evolution who enthusiastically focussed on the classification of parasitic wasps. As a consequence of his experience as a guest in a class for Marriage Preparation, he acquired a new research interest. He was an open-minded free-thinking person who encouraged students to ask questions about matters that were not commonly covered in biology texts or in classrooms, but were standard fare in dormitories, with the lights turned low. How much variation occurs in the length of the human penis" What proportion of couples have intercourse before marriage" How does the frequency of sexual activity vary with age" What fraction of the American population has had a homosexual experience"

Kinsey’s response was to say, "I don’t know. But I’ll find out". He couldn’t find the answers in books so he developed a research program. He began with the students in his classes, and invented a system of intensive private oral interviews. He believed that his skills were adequate to provoke valid answers even from reluctant respondents, particularly if he could have access to them before they were aware of his objectives. He gradually refined his techniques and expanded his sampled populations to include most sectors of modern society, from prostitutes and prison inmates to corporate board rooms. When I came to Bloomington in 1946, Kinsey was focussing on graduate students, particularly on graduate students in biology.

As I reported for the first time to the Zoology office on the fifth floor of the Science Building, the departmental secretary asked if I had met any of the Zoology faculty. When I answered that I didn’t know any of them or anything about them, she suggested that I should begin to get acquainted. She offered to take me instantly to see Dr. Kinsey, whose office was on the fourth floor just below, beside the elevator. The elevator was an unreliable relic enclosed in a wire cage that was operated by pulling on one of a pair of cables inside the cage. She explained that it was simpler to walk down the stairs. Kinsey’s light was on, but he did not respond to a knock. We returned to the office and I asked which faculty member I should call on next. The secretary responded that it wasn’t necessary to chase down the faculty; other things needed to be done. When I reported the following morning to the office, I asked whether I shouldn’t again try to meet Dr. Kinsey. She answered. "No, that’s not a good idea. He only wants to meet new students the first day they are on campus."

The episode was puzzling until I became better acquainted with some of the more seasoned graduate students. Kinsey was completing his sample of young male students. He believed that anyone who had been on campus more than 24 hours would have learned what he was up to and would have prepared defenses. I was never in fact introduced to Kinsey, though I read his Sexual Behavior in the Human Male which appeared amidst great notoriety two years later, without any input from me. Some of the single males in the dormitory took up a collection to buy a copy, at $6.50 yet, which we wore out passing it back and forth.

Kinsey did discover who I was a few weeks later when I was embarrassed by having to call on him for help in an awkward situation. The comparative anatomy lab was on the fifth floor of Science Hall, and Francis Gambell and I were cleaning up the lab from a late shift one dark Friday evening, wrapping the preserved rats and sharks in cheese cloth and stacking them in steel bins. When we went to the elevator to go home to our respective dorms, I was overtaken by an impulse to pull the "up" cable in the cage instead of the "down" cable. The elevator responded appropriately and carried us to the attic, a dark dusty storage space with no lights at all. My curiosity satisfied, I pulled the down cable, repeatedly, but the lift would not budge. The building seemed to be deserted, and cell phones had not been invented. All we could do was to yell ever more hoarsely down the elevator shaft to a deserted building. Seemingly hours later, but probably no later than 9 o’clock, we got a response. Doc Kinsey, who kept unconventional hours, had returned to his office to investigate some esoteric topic, and heard our cries for help. He called the maintenance crew, who brought a ladder and extracted us through a trap door in the ceiling. He leaned against the wall, watching us with what we interpreted as a sneer, but never said a word to us. He probably had too much information on the shenanigans of biology graduate students by this time to be very interested in what might have been going on the dark attic of Science Hall. In later years when my advisor’s laboratory and the Sex Institute were both moved to Wylie Hall, Kinsey would occasionally drop into the lab during his nocturnal meandering. But I never felt very comfortable with him.

Kinsey undertook an investigation into aspects of human behavior that at the time was considered by many to be out of bounds for public discussion, and unsuitable for academic inquiry or classroom discussion. Before many years had passed, however, the information was seen as relevant for responses to a changing society and to a major crisis in public health. The book was a best seller, though not at the time because of its public health significance. Kinsey’s project met with both applause and disapproval. The administration at Indiana University was besieged with complaints and threats, but the University under Well’s skillful leadership opposed the public sentiment in a conservative state. Wells insisted that the topic and the methodology were suitable for academic research, and refused to terminate or censure the studies. Kinsey subsequently established an institute for research in human sexuality, published a series of significant books, and assembled archives of pornographic films, books and art objects - reputed to be matched in size and quality only by the great collections in the Vatican. His rare open-houses generated a near stampede on campus.

The person who became my mentor, and the most significant person in my scientific life, was not one that I was even aware of at first. Tracy Morton Sonneborn (1904-1980) was also something of a sexologist, but he got his kicks looking in a microscope instead of interviewing naive human informants. His scientific research was focussed on Paramecium, a unicellular ciliated animal commonly found in pond water. Fernandus Payne had discovered Sonneborn in his scouting surveys, while he was working as a long term research associate in the laboratory of Herbert Spencer Jennings (1868-1947) at Johns Hopkins, in Baltimore.

Jennings was an important figure in the development of genetics in America, though his contributions have escaped the myth-making dominating the disciplinary accounts. Jennings discovered experimental biology when he spent time in Europe in the 1890s, particularly at Jena and Naples. In the process he also described the behavior of Paramecium under experimentally defined conditions. His mechanistic description of the "avoiding reaction", evoked when paramecia encounter an obstacle or a toxic substance, adopted an explanation that rejected the vitalism common at the time. Returning to America, he joined forces with people such as Jacques Loeb and C.O. Whitman in supporting experimental biology, research stations, and physico-chemical interpretations of biological phenomena. He enthusiastically welcomed the rediscovery of Mendel, contributed to theoretical population genetics, and made beginnings in microbial genetics with protozoa. He was nearly alone, however, except for a few other independent biologists, such as L.C. Dunn the mouse geneticist, in resisting publically the allure of genetic determinism as a solution to societal problems (Kevles,1985). The American Eugenics movement had flared in the first quarter of the century, fueled by industrial barons and elite educators, who underwrote the development of genetics laboratories and research programs in human heredity. The movement faded only when its racist premises became apparent in Hitler’s eugenic programs of the ‘30s.

Instead of acknowledging the aberrations of the eugenics movement, American geneticists organized a new genetics society (The Genetics Society of America), wrote a new genetics textbook (Sinnott and Dunn - The Principles of Genetics), and created a bowlderized history that papered over the shameful record, along with Jennings. I learned of this potted history only much later. The older genetic society continued through private sponsorship with a much lower profile. Thus, two professional organizations for geneticists were active when I finished my Ph.D. Both published journals describing genetic research, and I joined both. The American Genetic Association (AGA) published The Journal of Heredity and the Genetics Society of American (GSA) published Genetics. The Journal of Heredity in later years was devoted largely to practical applications such as the breeding of plants and animals, while Genetics was the major outlet of experimental laboratory studies. Some 20 years later I was nominated to be the President of the Genetic Society of America, but lost the election to Charles Yanofsky, a prominent Neurospora geneticist. After another decade or so I became president of the American Genetics Association, and only then began seriously to query the origins and significance of the two American genetics societies. What I discovered was a shock to me, and I discussed that history in my presidential address (Nanney, 1983).

The final chapter in my involvement with professional genetics societies came in 1985. The annual meeting of the GSA was scheduled on the campus of the University of Illinois, in Urbana, and I was asked to chair the local Program Committee for the meeting. The committee’s obligations included the recommendation of a number of symposia on topics of wide interest and with popular speakers. The list of symposia was to be submitted personally to the governing board of the GSA at a meeting in Boston MA. The topic I was most enthusiastic about - though I had some difficulty with my local committee - was ?Genetics - A Century of a Science in Society?. The subject matter was to include Francis Galton, the cousin of Darwin who founded the Eugenic Movement, even before the rediscovery of Mendel, and included the American eugenics movement that flourished before the Nazi applications brought it into disrepute. The program I proposed was intended to include the eclipse of the AGA as the major professional society of genetics and its replacement by the GSA. I suggested speakers who were historians and philosophers, and I had approached the current president of the AGA for joint sponsorship of this one symposium.

At the meeting of the governing board in Boston, I met a stone wall. None of the board members had ever heard of the American eugenics movement, or of the related conflicts over involuntary sterilization, and racial biases in immigration policy. One of the board members suggested that the story was a fabrication. They adamantly opposed enlisting historians and philosophers to talk about genetics. Their final judgement was that, even if my summary were historically correct, the eugenics movement should not be discussed in a public forum. Its disclosure could only damage the profession. The board disapproved this symposium topic. And I resigned from the program committee.

Tracey Sonneborn and David L. Nanney

Tracy Sonneborn (left); David L. Nanney (right)

The career of T. M. Sonneborn is an interesting study in its circumstantial trajectory. I assume that his family came to America in the later 1800s among the escapees of Russian pogroms. They, along with the Irish-Catholics fleeing from the Potato Famine, were sharing the difficulties of assimilation into the predominantly English-Protestant establishment. After an early predilection for a career as a rabbi, Sonneborn became fascinated with experimental biology, and completed a Ph.D. with Jennings, in 1928, just before the Stockmarket Crash and the onset of the Great Depression. Sonneborn’s economic situation was especially precarious because of still unresolved questions about the assimilation of Jews in American society. Even a decade later, when Jennings would have liked to ease Sonneborn into the chair he was about to vacate at Hopkins, he told Sonneborn that the "The time has not yet come when a Jew can be offered a professorship at Hopkins."

Fortunately, Jennings had obtained from private sources a long term grant for the support of genetic studies on Paramecium, and these funds allowed Sonneborn to continue his research career, if he was willing to study paramecia. Ironically, the funds came from the Rockefeller Foundation, an agency whose interests in genetics originated in its founders’ approval of the American eugenics program. Yet the funds supported the laboratory of the chief American opponent of the eugenics movement, and they provided the resources to nurture the development of an Ashkenazi Jew, at a time American higher education was significantly biased against Jews.

Even though Sonneborn continued to work through tight economic times, the years at Hopkins were not easy financially, with no raises, few job prospects, and a growing family. But Sonneborn was doing what he enjoyed, he persisted, and finally, after eight years with little apparent success, he discovered mating types in paramecia (Sonneborn, 1937). This discovery allowed him to conduct controlled mating experiments on the transmission of several interesting cellular characteristics. His break-through was near the beginning of microbial genetics, the next major epoch in the history of genetics. Controlled matings of fungi, of algae, of bacteria and even viruses were to be achieved within the next decade, and Sonneborn leapt from being a research associate without tenure to the leader of a significant new technology. Fernandus Payne was soon on his doorstep with an offer of a tenured professorship at Indiana. Payne treasured Sonneborn’s broad knowledge and his wise counsel. He followed Sonneborn’s suggestions for the development of a world-class biology department in Bloomington.

Sonneborn’s own research developed in a direction that brought him into an unexpected role as challenger to the orthodox genetic dogma that was rapidly being consolidated in the 1940s. One of the chief founders of that dogma and its most eloquent exponent was Herman Joseph (Joe) Muller (1890-1967; See Carlson, 1981), who paradoxically had been recruited at Sonneborn’s suggestion to come to Bloomington. My first social event after my arrival in Bloomington was a departmental banquet honoring Zoology Professor Muller, who had just returned from Stockholm. Muller’s Nobel Prize recognized his demonstration of the mutagenic effects of ionizing radiation in the fruit fly, Drosophila melanogaster. This was the second Nobel Prize awarded in genetics, the first going to Muller’s mentor, Thomas Hunt Morgan.

Muller was a participant in the fly-room at Columbia, a brilliant experimentalist as well as a social activist. Muller’s course in "Mutation and the Gene", which I took during my second year at Indiana, convinced me that genetics was the biological discipline most attractive to me. The breadth of his synthesis of biology, built around his central concept of the gene, gave coherence to the entire field of biology. Although the premature and biased claims of the eugenics movement had been rejected or repressed by most geneticists, Muller was still dedicated to improving human health and welfare through genetic science and biological engineering. He was a proponent of sperm banks, supplied by superior men (such as Nobel Laureates), whose genetic heritage could become a public resource. His concern over the deterioration of the human genetic endowment through physical and chemical abuse, led him to public attacks against atomic bombs and bomb testing, and against the irresponsible use of diagnostic and therapeutic x-rays by medical doctors.

Like Sonneborn, Muller was of Jewish background. He was also sensitive to having grown up in poverty in New York. During his course on mutation in the late 40s, he still expressed bitterness over having had to compete in the laboratory with A. H. Sturtevant, whose patrician background allowed him to give his full attention to research. Muller had to interrupt his studies to teach courses at Cornell Medical to make some money for tuition. Even after winning the Nobel Prize, Muller was bringing scraps of dated paper to class, and laying claims to ideas from the fly lab that were commonly attributed to Sturtevant or to Calvin Bridges, the other members of the triad that with him had established with convincing logic and brilliant experiments the location of genes on chromosomes.

Muller’s course in Evolution, which was not offered until I was ready to leave Bloomington, was in some ways the most impressive course I ever had. His comprehensive theoretical treatment began with the origin and evolution of the universe, as it was understood at the time. It then focused on the chemical and physical conditions prevailing on the nascent Earth (before the Urey-Miller experiments), proceeded through the origin of life as realized in the appearance of the primordial gene. He argued that the gene was nucleic acid; it was the Master Molecule, the dominant feature of living systems, all else of which is derivative. he considered viruses to be "naked genes". He paid scant attention to the Modern Evolutionary Synthesis which had been essentially completed by 1950, and which he considered a secondary matter.

One would think that students would flock to Muller’s laboratory, but this was not so. Muller had had an extraordinary career. He taught at Rice University in Texas for a time, but he was restless and sought a broader social arena. He was taken up with the theoretical promise of a socialist society and went to Russia in the 30s where he believed his science would be properly appreciated and implemented. He realized belatedly the merits of democracy when the charlatan breeder Trofim Denisovich Lysenko gained the support of Stalin and suppressed modern genetics as bourgoise, materialist nonsense. Muller escaped from the genetic purge in the USSR to a haven with friends in Edinburgh, and then to Amherst, before getting the call to Bloomington. Though Muller came to Indiana to a steady job and high recognition, he was not warmly received as a teacher or colleague. His classroom lectures were brilliant monologues, but he did not welcome questions, and many students claimed to be confused. He had sufficient research funds from various private agencies to run enormous experiments in well equipped laboratories. But very few graduate students were incorporated into his research family. Some of those who did join as graduate students were neglected, unable to speak to the chief for weeks at a time. A group of zoology graduate students with other sponsors became sufficiently incensed at Muller’s disregard of his students that we went together to the department head to complain.

Nucleocytoplasmic Interactions

I don’t know a lot about the relationship of Muller to the other faculty members, but I was given occasional glimpses in later years of his interactions with Tracy Sonneborn. Quite aside from possible personality conflicts, one has to take into consideration the trajectory of Sonneborn’s research program. Nearly all of his studies (Sonneborn, 1947) using controlled matings with paramecia yielded results that seemed to contradict the universality of the Mendelian modes usually revealed in higher organisms, and contrary to the dogma of nuclear (genic) preeminence in the cellular economy.

Sonneborn’s first crosses probed the genetic basis for the mating type differences that allowed him to control matings in the strains then referred to as "variety 4" and later called Paramecium tetraurelia. When mating type VII cells are mixed under appropriate conditions with mating type VIII cells, they exchange male pronuclei and establish chromosomally identical zygote nuclei in each of the conjugating mates. The conjugants separate, dismantle their old nuclear apparatus (the polygenomic macronucleus) and construct a new macronucleus. Under conventional mendelian expectations the genetically identical exconjugants would be expected to give rise to clones of identical hereditary characteristics. Yet, the type VII exconjugant regularly produced a clone of type VII, and the type VIII exconjugant gave a clone of type VIII. Occasionally exconjugants failed to separate at the normal time and were held together by an open cytoplasmic bridge. Under these circumstances both exconjugants usually gave rise to clones of mating type VIII. The clear implication of the experiments was that mating types were influenced by some cytoplasmic mechanism, and were not controlled directly by mendelian genes.

A second set of studies was triggered by the observation that when stocks of variety 4 from different origins were mixed, those from one source were killed by some toxic substance produced by the other. Killers and sensitives could be induced to conjugate without unilateral death under special conditions, and the mode of hereditary transmission could be assessed. Curiously, the exconjugant killer exconjugant produced a killer clone and the sensitive exconjugant gave a sensitive clone. Again, when cytoplasmic bridges were observed (or induced), both exconjugant clones were found to be killers. The killer-sensitive distinction involved the cytoplasm in some unexplained way; the nucleus did not seem to be in direct control of this trait, even though a nuclear gene, K, was essential for its manifestation.

Sonneborn began a third set of breeding studies, in this case involving responses to antiserum induced by injecting paramecia into rabbits. Strain A cells were immobilized in dilute antiserum prepared against it, but strain B cells would not. Strain B, but not strain A cells, were immobilized by anti-B serum. When strain A and strain B were crossed under appropriate conditions, the strain A exconjugant produced a type A clone, and the type B exconjugant gave a type B clone. When cytoplasmic exchange was induced during conjugation, the exconjugant clones were usually identical, both A or both B, depending on environmental conditions.

The results of these three studies were well established and widely publicized by 1947, particularly in a well-known summary in the first volume of Advances in Genetics. That presention was followed by a series of brilliant invited lectures to various learned societies. Sonneborn himself was lionized. The life cycle of Paramecium was diagrammed in every genetics textbook; the killer character was discussed in every genetics course. Sonneborn was elected a Fellow of the National Academy of Sciences, of the American Academy of Arts and Sciences, and of the American Philosophical Society. He served as president of several professional societies including the Genetics Society of America, the Society of American Naturalists, and the American Institute of Biological Sciences. Times had obviously changed when this "permanent post doc" from Hopkins was offered a professorship at Harvard. I only know of that offer because Sonneborn asked me, some time in 1950 I believe, if I would rather have a degree from Indiana or from Harvard.

One of the first messages from microbial genetics, and a brilliant exploitation of a unique experimental system, was that heredity at the level of the cell does not follow simple mendelian rules, but is under the control of unexplained cytoplasmic mechanisms of a very different nature. The Paramecium studies challenged the hegemony of the nucleus, and in particular the genic centrality that Muller had built into his theoretical masterpiece. And this was occurring in Muller’s own back yard, in the laboratory of the person who was most responsible for bringing Muller back into a respected position in the United States after his fugue to Communist Russia. The situation was ripe for conflict, and could have become explosive.

Muller, from what has been mentioned before about his personality, can be seen as a brilliant but insecure scientist, focussed on his own career and showing little tolerance of dissent. His speech at the Zoology Department banquet in 1946, when he returned from accepting the Nobel Prize, was an eloquent evocation of scientific idealism. Prizes, he claimed, "are not the objective of scientific careers, but are accidental byproducts, welcome but irrelevant." But he was clearly abrasive with his colleagues, and intemperate in conflict. Sonneborn eventually had to exclude Muller from visits to his laboratory because of their frequency and their contentious nature. Muller felt obliged to block the message coming from the Paramecium laboratory, but he had to do it vicariously, so as not to bring an open breach between colleagues.

One episode in the Muller-Sonneborn conflict played itself out while I was in Bloomington. John R. Preer Jr. was Sonneborn’s first graduate student at Indiana. His graduate work focussed in the physical basis of the killer character. The Sonneborn lab at this time was determined to resolve the apparent conflict between nuclear and cytoplasmic inheritance. Sonneborn had showed, for example, that mendelian genes are responsible for mating type potentialities, particularly the differences between one-type and two-type stocks of variety 1 (P. primaurelia.) And he demonstrated that the capacity for becoming a killer cell was associated with a nuclear gene, K, even though the killer trait could be spread by infection under certain conditions to appropriate strains. Preer attempted to identify the nature of the infectious agent; leaning on Muller’s expertise with x-ray technology and target theory, he inactivated the particles responsible for killer infectivity with x-rays and calculated that infectivity was associated with large particles, of the approximate size of bacteria, and that they must exist in killer cells in numbers up to several hundreds (Preer, 1948). Indeed bacterial stains soon showed visible particles (kappa) in killer cytoplasm that was stained with DNA-specific Feulgen dyes.

Muller was delighted. Kappa particles were bacterial symbionts. He was glad to accept parasites as the explanation for cytoplasmic inheritance, and to dismiss all the cases as similarly trivial deviations of transmission patterns - not fundamental challenges to the supremacy of the gene. He did not go public with this interpretation, however. Instead he apparently invited an old friend from the fly room, Edgar Altenburg, to come to Indiana for a visit, and to spend time with Sonneborn, learning the emerging details of the kappa story. Sonneborn welcomed Altenburg to his lab, and freely told him about current developments in the killer story. Altenberg, who had written a popular but conventional genetics text popular at the time, expressed no reservations about the significance of his cytoplasmic stories during his visit. However, he quickly prepared a review of the killer story, and published it in The American Naturalist. It was a skillful polemic that developed the parallels between the emerging studies of plastid inheritance in plants and symbiotic algae and bacteria in ciliated protozoa. Altenburg pressed the dismissal of cytoplasmic inheritance as misleading relics of infections that were of no theoretical relevance in terms of genetic fundamentals (Altenburg, 1946, 1948). Sonneborn believed that Altenberg had pretended to be interested and supportive in order to ferret out details on the visit that might be used to discredit the Paramecium work. The already strained relations between Muller and Sonneborn were damaged even further.

This little episode in the Paramecium tale is linked to other episodes in the emerging exploration of the microbial world. The distinction between infection and heredity was seized upon by Joshua Lederberg (1952), the enfant terrible of bacterial genetics, who had first demonstrated genetic recombination and visible conjugation in Escherichia coli. The more flexible genetic systems of microbes allowed genetic transfer of parts of genomes under special circumstances and revealed phenomena that required redescription of the essential properties of the genetic apparatus and blurred the very words used to describe genetic relationships. Were inducible phages part of the genome of the bacteria that harbored them, or were they parasites" The issue of symbioticism in both heredity and development did not go away. Indeed, Lynn Margulis (1981) eventually made the episodic fusion of alien genomes the foundation of her vision of saltatory evolution, and of her career.

The issues with respect to the Sonneborn laboratory, however, were not resolved by a rationalization of the killer story. Sonneborn was unwilling to sweep the mating type and serotype phenomena under the same rug as the killer phenomena, and to dismiss them as irrelevant in the discussion of cellular governance; what is the role of the cytoplasm in the exploitation of multiple capabilities in the heterogeneous microbial environment" What is the role of the cytoplasm in the acquisition of cellular specificities in the development of metazoa"

Why was Muller hostile to the efforts of the Sonneborn lab" One of the challenges of being pushed into the spotlight of public attention is to find an appropriate encore. Although Muller again had a big laboratory, star status, and adequate funding during his years at Indiana, I’m not aware of any major discoveries that he made in those days. Part of the problem had to do with his instrument. Drosophila was not the useful tool that it had once been, and that it would be again. Drosophila was at the center of experimental genetics in the 1910s, as each of its chromosomes was marked with genes, and recombination was demonstrated to be the consequence of physical exchanges of chromosome parts between homologues. But once its cytogenetic contributions had been consolidated in the 20s, it became less interesting as a research object. It is true that Dobzhansky and other population geneticists by the 1930s had brought new uses to Drosophila in the union of genetics and evolution that was called the evolutionary synthesis. The Drosophila population studies, however, had run their course pretty much by 1950, and Muller had little active role in them anyway. The demand for Drosophila geneticists had slumped; universities were now recruiting bacterial and virus geneticists. Drosophila genetics was a depressed industry until after molecular genetics had matured, and the homeotic mutants of Ed Lewis again resurrected what had become a morbid organismic technology. Muller was married to Drosophila, but the fruit fly had no suitable flying arena in the 1940s.

Muller may have been displaying some organismic envy as microbial genetics began to move into the spot light. At this time Seymour Benzer was doing remarkable genetic tricks with T4 bacteriophage at Purdue, over in West Layfayette, Indiana, before he moved to California to begin his innovative studies in behavior genetics. He was invited to Bloomington sometime about 1950 to give a talk. In his lecture he described inversions, deletions, transpositions, and other genetic rearrangements in the virus, and then commented: "All these things were first described in Drosophila, as Dr. Muller will be the first to tell you". The comment brought chuckles from an audience well aware of Muller’s sensitivity about priorities. But the work of Muller’s that Benzer referred to was classic work, over a quarter century old, and was of only marginal significance to Benzer’s dissection of the gene.

As a graduate student I was not ordinarily privy to relationships behind the faculty facade, but I suspected that external events sometimes had cryptic roots. Julian Huxley, who popularized the modern evolutionary synthesis, was a close friend of Muller. He once described Muller as the "greatest all-round biologist the 20th century has produced". Huxley, as the grandson of Darwin’s bulldog, and as a master synthesizer, with a British accent yet, enjoyed enormous international prestige. During this time he was brought to Bloomington by Muller as a visiting professor, and gave a series of lectures, which he subsequently published as a book, Evolution in Action (1953). It turned out that American college students were not particularly impressed with his style, and graded him a terrible lecturer. The chemistry lecture hall was packed for his first lecture, but after a few sessions only the students who had signed for credit were found clumped in the front rows of a near empty hall. Huxley’s lectures never got around to dealing with cytoplasmic inheritance or the mechanisms of development; they seemed designed to stroke Muller’s sensibilities.

Another notable faculty member at Indiana in the late 40s was Salvador E. Luria (1912-1991), a refugee from Italy, who was laying the foundations of bacteriophage genetics with the German expatriot, Max Delbruck (1906-1981). Delbruck had been trained as a physicist but was attracted to the challenge of explaining the properties of living systems, using more sophisticated analysis and simpler biological systems. Historians have noted the influx of physicists into biology in the 40s and attribute much to the stimulation of Erwin Schroedinger’s What is Life (1944). Whether the book was a cause, a symptom, or an excuse, scientific interest was shifting away from the technology that permitted the atomic bomb and toward the technology that would make life more comprehensible. Physicists made major contributions to the development of molecular biology at this time, and Delbruck was the intellectual focus of much of this integration. Luria and Delbruck worked together to establish a quantitative methodology for bacterial and viral genetics, using bacteriophage mutants, visualized in plaque morphology and host range restrictions, to study mutation and recombination in the simplest forms of life known.They argued for an essential similarity in the mechanisms of replication and of genetic change in viruses and higher organisms, and hence for the appropriateness of extrapolation from the simple to the complex. Microbial genetics opened the door to molecular biology and promised to lay bare the essence of life in its most primitive aspects, and thus to fullfil Muller’s characterization of viruses as "naked genes".

Delbruck, who had been at Vanderbilt for awhile, moved to Cal Tech where he finished his career, while Luria came to Bloomington. Luria didn’t stay at Bloomington much longer, but moved on to Champaign-Urbana, Illinois, in 1950. The shortage of trained biologists that brought me into graduate study, and that provided professorships in American universities for European professionals, was also affecting the economic status of the professions, and encouraging faculty mobility. In order to recruit a faculty member in an emerging discipline, universities were required to, and were willing, to offer salaries and facilities better than they could provide to all the faculty members who were already in place. The result was that major improvements in benefits could usually come to an individual only with a move. Most faculty members moved at least once after their first appointment, and some moved several times. Luria moved from Bloomington to Urbana, taking with him another Indiana faculty member - I.C. Gunsalus - when the University of Illinois undertook a major development in molecular genetics in the 1950s. Luria didn’t stay very long at Illinois, either, but moved to MIT in 1959, the year I moved to Urbana from Ann Arbor, Michigan. This frequent movement of staff between universities, effected by what was called the "Star System", made biological research much more homogeneous across the nation, but it undercut local traditions and loyalties. The demographic forces also placed priority on national visibility in research, and deflected universities from their committment to education. The economic status of the research professions leaped ahead, but at a cost to local communities of scholars and to educational programs.

Though Luria and Delbruck were separated institutionally, that seemed to make little difference. They presided over a coherent and rapidly growing international "phage group" that frequently convened at various places. Luria was a careful and patient experimentalist, capable of conducting tedious and lengthy experiments. Delbruck was the critic of the group who insisted on theoretical rigor. They were to share the Nobel Prize for their work (with A.D. Hershey) in 1969. I don’t know much about Luria’s relationship to Muller, Sonneborn, or other faculty members at Indiana. The membrane between graduate students and faculty was almost impervious in those days. Sonneborn, at least, was very reluctant to discuss any administrative matters or faculty relationships with his students. Only in later years did he occasional let slip observations about relationships that were not entirely as they seemed on the surface. The "information" I received from the graduate student rumor mill was that Luria was an autocratic director, that his left-wing sympathies made his collegial relationships suspect, and that he had a bad case of Nobelitis.

Though Luria served on my doctoral committee, I don’t remember that he had any input on my thesis or comments about it later. My impression was that relationships among the faculty members were so sensitive as to require a "hands off" policy with respect to their students. As in the case with Muller, I don’t know that Luria ever even read my thesis. Only one time did I go to Muller with a problem about my thesis. I was curious about the basis of "selfer" clones, and was considering an explanation based on the properties of heterogeneous compound macronuclei. I wanted to estimate the number of assorting units in an "unstable" mixed nuclear system in paramecium by measuring the rate of production of pure units at cell division. Muller brushed the question off as incalculable. He said that he had himself tried to deal with an equivalent problem back in the 20s and had consulted Sewall Wright about its solution on a long driving trip. Wright told him it c ouldn’t be solved. It was a problem that was eventually resolved by simulation on an early computer at Michigan in a collaboration with physicist and a mathematician (Allen and Nanney, Schensted, 1958) , and a different protozoan - Tetrahymena.

Luria, like Muller, did not attract many graduate students. He had an active laboratory that included several workers scheduled for distinction, such as Renato Dulbecco, who later also received a Nobel Prize for his work on the genetics of cancer. His most famous student was James Dewey Watson, who arrived on campus a year after I came to Indiana. Though when he applied for admission, Jim had declared his intention to study ornithology, he quickly sensed that the excitement and the prizes were in the explosion occurring in genetics. He registered in the courses of the BIg Three, and considered working with Sonneborn or Muller. For at least two years he regularly attended the Friday evening seminars held at Sonneborn’s home, which were the spiritual center of the Sonneborn research group. Jim eventually chose to do his thesis with Luria. As he explained, one has to have an "instinct for the jugular", and he saw in the Luria-Delbruck program the opportunities that would provide him more rapid advances than through the murky phenomenal waters of cytoplasmic inheritance.

Left to right: David L. Nanney, Dick Siegel, Jim Watson and P.K. Cho

Before that, however, while we were being vastly stimulated by the faculty interactions at the weekly Biology seminars, Jim expressed an interest in fooling the pompous authorities. He suggested that together we might pretend to be doing some secret research, and after a period of time might announce that we would like to speak at the Biology seminar. We never discussed what the hoaky seminar might be about, but as a cover story we decided to do some radiation studies. We inactivated the killing activity of breis of killer cells, assayed by exposing sensitive cells to diluted breis and counting the killed cells. We ran the x-ray exposures late at night when most people were leaving the lab, and I would count the killed cells by removing the blistered rounded corpses with a micropipette, over the next few days. We wanted people to know we were running an experiment, but we didn’t want them to know what the experiment was about. Jim wanted ample room to construct the hoax from whole cloth. But then Luria sent Watson to Cal Tech or Cold Spring Harbor for polishing with others in the phage group, and Jim never got around to preparing his announcement designed to astonish and amaze, at least not until he got his hands on another and more substantial data set.

I was left with a set of data on the inactivation of the toxin (paramecin) associated with the killer particle kappa. The killing action was much harder to inactivate than was the replicative ability of the kappa particle. I concluded that only a fraction of the killer particle was necessary for the killing action, and stuck the data in my thesis, along with my mating type and serotype studies, when I submitted it. I didn’t consider that the work was very important, and apparently Sonneborn didn’t either. A few years later, however, I got a letter from Jim, after he had published the double helix with Francis Crick (1953). He had come upon his copy of the data on paramecin inactivation, and wondered if I had ever published the study. He offered to write a paper about it if I had no intentions of doing so. Well, I had written a paper about the work and submitted it in my thesis, but I had never sent it out to be reviewed for publication. I was flattered that Jim thought the work definitive enough to be published, but I wasn’t wild about his writing a paper about it. So I submitted the paper to Physiological Zoology, whose reviewers complained about "half-finished pieces of work that graduate students sometimes submit" , but they agreed to publish it (1954). I added an acknowledgement to "J. D. Watson who ran the x-ray machine for the experiments".

I am not aware that Luria, any more than Muller, was responsible for any major scientific advances after he came to Bloomington. Nobel Prizes are usually awarded for work done much earlier, for work that has already been acclaimed in terms of prestige and economic status. The scramble of univerisities for "stars" who have received or might be candidates for prizes and other marks of public recognition is often disappointing in terms of continued scientific or educatiional contributions. I don’t know if prizes stifle scientific productivity, by making scientists overly self-conscious, or whether biologists - like mathematicians -are often able to make serious scientific contributions only while they are young and hungry. Their more mature contributions tend to be in scientific administration.

Though Luria was technically in the Bacteriology Department, the biology faculty in Bloomington was effectively one faculty at the graduate level. Particularly, all the students interested in genetics took Luria’s Virus course, Muller’s course in Mutation and the Gene, and Sonneborn’s Genetics of Microorganisms. A more contrasting set of courses could scarcely be imagined. They differed not only in their organismic focus but much more strikingly in their approaches to the subject. Sonneborn’s was the least predictable in content, but completely consistent in style. Every year he would choose a different organism as the basis of his presentations, Neurospora, or Chlamydomonas, Saccharomyces or Paramecium. Sonneborn’s course was the most "bottoms up" science course I ever took. It always began with a description of the organism, and of the phenomena that were manifested, evaluated and quantified. He tried to read every paper ever written describing the phenomena and their analysis. He would sometimes find errors in published tables, mistakes in calculations. He would raise queries about biochemical inferences, pose alternative interpretations, and in some cases correspond with the authors of those papers. He would provide a critique of a field of research seldom provided outside the small group of experimentalists who had done the original work. He may well have known more about the status of "Genetics of Microorganisms" than anyone else in the world at the time.

Muller’s course was exclusively top-down. He painted with a broad brush, selected telling details to support his interpretations, but always from secondary sources. He may have read as much as Sonneborn, and as widely, but with very different intent. I suspect that Muller’s critique of the data from his own laboratory was as critical as Sonneborn’s, but his use of the published data of others was much more instrumental, and selected to make the case he was in the pulpit declaiming. I don ‘t believe that Muller’s style was constrained by his organismic commitment, though students would joke (behind his back) that Muller had come to look like a fruit fly, particularly when he was describing mutants with unusual spots on their wing or thorax, and would point to analogous points on his own anatomy.

When I later met A.H. Sturtevant at Cal Tech, I was surprised to see how much temperamental differentiation was possible among the participants in the fabled fly lab. Sturtevant did most of his experiments in his head, sitting with his feet on his desk and his pipe in his mouth. He didn’t want technicians cleaning his glassware and making noise. He didn’t need that many bottles or petrie dishes anyway. He ran small well-designed experiments looking for particular theoretical consequences. Muller ran a fly-factory with large laboratories and many assistants. He was so eager for more data to crunch that he recruited the students in his lab course to run parts of his big experiments, though not until the end of the course were we ever allowed to understand the design or the purpose of the experiment. And he didn’t trust us to make the discriminating diagnoses of the multiply marked phenotypes of the flies. When critical readings were called for, his assistants came in to verify our classifications and counts.

I have a harder time describing and evaluating Luria’s course on "Viruses", perhaps because I took only the one course from him. He spent a lot of time in class on semantics. I particularly remember the long discussion of whether a virus is "alive", attended with an expansion on the essential characteristics of "life". I don’t remember lots of time spent on discussions of experimental protocols or analysis of data. In retrospect I have difficulty relating my memories of Luria’s course to his celebrated achievements in establishing the operational basis of bacteriophage genetics. At the time I had never met Max Delbruck, or even heard that much about him. I had little inkling of his role in phage genetics.

Delbruck came to Indiana and gave a lecture in 1951. (I had stayed on for an extra year after completing my degree requirements, because I had a fellowship, and a girlfriend, and was having fun.) Jim Watson had spent the summer of 1949 under Delbruck’s tutelage at CalTech, and had returned to Bloomington a new man. He had come in 1947 from Chicago with the characteristic marks in speech, dress and behavior of an adolescent Chicago kid. He returned from Delbruck’s finishing school with a clipped b urr haircut, a new wardrobe and a distinctive German accent. The grad students at Indiana thought it was funny but we didn’t understand the transformation that had occurred when Jim came back to finish his dissertation. At Delbruck’s lecture the next year I got tickled because my first impression was that Delbruck was doing a number on Jim Watson. I quickly realized that I had got it all wrong. Jim Watson, among other things, was an intellectual chameleon; he had the capacity to absorb completely the nuances of a scientific disciplin e, and to express them unselfconsciously down to his very skin. Jim absorbed Luria and Delbruck and phage genetics before going on the Europe, to Denmark, and the Cavendish. The next time I saw him, he was with Francis Crick at Woods Hole one summer after the double helix had been announced. Even in his swim suit Jim Watson looked and sounded like a premature evocation of the Beatles, long hair and all; he was now as English as Francis Crick. Those who had not known him before thought he was British, He had absorbed another scientific culture, and was expressing it as his own.

What all this might have to do with Luria’s virus course back in 1948 is pretty hypothetical. I now suspect that Delbruck’s intellectual authority was difficult to resist, even by a mature scientist such as Salvador Luria. Perhaps Luria was trying his hand at Delbruck’s game in Bloomington, trying to be a rigorous theoretician and tactician of biology, and not involving the class unduly in the laboratory manipulations at which he was a master. He didn’t attract many graduate students with the exercise, but he apparently convinced Jim Watson that this was the fastest game in town, even if the spiel was second-hand.

I had a tangential brush with Max Delbruck several years later. I have to admit that I too was fascinated by phage genetics, though I was exposed to it only after having made a commitment to Sonneborn. When I was planning to leave Bloomington in 1951, I applied for an NRC Postdoctoral Fellowship to go to Cal Tech and work with Delbruck. I decided not to take the fellowship, however, but to accept an assistant professorship at Michigan. Years later, still a bit dazzled by the glamour of the phage legend, I wrote to Cal Tech again asking for permission to come to Cal Tech for my first sabbatical leave from the University in Ann Arbor in 1958-59. George Beadle, who was then Chairman of the Biology Division, suggested that it might be more appropriate for me to work with Ray Owen, and to continue my work with Tetrahymena, utilizing the antisera against Tetrahymena that Owen and his associates had accumulated. This was a highly satisfactory suggestion, and led to productive research and to a warm and continuing relationship. And I did get to meet Max Delbruck, sort of. I gave a lecture at Caltech, on mating type determination in Tetrahymena thermophila. Delbruck took a prominent seat near the front on the aisle before the lecture began. I had been warned that Delbruck delighted in skewering guest speakers, and to be prepared. As I remember, Delbruck asked one question with a self evidence answer at the end of the lecture. I never had a personal conversation with him.

At the time I was at Cal Tech, I was growing impatient with what I considered the sluggishness of the administration at Michigan, and was considering alternatives. I did in fact resign from my position at Michigan during my sabbatical, a procedure I do not recommend, though some get away with it without penalty. I half hoped that Cal Tech might be interested in a ciliate geneticist, but that was not to be. I have sometimes wondered if Delbruck had any influence in these matters, and whether I could have been a better salesman at the lecture he attended.

The courses of Sonneborn, Muller and Luria provided distinctive display cases of modern genetics against which genetics students would test their mettle and shape their ambitions. I do not want to give the impression that biology at Indiana was monolithically genetic, however. The largest group of graduate students when I arrived was probably that directed by William Breneman, who did endocrinological research on chickens. Breneman had been an undergraduate student at Indiana, and a football hero. He was the dramatic and highly popular instructor of the general zoology course. His lecture on the time course of evolution - From Kalamazoo to You - was famous. The students were greeted one day with a string stretched from the podium of the lecture hall across the room and out the north window, in the direction of Kalamazoo, Michigan. Attached to the string at appropriate points were ribbons symbolizing important events in the history of the universe and of life. The last ribbon, inches from the podium, represented the appearance of humans (?Man? in those prefeminist days).

Another substantial group of students was attached to Ralph Cleland (1892-1971) whose research was also relevant to the interests of genetics students, and provided some perspective into the diversity of genetic mechanisms employed by other organisms. Cleland was a prominent botanical geneticist, and National Academician, whose research reputation was based on the rationalization of unusual patterns of genetic transmission found in Oenothera lamarckiana, the evening primrose. The evening primrose was one of the plants subjected to systematic breeding analysis early in the genetics game, but it had produced results entirely different from the regular patterns described for Mendel’s garden peas. See Cleland (1981). Hugo de Vries, the important Dutch plant breeder at the turn of the century, one of the "rediscoverers" of Mendel in fact, had observed dramatic saltatory changes involving multiple traits in Oenothera. He recognized that these results did not conform to Mendel’s observations or to the behavior usually described for chromosomes. De Vries developed his Mutationtheorie (1901), as an alternative to Darwinism. This theory was a notable challenge for a time during the emerging consensus concerning genetics and evolution, one of the ?exceptions that prove the rule.?

Cleland was the careful microscopist who, followed some critical initial observations of cytogenetic irregularities reported by some German cytologists, led by Otto Renner, undertook a detailed interpretation of Oenothera’s bizarre transmission patterns throughout the North American races. All of Oenothera’s chromosomes are involved systematically in segmental interchanges, so that each chromosome shares homologies with halves of two other chromosomes. Consequently, when chromosomes pair homologously in meiosis, they line up in a single giant ring (catenation), and separate at anaphase into only two kinds of gametes. This limited recombination pattern is affected by a system of balanced lethals, and is scrambled by occasional recombination events, that break up the linked complexes and result in the sudden appearance of several new phenotypes. Cleland’s students carefully worked out the genomic complexes in the many races of the species. Oenothera had made its mark on genetic research, and helped confirm the Chromosome Theory. But, unlike the Drosophila organismic system that was revived for evolutionary, developmental and neurobiological applications later, Oenothera’s contribution was essentially dead-ended. It, like many other useful organismic explorations, has been forgotten in the rush of -events in the second half of the century.

The Sonneborn Research Group

When Fernandus Payne told me to select a research mentor, or get lost, I began seriously to consider my options among the faculty. I thoroughly enjoyed Bill Ricker, the ichthyologist who taught the biological statistics course, but he was probably the worst teacher I ever had, and I cared nothing about fish. Indiana had a long tradition in ichthyology, going back to David Starr Jordan, who had gone to California to head up the new college being built by Leland Stanford with railroad money. I was told stories about another Indiana ichthyologist, Carl Eigenmann, who believed the human brain had a finite storage capacity, and decided he shouldn’t waste his capacity on the names of students. Since Eigenmann’s fame faded over the decades, and Jordan’s persisted, the story eventually migrated west. In the 90’s I read that Jordan had refused to learn the name of a student for fear of forgetting the name of a fish. A good story, particularly a good fish story, is too good to waste. Ricker wasn’t around long enough to be a mentor anyway. He bounced back to Canada and a distinguished career in Canadian Fisheries.

I took LaMont Cole’s course in Entomology, and was sufficiently intrigued with his style and subject matter to consider seriously the prospect of becoming an economic entomologist. Cole had just written a clever analysis of the perils of interpreting population cycles. The catchy title, "The Horn of the Unicorn", brought him national attention and he soon moved to Cornell. I was interested enough in entomology to teach an entomology course at OBU one summer, as I was allowed to do to gain a little extra cash - at $100 per course. In the summer of 1948 I tried to shore up my fiscal base and taught three courses: Entomology, Embryology and Genetics, all with labs and no assistants. By this time I had been awarded a national predoctoral fellowship, and no longer had to teach courses at Indiana. So I decided that it was probably time to cut my ties with OBU. Dr Trent had told me that I would have to watch my language if I was to teach biology at OBU. The forbidden word was "evolution". Trent advised that the problem was not so much the concept as the word, and that "phylogeny" was a perfectly suitable substitute. Tired of the heavy teaching loads in a small college, and with my predoctoral fellowship in hand, I scheduled a public lecture on ?Darwin and Genesis?. That did the trick. I was not invited to return for the following summer session, and the prospect of an eventual permanent appointment at Oklahoma Baptist vanished in the summer breeze.

By this time, in fact, I had chosen a research advisor, strictly by the process of elimination. I had only seen Sonneborn at the weekly biology seminars, where the professors crossed swords with visiting speakers, and with each other. I knew that Sonneborn worked with paramecia, but I had had no formal introduction to any microbial life. I had heard that he studied cytoplasmic inheritance, but I had no perspectives on the significance of that topic. I found out where his office was located - in the Chemistry Building - knocked on the door and entered. I told him my name, and asked if he would serve as my research advisor. Apparently Dean Payne had briefed him on the qualifications of the graduate students even before I arrived. He asked me no questions but led me into the big open research lab, pointed to a desk on which sat a dissecting scope, and a water heater with a beaker of water on it. He went into another room and returned with a test tube rack containing two test tubes, labelled 51VII and 51VIII, and containing an amber liquid. He handed me a couple of reprints describing basic methodologies, and said that I should feel free to ask any questions as they arose.

I don’t remember who actually introduced me to lab procedures, inoculating the baked lettuce medium with bacteria, adjusting its pH the next day, pulling pipettes, disposing of used depression slides in buckets of water, staining cells. Ruth Dippell was probably involved. She was Sonneborn’s chief technician and laboratory manager for many years. Ruth published papers with Tracy, and before he retired she acquired a Ph D and a teaching position at Indiana. The only other regular member of the lab at that time was Bob Geckler. Johnny Preer had completed his work on the kappa particles before I joined the lab, so I did not get acquainted with him until years later. Bob Geckler’s project, like Johnny’s, might have been influenced by Muller. At any rate the work consisted of an attempt to induce mutations in paramecia. Mainly, that first semester I was absorbed with my own interactions with the protozoa; pulling pipettes, making crosses, isolating conjugating pairs, separating the exconjugants into individual depression slides, isolating caryonides when exconjugants divided, testing starved cultures for autogamy with aceto-carmine stains, growing tube cultures to sexual maturity, testing for mating types. Within weeks I was compulsively engaged with the challenge of asking questions to the microscopic animals and interpreting the answers that they provided. The experience of directly experiencing vital phenomena in an interactive way transformed my perception of science and fixed my philosophy of biological education for the rest of my career.

My total absorption in the manipulation of paramecia blocked out most other activities, and was probably responsbile for my burning the bridges at OBU the next summer. I spent late hours in the lab instead of in the Rogers N lounge playing bridge. I didn’t spend much time with Sonneborn and didn’t bother to read about what he had done, or what he thought he was up to. Nearly a year after coming into the lab I proudly presented him with a summary of my studies of mating type inheritance in variety 4 of P. aurelia. He handed it back to me the next day, saying "You could save some time by reading the literature." He then handed me some reprints of his published work on the subject, acting a little annoyed. But even at the time, I felt that he really approved my autonomy, and was pleased to have a truly independent confirmation of his work on the subject.

The Sonneborn Seminars

I didn’t start becoming socialized in the practice of science until the following fall. Students were beginning to reach graduate schools in significant numbers. In the fall of 1948 suddenly the Sonneborn lab was filled. There were Richard W. Siegel, Henry M. Butzel, Paul David Skaar and Myron Levine. There wa s even a Chinese student, Pao Kuo Chao, contending with the English language, American food, and a different kind of academic hierarchy. Post doctoral associates were also beginning to come in and participate. I don’t remember exactly who came when but I especially remember Mary Austin, the prim lady professor from Wellesley and Geoffrey Beale who brought British style and diction from Edinburgh. The Sonneborn laboratory had become a scientific society, which was called to order at sundown on Friday nights at the modest home of Ruth and Tracy Sonneborn on S. Mitchell Street.

Ruth Sonneborn was the hostess who lubricated the connections, provided good cheer and refreshments. She took the unscientific spouses out of sight and sound when the business meeting began. Sometimes Tracy talked. Sometimes students reported on published papers that were considered significant to the goals of the lab. Later the students would present their own data and be questioned about how the experiments were conducted and how they were interpreted.

As 1950 approached, and the centennary of the rediscovery of Mendel loomed ahead, geneticists from across the country and around the world came to centers of the genetic industry to celebrate and to tune in. They came to In diana perhaps primarily because of the prominence of H. J. Muller, who was something of the high priest of the gene. But some of them came because they were aware of a challenge to the hegemony of the Master Gene.

I recall particularly, because I was personally involved, a visit of Peter Medawar. Medawar was not yet known for the immunological insights that later yielded him the Nobel Prize . Nor was he known as the essayist and distinguished spokesman of science he became in later years. His publications to that time had dealt largely with phenomena of tissue rejection in skin grafts of guinea pigs, work that he had published with a younger colleague Rupert Billingham (1948). Medawar had been impressed with the possibility that plasmagenes, hypothetical autonomous copies of genes released under critical developmental stimuli into the cytoplasm, might be agents of cellular differentiation - agents derived from genes, but reproductively independent after their release. The work of Billingham and Medawar focussed on the transfer of pigmentation from grafted patches of black tissue into adjacent white skin. Cytological examination suggested that melanocytes (containing pigment granules) in the black skin could anastomose with amelanocytic dendritic cells &Mac249; in the white skin, and transfer pigment - and pigment forming ability to them. Pigmentation spread into the white skin but not into the hair follicles; white hairs remained while the pigment spread in the skin.

Medawar had given a public lecture and had been entertained by some genetics professors afterward, but had promised to visit with the Sonneborn research group in the evening. As I walked in the front door of the Sonneborn home I was met by Tracy, holding a cluster of reprints. He thrust them into my hands, pointed to the stairs, and whispered that I should go up to one of the bedrooms and refresh my understandings of the transplant work. I had given a presentation to the group on pigment spread, so was more familiar with the phenomena than anyone else. Tracy said that Medawar was one of the most unreactive guests he had ever entertained. He got few questions after his talk, and he managed not to answer them cogently. With the professors he mainly asked questions and smiled. Tracy told me to come up with something that would make him talk. Fifteen minutes later I slipped down the stairs and into the parlor and started quizzing Peter Medawar. At first he parried my questions, but he soon realized he was under attack, and rose in animation to the defence of his plasmagene interpretation. A good time was had by all, especially me. To my knowledge, this was the end of the plasmagene explanation of pigment spread, an unnecessary complication in the story of the simple diffusion of formed pigment granules from pigmented to nonpigmented cells.

I happen to recall the Medawar visit particularly, because of my special involvement with him, but the Sonneborn parlor hosted many distinguished biologists in an informal social setting with free interaction with students and staff. Jim Watson, though not a declared student of the Sonneborn group, actively participated in the Friday night activities, which were the most distinctive feature of the educational program for young geneticists at indiana in those days.

The interaction among Sonneborn’s students and staff in the laboratory was continuous into the evening and into the weekend. The students brought experiences and expectations every much as distinctive as mine, and these were compared, abraded and swapped during incessant intercourse. Dick Siegel had been raised in the liberal New York Ethical Culture society. Myron (Mike) Levine was from an observing New York Jewish background. Henry Butzel was a scion of a privileged and cultured Michigan legal family, of Jewish but non-observing background. Dave Skaar came with no expressed religious prejudices. William Tepe was a devout Catholic who wanted to teach biology in a Catholic university. He contended for awhile with the epistemological noise, but eventually was pushed into complete commitment to his faith. He left Indiana, went to a seminary and was eventually content to be called Father Tepe, and to baptize babies into the Catholic fold. Jim Watson’s mother was a practicing Catholic, but his agnostic father would take him birding on Sunday mornings to avoid Mass. P. K. Chao had been raised in a Presbyterian compound in China, but he seemed not to observe religious practice of any kind in Bloomington. The persistence of his cultural connections were manifested in his marriage, with a bride selected and a wedding arranged for him by his brother in Chicago, after he had completed his Ph D.

I mention the religious backgrounds of the participants in the Sonneborn seminar group, not because religious belief or practice was a common topic of dispute or discussion, but simply because the tags provide some indication of the cultural diversity that we represented. Only occasionally did problems arise with respect to religious practice. I remember, for example, Mike Levine’s concern at first over having to assist in classes scheduled without appropriate regard of Jewish high holy days. I also remember one dramatic eruption on a theological topic at lunch one day in the Roger’s dining hall. For some reason the discussion drifted into the topic of predestination. Bill Tepe, and another graduate student working in Chemistry, were discussing divine constraints on personal salvation. Dick Siegel, always the most spontaneous and volatile of the Sonneborn crew, suddenly pushed back his plate, his lunch half finished, stood up and declared to all in hearing range, ?God Damn God Anyway?. He stalked out of the dining room alone. I noticed particularly that he did not buss his dishes, as we were supposed to.

My most memorable embarrassing experience came from what I consider an awkward management of a cultural divide. It happened at the end of one of the Friday evening sessions. After a talk had been presented and the serious discussion completed, Ruth Sonneborn would appear, along with any spouses who might have been waiting out the working sessions. Beer and soft drinks were provided and the conversations broke up into smaller groups, and less scientific subjects. Although I had been away from Oklahoma Baptist for two years at this point, and although I considered myself something of a free-thinking independent, I had never drunk a beer or tasted a distilled drink. Most of the seminar participants drank beer, but I was too inhibited. After some months, however, I was feeling comfortable enough to grab a beer when they were handed around, and tentatively took a swig. I was sitting in a back row trying to imbibe the distasteful brew. I was sitting next to Willem J. van Wagtendonk. Wim was the biochemist whom Sonneborn had recently employed. He had been added to the group to develop a defined culture medium for paramecia and to make a beginning of biochemical genetics for this organism. Wim was a big gruff guy from the Dutch East Indies who eventually developed an adjunct group of graduate students. He was of a phlegmatic nature and difficult to ruffle. His equanimity was significantly disturbed, however, when the half bottle of beer I had consumed was released into his lap. To this day I do not know whether I am allergic to beer or if my ingrained inhibitions promoted stomach contractions. I have never learned to drink beer.

I did not think much about the quasi-religious significance of the sundown ceremonies on South Mitchell Street until the time our grandson, Brian, told about the social constraints enjoyed by his high school friend Rafael, whose family had recently decided to become observant Jews. And then I thought of the young Tracy Sonneborn, turning his back on his calling to be a rabbi, and later establishing in his home an awesomely effective ritual enacted every Friday night in the name of disinterested experimental biology. Every one of us came to that symposium with lots of baggage, but the main brew of the evening was not liquid but intellectual, and it washed over the distinctions we had brought with us and made us an ecumenical whole. A sign of our community came to be the tobacco. Tracy Sonneborn smoked a curved pipe, and one by one curved pipes appeared in the mouths of most of his committed graduate students. There were a few exceptions. Dave Skaar stuck with his cigarettes, and died young. P. K. Chao couldn’t overcome his cultural aversion to tobacco, any more than I could drink beer.

More importantly, we hung on the words of Tracy Sonneborn, eager for signs of approval, and were mortified by imagined slights. We often considered suggestions to be commands, and queries to be research programs. P.K. Chao’s research project was an example of the transformation of a casual wish into a program. John Preer’s work had led to the identification of microscopically visible Feulgen-positive particles in the cytoplasm of killer paramecia. One Friday night Tracy mused, "I wonder how many kappa particles there are in a cell, and if the number is different in cells heterozygous and homozygous for the K gene." P.K. thought he had been directed to count the kappa particles. The counts had to be made with oil immersion lenses and with optical grids. With several hundred particles per cell, often obscured behind the nuclei, this would have been an heroic task under the best of circumstances, but P. K. started on the project without a complaint. He had counted kappa in a number of cells, but finally had to quit because of illness. He had developed nausea and painful headaches that were interfering with his progress. Someone supplied him with a bottle of whiskey to keep in his desk, hoping that might help.

Finally one day Ruth Dippell asked P. K. if she could look at one of his slides, thinking that the quality of the staining might be part of his problem. What she discovered was that P. K. didn’t know how to use binocular microscopes. All the microscopes he had access to in China were monocular. He didn’t know that one of the oculars had to be adjusted to provide appropriate binocular vision. He had essentially been doing kappa counts with one eye, while trying futilely to see something in the other badly adjusted lens. Once he learned to use the microscope properly, his headaches disappeared, and he completed the counts quickly. Paramecia homozygous for the K gene had twice as many kappa particles as paramecia th at were heterozygous.

This simple dose response suggested that paramecia provided some substance that was necessary for the reproduction of the kappa bacteria, and that KK animals could support twice as many particles as Kk animals. This result was not particularly astonishing, but P. K. also mentioned another observation that was quite unexpected. Since conjugating pairs of variety 4 consist of one cell of type VII and one of type VIII, and yield exconjugant clones of types VII and VIII, one could also ask whether the mating types of the cells were in any way related to the number of kappa particles they maintained. P. K. ‘s results indicated a positive answer. In fact, type VII KK homozygotes had almost exactly twice as many kappa particles as the type VIII KK cells. And type VII heterozygotes had twice as many kappa particles as type VIII heterozygotes,

At that point Tracy turned to me - who had been repeating his studies on mating type determination in these same stocks. "Dave, how do you explain that the type VII cells have twice as much kappa as the type VIII cells"? I could only shake my head. But in the night I woke up, and KNEW that the type VII cells had twice as many K genes as the type VIII cells, and that the essential difference between cells of the two mating types was that the "odd" mating type had twice as much of "something" as did the "even' mating type, and that this quantitative difference explained in some way the expression of alternative mating types. This was the first "Eureka" moment in my scientific career. Maybe it was my only eureka moment; I can’t recall another that was so dramatic.

The interpretation came fleshed out in full detail, very much as it was described in my thesis when it was eventually published (Nanney, 1953). The quantitative genetic difference I referred to as a “ploidy” difference, though I was aware that a dosage difference could be characteristic of the whole genome, or of only a part. And I knew that the polygenomic macronucleus was a complicating factor that would have to be dealt with.I perceived the cytoplasmic feed-back of mating type determination in terms of a threshold effect on the doubling of some macronuclear component in its development. The whole scheme was so apparent that I felt chagrined. I suspected that Tracy had seen it, had given me a chance to speak it, and I had failed the test. Then a small hope arose. Maybe he hadn’t seen the explanation. Maybe I still had time to say the obvious before it occurred to him. In the small hours of the night I wrote out the explanation. With dawn I rushed to the lab, hoping that Tracy would be in his office early on a Saturday morning. He was not, but I pushed my explanation under his door, and walked back to the Rogers dining room for breakfast.

To this day I also don’t know whether some genetic component in P. tetraurelia is present in the macronuclei of mating type VII in twice the quantity as in the macronuclei of mating type VIII. I am as convinced today as I was that night in Bloomington that mating type continuity in this species is maintained by an epigenetic control system (though that term was not established as appropriate for another decade) consisting of “messages” derived from old differentiated macronuclei that induce corresponding activity states in newly developing macronuclear anlagen. “Cytoplasmic inheritance” in this situation is not a manifestation of an alternative system of heredity but a manifestation of the sophisticated regulation of nucleic information.

To this day, also, I don’t know if Tracy Sonneborn was independently aware of the kind of interpretation I was proposing. He subjected me to grueling interrogation at successive Friday evening sessions, and critically proposed amendments to my clumsy first drafts of manuscripts. But I felt that he was in fact, from the beginning, but reluctantly, in agreement with the thrust of my arguments. I repeatedly tried to get him to coauthor my interpretive essays, but he obstinately refused. I don’t know whether he was reluctant to abandon the vague concept of a protoplasmic memory for fundamental vital properties, that may have appealed to him as a counterweight to the Mullerian-nucleic hegemony. Or whether he recalled his own disappontment, maybe outrage is a better term, when H.S. Jennings - his mentor - announced his (Tracy’s) discovery of mating types in Paramecium in the course of a lecture he had been scheduled to give, and without asking permission. Tracy swore he would never be accused of misappropriating credit for work done in his lab. Perhaps the "epigenetic" explanation of many cases of cytoplasmic inheritance was as much his as mine; I always assumed so. Perhaps if he had acknowledged his ownership and helped with the explanation, the delay in acceptance might have been shorter and the ciliated protozoa might have received more credit for their contribution.

The Sonneborn research group began in the 1940s and continued for many years with continuous turnover of personel. It was the fountainhead of an international complex of workers focussed on ciliate genetics. The number of graduate students continue to rise for several years, overlapping with those who made up my cohort. Immediately following us were Paul Margolin, Earl Hanson, Elias Balbinder, Bruce and Barbara Burns, Audry Barnette, Dot Widmeyer, and on and on. From overseas in addition to Geoffrey Beale with his Scottish connection, Ko Hiwatashi came and went from Bloomington back to Japan to establish a Japanese Paramecium genetics industry. Renzo Nobili returned to Italy and developed at Pisa a remarkably industrious and metastasizing research group. Janine Beisson headed the French Paramecium efforts. Klaus Heckmann took ciliate genetics to Munster, Germany. And once socialized into the Sonneborn society, the web of interactions usually continued with increasing strength. Participants often returned for visits and for sabbatical leaves. Manuscripts were transmitted, englished, criticized and returned.

But this remarkable society eventually faltered. The placement of Sonneborn graduates became more difficult. Students often had to take postdoctoral training to learn new organismic techniques. Many had to take teaching positions in smaller educational institutions with heavy teaching loads and less satisfactory research facilities. John Preer took a position at Pennsylvania and returned to Bloomington as Sonneborn approached retirement. I taught at Michigan before going to Illinois. Dick Siegel went to UCLA. But most of Sonneborn’s later students had restricted opportunities to continue in the activities for which they were trained. We will return to consider this phenomenon after a detour.

The Sonneborn-Jennings Connection

Tracy Sonneborn’s relationship to his sponsor has long been a puzzle to me, and the few shafts of light came through cracks in a facade. Tracy never really said much about his mentor, and the only papers of Jennings that I read as a graduate student were those on mating type inheritance in Paramecium bursaria. These were products of Jennings’ late years, after Tracy had discovered mating types in P. aurelia. Jennings’ studies did not show the differences between sister conjugants or between sister caryonides that made the P. aurelia phenomena a challenge. As I remember, the bursaria results suggested direct genic control of mating types in this species complex, but with unresolved complications, particularly with respect to selfing clones.

Only recently have I become aware of an interesting insight into Sonneborn-Jennings relationships, and that derives from further work on P. bursaria. As a graduate student Dick Siegel worked on "mate-killers", a phenomenon related to the better known killer trait, except that the killing occurs when cells are conjugating. Like many graduate students, including me, Siegel needed some emotional distance from the intense relationship that Sonneborn developed. As I shifted from Paramecium to Tetrahymena, Dick moved from the P. aurelia complex to the P. bursaria complex, and consequently came into indirect relationship with Jennings, through his work on P. bursaria. Siegel was not able to replicate some of Jennings’ observations, particularly the erratic appearance of selfing within his cultures. After examining Jennings’ experimental methods, Siegel suggested that the observed selfing (mating within a culture) was not in fact true selfing, but evidence of contamination between cultures that were maintained in stacked casters. After explaining away the selfing phenomena, Siegel was able to rationalize his breeding results in a coherent synthesis that accounted for direct genic control of mating types in this species cluster.

According to Siegel, Sonneborn was not pleased with this clarification, but was angry that Siegel had criticized Jennings’ experimental work. While Siegel had maintained in some ways the closest relationship of any of Sonneborn’s students, his criticism of Jennings seemed to have provoked a disproportionate response, so unexpected as to damage his continued participation in the ciliate community. Siegel was on the staff of UCLA at the time. He terminated his Paramecium studies, made contact with Seymour Benzer, and made his later experimental contributions searching for behavioral mutants in Drosophila.

I interpret Sonneborn’s relationship to Jennings as seriously conflicted. Jennings provided him an academic home when few opportunities were available, and gave him freedom to pursue studies that were initially fruitless. Once he arrived in Bloomington, however, he seldom mentioned Jennings, at least to his students. Jennings died in 1946, and Sonneborn as his most distinguished student was expected to write his major biographical summary for the American Philosophical Society. Sonneborn did not finish the job for nearly 20 years, and it is 'not the warm recollection expected from a student, friend and colleague.

I was shocked much later to read in Dan Kevles’ 1985 book on the American Eugenics Movement that resistance to applying genetics to human breeding was rare among professional geneticists in America. He identified Jennings as a lone voice in American genetics criticizing the premature application of genetic knowledge. In England, where eugenics first took root in the land of its originator, Francis Galton, several prominent geneticists openly opposed state sponsored genetic selection, particularly J.B.S. Haldane, Julian Huxley, and Lancelot Hogben. Meanwhile, wealthy corporate Americans underwrote genetics research and supported involuntary sterilization of the unfit and criminals. They opposed the immigration of genetically inferior peoples. And American geneticists waited until 1933 to shed the mantle of eugenics, and return to a more scientific foundation. Except for Jennings. If my professor had played such a role, I would have let people know about it.

Even more recently in the work of Schloegel and Schmidgen (2002) I discovered that Jennings had substantial influence in shaping educational theory, particularly by emphasizing the significance of education in shaping human behavior. He was once awarded an honorary degree, along with Sigmund Freud, for his contributions to educational practice. Sonneborn never told me about such things.

One final note on the Jennings-Sonneborn relationship needs to be added for completeness, though its source is problematic and its significance uncertain. Gary Grimes was one of Sonneborn’s later students who revelled in the morphogenetic complexities that Sonneborn elucidated during his later years. I once asked Gary what he knew about Jennings. Gary responded, "Did you know that Jennings was gay"? My reaction was utter skepticism. But Gary said, "Oh yes, Tracy told me." Tracy didn’t tell me. Ruth Sonneborn’s rare references to the Jennings were to "Dr. and Mrs Jennings...", and I had no picture of the Jenningses as anything other than fussy old fuddy-duddies. Jennings married late, to his brother’s widow, and had no children. If one were a novelist one could spin a web. But all I have is circumstantial evidence and unsupported hearsay.

Cytoplasmic Inheritance Goes International

The local clashes over cytoplasmic inheritance at Indiana began to acquire an international flavor as the Mendelian semi-centennial approached. Maternal inheritance of organismc traits was not a new phenomenon. Indeed, Carl Correns, the German plant breeder who was one of the rediscoverers of Mendel's work, described the maternal inheritance of plastid characteristics by the end of the first decade of the century. This particular Mendelian exception had been "tamed", however, by a developing consensus that the plastids of higher plants are relics of an ancient symbiosis with unicellular algae. Plastids were thought to have parts of their own original genetic systems. DNA had not been firmly established as the genetic material yet, and DNA had not yet been demonstrated in chloroplasts, but the Pneumoccocus work on DNA transformations, and related work on viruses, was being widely accepted as critical evidence for a common genetic substance and a single basic hereditary mechanism in all organisms. Other exceptions were also being tamed on the grounds of secondary properties of evolved genetic systems. Work on the catenated chromosomes of Oenothera, for example, like the studies on the sex chromosomes of mammals and insects, showed that the exceptional transmission of traits often corresponded to exceptional behavior of chromosomes.

The most persistent challenge to the hegemony of the nucleus (and of its nucleic acid components) came, however, from developmental biology rather than from genetics. Developmental biologists had long maintained that cells acquire specific traits in the course of embryonic differentiation, and that these traits are "hereditary" in the operational sense that differences are maintained indefinitely during cell division in the same environment. An early proposal, by Weismann, to explain this phenomenon supposed that chromosomal material might be actually sorted out to differentiated tissues. That suggestion failed with increasing evidence for the reliability of mitosis and the chromosomal constancy of cells throughout the organism.

The Developmental Paradox and Microbial Systems

The developmental paradox - the chromosomal equivalence of cells with hereditary differences - could not be approached by classical breeding procedures because the techniques of "breeding somatic cells had not been developed. Microbial geneticists seized on the fact that "somatic' cells can serve as gametes as an opportunity to cross "differentiated" cells and thus to resolve the paradox. If mature paramecia are "differentiated" with respect to mating type, or to serotype, then controlled matings should re 1veal the physical basis of these differentiated traits. Sonneborn’s evidence for the cytoplasmic basis of these traits, even when combined with clear evidence of nuclear control of the repertoire of cellular characteristics, was compelling, and discomfiting.

Other workers, using organisms with more conventional nuclear organization, employed a similar strategy and came to similar conclusions. The green alga Chlamydomonas reinhardii, for example, in the hands of Ruth Sager provided evidence of cytoplasmic inheritance of several characteristics, with physical locations in the chloroplasts. And Boris Ephrussi in France undertook the analysis of respiratory traits in baker’s yeast, Saccharomyces cerevisiae. In time these examples came to be understood as evidence of yet a third reservoir of genetic information in the cell, housed in the mitochondria, the "power houses" of the cells. Mitochondria were found to contain DNA, and DNA whose evolutionary origins could be traced to an ancient incorporation of a bacterial symbiont into the eukaryotic lineage. All eukaryotes have two nucleic reservoirs, though in a few the reservoirs have degenerated almost to nothing. Green plants have a third reservoir constituted as relics of ancient algae. Many organisms also contain less general and less integrated symbionts that affect their physiological activities and reproductive behavior. None of these studies diminished the magisterial role of DNA in the cellular economy, only questions about its placement in the cell and its transmission patterns.

But none of these explorations cast much light on the developmental paradox, and some studies of microbial cellular differences refused to be explained on the basis of plastids, mitochondria, or more recent symbionts. These examples included the classic mating type and serotype studies in paramecia. The eventual solution to the paradox came from an unexpected source, in the intellectual concerns of engineers in developing and interpreting control systems for complex instruments and machines. Norbert Weiner’s Cybernetics: or Control and Communication in the Animal and the Machine, was published in 1948 and reflected the considerations being developed by Shannon, Weiner and others on the measurement and management of “information”. I still have my copy of Weiner’s book, still wrapped in the protective plastic cover provided by the IU Union bookstore. It is one of the few books that has survived the constant turnover of a bookoholic’s career, though I haven’t opened it in half a century.

The first usually acknowledged application of cybernetic principles to the maintenance of a particular physiological states in cells was that of Max Delbruck at a conference on cytoplasmic inheritance held in France in 1948. Sonneborn had sent his younger British colleague - Geoffrey Beale - to this conference to present the phenomena associated with serotype expression and transmission in Paramecium. Sonneborn, like most Americans, had little experience in speaking a language other than English, and English was not yet the lingua franca of science. Beale was a capable spokeman for the laboratory, and described the evidence for cytoplasmic transmission of patterns of nuclear expression. One of the mechanisms considered was that of plasmagenes, initiated as copies of nuclear information that were replicated autocatalytically in the cytoplasm. Delbruck spoke in the discussion session following the presentation and drew a simple diagram of two parallel enzymatically catalyzed metabolic pathways, with mutual negative feedback loops between the pathways; he suggested that a product in each pathway could interfere with a reaction in the competing pathway. An established pathway would prohibit the development of the other. The expression of a pathway would be maintained in a "steady state", "remembering" past experiences, and resisting perturbance. Hereditary expressions could be maintained in the absence of genic or environmental differences, and without plasmagenes in the cytoplasm. At first this suggestion was a picture, a diagram with no identifiable molecular components. It was important, however, in raising consciousness about another way of considering cellular homeostasis.

Franco American Relationships

In 1950, Andre Lwoff , at the Institut Pasteur in Paris published a little book called Problems of Morphogenesis in Ciliates: The Kinetosomes in Development, Reproduction and Evolution. The book was based on the Dunham lectures that Lwoff had given at Harvard University in 1947-48. It covered the morphogenetic studies initiated by Edouard Chatton, using the silver-impregnation procedures that are known universally as the Chatton-Lwoff techniques. The controversial thesis of the book was that another cytoplasmic organelle, the kinetosome, or basal body, located at the base of the cilium and eukaryotic flagellum, is a "Visible Cytoplasmic Unit Endowed with Genetic Continuity". Edouard Chatton was a prominent French protozoologist who made major contributions to the technology of descriptive protozoology, beginning in the 1920s. He was also the person who first introduced the terms prokaryote and eukaryote to designate the discontinuity in living systems that Roger Stanier and C.V. Van Niel later elaborated into the most fundamental distinction within the kingdoms of life. Three of Chatton’s students became outstanding leaders and rivals in French biology, particularly after WWII: Andre Lwoff (1902-1994), Jacques Monod (1910-1976), and Boris Ephrussi (1901-1979). Two of them were named as Nobel Laureates, and the third almost made it. All three were involved in different ways with the resolution of the nuclear-cytoplasmic dilemma in the 1950s. Their relationships with Americans during the negotiation of the definitive role of the nucleus provide interesting examples of agonistic interaction on an international playing field.

Sonneborn was invited to review the Lwoff book. His review (1951) could be generally interpreted as favorable, but perhaps "picky". By 1950 the struggle between the nucleus and the cytoplasm had taken on political overtones, not totally independent of the international cooling with the Cold War, and the domestic warming with Joseph McCarthy and the Red Scare. Trofim D. Lysenko had abrogated international standards of scientific evidence in overthrowing "Western" genetics in the USSR, and won Stalin’s connivance. Lysenko's criticism of the western genetics was based on no substantial evidence, but could be confused in its rhetoric with complaints about the inadequacies and confusions surrounding the concept of the gene, particularly in his criticism of Weismann and Morgan, and the concept of the gene. Sonneborn resented being lumped with Lysenko and other heretics of modern genetics, and urged the most rigorous application of the rules of evidence. He was sufficiently sensitive to subtle charges that he was a secret sympathizer that he accepted an invitation to give lectures in South America. He thought that if he spoke out against Lysenko while away from CIA surveillance that his stance on cytoplasmic inheritance might be more credible.

In applying his rigorous scrutiny to Lwoff's thesis, he had to express doubt that the evidence for genetic continuity of kinetosomes was sufficiently secure. This position was related to his rising interest in the continuity of the ciliate cortex. The hereditary inertia of the compound ciliate cortex was obviously a formidable challenge, and one related to the functions of the rows of basal bodies along the ciliary rows. But Sonneborn doubted that postulating genes in kinetosomes solved any problems, and didn’t think the evidence for them was compelling.

Sonneborn had not been particularly interested in the ciliate cortex before the publication of Lwoff's book, but much of his subsequent research career was dedicated to understanding ciliate morphogenesis. This work got off to a promising start with a joint study carried out with Janine Beisson (Beisson and Sonneborn, 1965), an intelligent and technically skilled young French investigator. Their work on doublets and cleverly constructed cortical mosaics eventually eclipsed the ciliate work on mating types and serotypes, and indicated yet another kind of hereditary mechanism at the cellular level. But the immediate effect of Sonneborn's review was a rupture of relations with Lwoff. Sonneborn had visited Lwoff in Paris, and had previously enjoyed cordial relations with him. Those relationships were over. At international gatherings Lwoff would refuse to speak to him; he would leave the room if he saw him from a distance. Mutual friends explained the situation as a cultural clash. Sonneborn believed that a request for a review was a request for a candid appraisal. The French believed (his intermediary explained), that only two kinds of reviews exist: friendly and hostile. To accept an invitation and give a qualified approval was tatamount to declaring oneself an enemy. Lwoff claimed that Sonneborn deliberately misrepresented his views and quoted him out of context.

The scientific questions concerning the roles of the kinetosome in cortical heredity were dragged out awhile, quite aside from the issue of personal relationships. For awhile some investigators thought that the issue would be solved if DNA could be found in basal bodies, and several hopeful reports were published, and then contradicted, before the realization dawned - in agreement with Sonneborn's critical opinion - that having DNA reservoirs at a few cytoplasmic depots was not going to explain very much about morphogenesis even if it were true. The story of ciliate morphogenesis remains an unresolved challenge (See Joseph Frankel’s 1990 review).

Andre Lwoff's Nobel Prize (in 1965 with Jacob and Monod) didn't have much to do with ciliates or with genes in the cytoplasm, however. He shared the Prize for studies concerning the control of gene action, particularly with respect to the control of prophages in the bacterial chromosome. How he become included in the French award is a curious story in itself, bearing as it does on international relations in science and the politics of high prizes. I have to jump ahead a few years to pick up that story. I came to the University of Illinois in 1959, in part because of the influence of Sol Spiegelman (1915-1983), who was a professor of Microbiology at Illinois at the time. Though I was brought in as a Professor of Zoology, the &Mac176; molecular biologists - concentrated in the Microbiology Department - were in the ascendancy and had a new building. Space was made available to me in Burrill Hall, next to Sol's laboratory, and I was regularly entertained with his curious soliloquies.

When I first arrived in Urbana, Spiegelman was fearful that Jacques Monod might acquire a Nobel Prize and that he would be excluded. Monod and Spiegelman were a lot alike, short, brilliant, competitive, and engaged in similar work, particularly in the control of physiological proceses in bacteria, Spiegelman's best known work was on cell-free viral replication, and on induced enzyme syntheses. One day Sol called me into his laboratory in great excitement. His "spy" in Paris had reported that the next Nobel Prize was to be given to Jacob, Monod, and Spiegelman. Charles de Gaulle had decided that France needed a Prize, and had appointed a commission to determine the most promising work to be proposed. His commission reported that the research most likely to be successful was the work at the Pasteur Institute on the control of gene expression. Although such a formulation could not produce a purely French prize, De Gaulle had accepted the imperfect package, and had invited the Swedish Academy to be wined and dined in Paris. Spiegelman's reporteur believed it was a done deal. But when the prize in physiology and medicine was announced that year bypassed molecular genetics entirely and was awarded in another area.

A couple of years later, however, the lightning did strike, but this time the prize was formulated so as to include Andre Lwoff, and to exclude Sol Spiegelman. Sol was devastated, pronounced that his teaching days were over (in the middle of a course in Microbial Genetics that he was teaching with me). He announced that he would hereafter study mouse tumor viruses, and try again for a Nobel Prize in a different but related area. His colleagues at Illinois told him he would have to wait until appropriate confinement quarters were constructed before he could work on mammalian tumors. Thereupon Sol announced that he would move to any institution that would make it possible for him to make the research transition quickly. He moved to Columbia University (1969), and after a few years died of pancreatic cancer.

My only personal interaction with Lwoff occurred at a conference on cytoplasmic inheritance in Gif-sur-Yvette in 1957. I had given a presentation on mating type inheritance in Tetrahymena. At the end he asked how could I be talking about mating types controlling mating in Tetrahymena when Chatton had demonstrated in the 1920s that one can induce mating in Tetrahymena by controlling the food supply. He was referring to a report by Chatton that mating in these organisms (only subsequently named Tetrahymena by Furgason - in 1940) depended on certain environmental conditions. I explained that both intrinsic and extrinsic conditions must be appropriate for mating. Some tetrahymenids, such as the Ann Arbor strains I first studied, are constitutional selfers and will mate readily whenever their food supply is restricted. Other strains, such as the Woods Hole strains that I was reporting on at Gif, will mate only when clonal cultures of appropriate mating types and clonal age are mixed, again when food is restricted. Nature doesn’t tolerate sex when multiplication is possible.

After my lecture I asked Sonneborn why Lwoff asked such a stupid question; surely he knew about mating types" Sonneborn’s response was to say, ”Obviously this was the first time Lwoff has heard you lecture. He asks that question whenever he hears the term ‘mating type’ mentioned.” This was apparently a ceremonious evocation of the name of his mentor, a reminder that Chatton, and the French, were there first.

The Ephrussi Business

The Ephrussi interactions that I was aware of also began with a little book, and also involved nationalistic interests and claims of intellectual property. The book was Nucleo-cytoplasmic Relations in Microorganisms, published in 1953 by the Oxford University Press. It was based on a series of lectures that Ephrussi had given at Yale. The lectures dealt with the French studies, primarily by Ephrussi and Slonimski, on certain respiratory mutants in yeast, whose simplest phenotypic expression was the formation of small (petite) colonies on agar plates. These mutants could be specifically induced by treatment with certain dyes and were irreversible. Evidence was accumulating that in fact the mutations involved the damage or loss of mitochondria. The extensive parallels between the petite phenomena in yeast and the killer phenomena in paramecia were developed in the text. In the case of the yeast cells, however, the phenomena involved a more fundamental vital property. Ephrussi used the peti te studies as the platform to survey other cases of apparent cytoplasmic inheritance and to place them in context. With this book Ephrussi became the principal European exponent and interpreter of cytoplasmic hereditary phenomena. His theme of “nucleo-cytoplasmic interactions” echoed Sonneborn’s “partner of the gene” theme and was close to the title of Sonneborn’s presentation “Patterns of nucleo-cytoplasmic integration in Paramecium”, which was to be presented at the 9th International Congress of Genetics.

A few years later, in 1956, an international conference on ?The Role of The Cytoplasm in Heredity?, was organized at John Hopkins University. Boris Ephrussi, Joshua Lederburg, and various other leading geneticists were to participate and Sonneborn was invited to speak. He demurred, however, and recommended that I speak on behalf of the ciliate community. This was my first major opportunity to present a brief on behalf of the control systems approach to cellular differentiation and cytoplasmic inher itance. I suggested that genic information was the foundation of vital phenomena, but that the genes could be either in the nucleus or in the cytoplasm - as in plastids, mitochondria or symbionts. The management of states of expression of genetic capabilities could mimic genetic phenomena, but have inertial properties based on physiological states. I suggested that the mechanisms of maintenance be referred to as "genetic" and ?"paragenetic, and that the cellular location was not the significant distinction. Lederburg didn’t care for my distinction, and proposed instead that the terms should be "nucleic" and epinucleic". I was not able hear all the papers or to participate fully in the discussions. After my talk I received a phone call saying that Jean had gone into labor and that I was needed back home. I left Baltimore quickly and returned to Ann Arbor, and trauma. Our first born son died at birth in the hands of an unskilled trainee at the University of Michigan Medical School.

The Conference at G if-sur-Yvette

The conference at Gif-Sur-Yvette was convened in 1957, and was a more partisan gathering of sympathizers with cytoplasmic inheritance. It was organized by Boris Ephrussi as a closed meeting with strict rules of confidentiality laid down as to protect the intellectual property of those attending. Tracy Sonneborn and Ruth Sager were present from the United States, and Sonneborn had requested permission to bring me along also. It was my first trip abroad, and I was in culture shock. I met some of the French biologists whom I had only known by name: Lwoff, Ephrussi, Jacob, Monod. I remember a few of the British who were there, particularly Pontecorvo, because he criticized the Greek derivation of the term I proposed for cellular control systems. He said that ?paragenetic? (my term) was less appropriate etymologically than ?epigenetic?. Unfortunately, he said, Waddington was the person responsible for the modern usage of the term epigenetic, and Waddington was, unfortunately, unable to come to the meeting. If we were going to extend "epigenetic? to include cellular control systems - far beyond the usage of Waddington, much less Aristotle, we had to have Waddington’s permission. Aristotle was clearly out of reach.

All this discussion began on the first evening of the conference, after a long and dramatic call to arms by Boris Ephrussi. He proclaimed that the conflict between the nucleus and the cytoplasm was growing increasingly desperate, and the defenders of the cytoplasm needed to gather forces and develop new strategies. At end of his peroration he asked if any of the participants had further comments before we adjourned to dinner. I was time warped and culture shocked, but I rose and told the organizer that he had it all wrong. The silly argument over territorial imperatives was really beside the point. I explained a little about the properties of control systems, and proposed some terminology that I thought might be helpful.

Ephrussi did not respond to my cheeky remarks, and in Ôdeed managed not to say a word to me until after the last session some three days later. Then, he capitulated in his final summation. He said that he knew the conferees would be startled if Boris Ephrussi should admit that he had been wrong. But, he said, unfortunately, Nanney is correct. The “geographical question” is no longer relevant. He himself would no longer consider it appropriate to defend cytoplasmic inheritance in the sense that had been posed. I arrived in Paris afflicted with jet lag (although the propeller plane I flew took 14 hours from New York to Paris); I departed Paris in euphoria.

Some months later I received a call from Tracy Sonneborn. He and Ruth were spending a sabbatical leave at the Institute for Advanced Studies in Princeton. Would it be possible for Jean and me to take a spring break from Ann Arbor and make a quick trip to see them" We could stay in their guest room.

When Jean and I, and our infant son, Douglas, arrived in Princeton, the cocktail party at the Sonneborn apartment was in full swing. Tracy met us at the door, grabbed me by the arm and said, ?Come, we have to meet Ephrussi’s train in just a few minutes.? On the way to the station he explained that Ephrussi was giving a major address at the Gatlinburg Conference at Oak Ridge, and was stopping by for a visit in princeton also. As Ephrussi came off the train, he handed me a manuscript and said ?Read it; I know that you will approve.? We rejoined the party and Tracy pushed me into his little study with the manuscript, and reiterated Ephrussi’s instruction, ?Read it.?
So I read it, with the distant sounds of the party fading:

I was asked to speak on ?The Cytoplasm and Somatic Cell Variation? And I apologize for not adhering strictly to this title. To a large extent this is because a Conference on ?Extrachromosomal Heredity,? recently held at Gif, greatly influenced my thinking on this subject. The views I am going to express here probably do not represent the cross section of opinion at the Gif Conference, but they bear its mark. In particular, they are influenced by some arguments of Dr. D. L. Nanney, in part unpublished (see however Nanney, ‘57) to whom I therefore owe an apology for using them, explicitly or implicitly. One of Nanney’s points is that the usual classificaton of genetic mechanisms on what he calls ?the geographical basis? may be very misleading. Nanney’s arguments appeared to me so important that, in preparing this paper, I was unable to limit myself to the consideration of the cytoplasm alone.

I was returned to euphoria, at least momentarily. The door opened and Sonneborn and Ephrussi came in. Tracy said, ?Well. What do you think"? I answered that I thought it was a very good paper. It should be because it was a fleshed out version of my comments at Gif. I could hardly disagree. Sonneborn, however, was not satisfied. He conten Yded that Nanney was mentioned in the first paragraph, and the rest of the presentation was pure Boris Ephrussi. Ephrussi had appropriated my perspective, and his public presentation of my ideas was inappropriate, particularly since Ephrussi had not asked my permission to use the ideas in this way, despite having been repeatedly urged to do so by Sonneborn. Ephrussi’s rejoinder was interesting. He said that the person who gets the credit for the ideas is not important. What is important is that geneticists, developmentalists and physiologists get on with their common tasks, and not waste time in further territorial arguments. He contended that if Nanney were to say these things, no one would be impressed. If Boris Ephrussi made the announcement the stupid battle would be done. This argument infuriated Tracy Sonneborn, and the voices drowned out the chatter in the next room. Tracy insisted that Ephrussi not be allowed to give this speech (the following week), unless Nanney had an opportunity to publish the ideas before the Gatlinburg proceedings were published. I was sent from the room, to visit with the departing guests, who were curious about the shouting match taking place in the study.

The dispute ended with phone calls to the editor of the Proceedings of the National Academy of Sciences. A paper by Nanney could be published in the PNAS before the Gatlinburg conference was published if it were received within three weeks. So Boris Ephrussi gave his paper in Gatlinburg, Tennessee, and I went home to Michigan to write “Epigenetic Control Systems”.

Later, Tracy told me that he had been furious with Ephrussi about the lectures published as Nucleocytoplasmic Relationships in Microorganisms. Ephrussi had discussed the material on yeasts and paramecium with Sonneborn before preparing those lectures, and had been told of Sonneborn’s perspectives, and of his intention to write a small book on the subject. The book that Ephrussi wrote essentially preempted the book that Sonneborn had in mind. As soon as Ephrussi announced the Gif Conference, Tracy had been suspicious that he was looking for a subject matter to be given at the talk he had already agreed to present at Gatlinburg. Seeing how Ephrussi responded to my remarks at Gif, Sonneborn suspected that he was planning a talk about control systems, and was determined that he would not get by with this unchallenged.

Ernst Mayr and the Species Problem in Protists

In trying to answer Johnny Preer’s question, ?Whatever happened to Paramecium Genetics?, several possible answers have been suggested. One of them is that Paramecium, like Oenothera, or Lymantria, or Pisum for that matter, has performed its assigned role on the stage of genetics, and has now retreated to the wings. Despite the efforts of W. J. van Wagtendonk and his students, cultivating Paramecium in a defined medium was not an easy task, and what was done came too late for Paramecium to participate fully in the era of molecular genetics. One could even argue that the rise of Tetrahymena genetics, simply because of Tetrahymena’s dietary tractability, doomed Paramecium to a secondary role. A precedent for the displacement of an organismic technology by that of a more tractable relative is provided by the example in the ascomycetes; the pioneer Neurospora led Beadle and Tatum to the one gene: one enzyme formulation, and the Nobel Prize, but the work horse of fungal genetics in the era of molecular genetics is the budding yeast, Saccharomyces cerevisiae , that grows more conveniently in liquid medium.

The mouse genetics industry precludes the extensive use of guinea pigs.
Another explanation for the fading of Paramecium genetics, and one that I have entertained myself, finds Paramecium a victim of the Cold War and the tar-brush of Lysenko. Sonneborn was obviously concerned with being labeled a heretic, contaminated with political improprieties as were J. B. S. Haldane, J. Brachet, and some other European left-wing sympathizers. I am less inclined to give credence to that interpretation now, however. Political nonconformists - such as Richard Lewontin - seem not to be punished severely so long as their scientific work seems uncontaminated. Certainly organisms have temporarily lost scientific credibility when they have been misused for inappropriate personal or theoretical purposes. One remembers, for example, Carl Lindegren’s irreproducible studies with Neurospora, and Franz Moewus’ imaginative adventures with Chlamydomonas. In both cases the organisms recovered their utility, though the biologists who misused them have been stripped from the canonical records. To the best of my knowledge, however, no one has seriously questioned Sonneborn’s scientific integrity or doubted the reproducibility of his work.

An episode in Sonneborn’s career that I had previously paid little attention to now seems to me to be more weighted with consequences than I had earlier imagined. Sonneborn’s reputation was seriously damaged at a critical time by the claim not that his work was fraudulent, but that he was incompetent. Sonneborn was accused by Ernst Mayr, the grand arbiter of evolutionary theory, of misunderstanding the most significant generalization of modern biology - the Modern Evolutionary Synthesis. Judy Johns Schloegel (1999) has recently studied the confrontation between Sonneborn and Mayr at a symposium organized by Mayr for the American Association for the Advancement of Science on The Species Problem (1957). In her account one finds evidence of another ?struggle for authority? between the defender of a synthesis based on studies of higher organisms and the proponent of a wider evolutionary perspective from one of the earliest explorers of evolutionary processes in microorganisms. Mayr’s long life and autocratic control of historical accounts of evolution delayed attempts to deal realistically with problems of genetic economies, and levels of selection until advances in molecular evolution have forced the confrontation of the imperfections of the Synthesis (See the discussions of Mayr and Woese in the PNAS articles of 1998).

I don’t know why I failed to see the significance of the Mayr-Sonneborn conflict at the time. Obviously I was embroiled in my own professional and personal struggles at the time, and focussing on the attempt to domesticate Tetrahymena into a proper genetic instrument. I was not sensitive to the difficulties that later Sonneborn students encountered in finding positions in research universities. It was only much later, after Sonneborn’s death. that I learned of Mayr’s prejudicial handling of my own nomination to the National Academy of Sciences. I did realize the difficulties that my own students had in finding positions and obtaining grants, and finally belatedly began to unde ìrstand the pervasive misunderstanding of evolutionary processes and species definitions among the protists, still hanging over in the obfuscations apparent in Bland Finlay’s (2002) rejection of Sonneborn’s concept of multiple species in a fine-grained and evolutionarily dynamic ecosystem.

In the end, I’m afraid I can’t confidently assign a single cause to the possibly temporary eclipse of Paramecium and Tracy Sonneborn from the perview of modern genetics.Probably “all of the above” are involved, and I will leave the arguments to the metabiologists.

Bibliography

Allen, G. E. 1978.
Thomas Hunt Morgan: The Man and His Science. Princeton Univ. Press, Princeton.

Altenburg, E. 1946. The symbiont theory in explanation of the apparent cytoplasmic inheritance in Paramecium. Amer. Nat. 80:661-662.

Altenburg, E. 1948. The role of symbionts and autocatalysis in the genetics œof ciliates. Amer. Nat. 82:252-264.

Beale. G. H. 1954. The Genetics of Paramecium aurelia. Cambridge Univ. Press.

Billingham, R. E., and Medawar, P. B. 1948. Pigment spread and cell heredity in guinea pigs’ skin. Heredity 2:29-48.

Beisson, J. and Sonneborn, T. M. 1965. Cytoplasmic inheritance of the organizatin of the cell cortex in Paramecium aurelia. Proc. Nat. Acad. Sci., Wash. 53:275-282.

Carlson, E. A. 1981. Genes, Radiation and Society: The Life and Woork of H. J. Muller. Cornell Univ. Press, Ithaca and London

Chao, P. K. 1953. Kappa concentration per cell in relation to the life cyclt, genotype and mating type in P. aurelia, variety 4. Proc. Nat. Acad. Sci., Wash., 39: 103-112.

Cleland, R. E. 1972. Oenothera: Cytogenetics and Evolution. Academic Press, London and New York.

Correns, C. 1909.Vererbungsversuche mit blass (gelb) grunen und bluntblattrigen Sippen bei Mirabilis Jalapa, Urtica pilulifera und Lunaria annua. Zeitschrift fur induktive Abstammungs und Vererbungslehre 1:291-329.

De Vries,H. 1901. The Mutation Theory, trans. J. B. Farmer and A. D. Darbishire, Open Court, Chicago.

Ephrussi, B. 1953. Nucleo-cytoplasmic Relations in Microorganisms. Oxford Univ. Press.

Ephrussi, B. 1958. The cytoplasm and somatic cell variation. J. Cell. Comp. Physiol. 52:35-53.

Finlay, B. 2002. Global dispersal of free-living eukaryotic species. Science 296:1061-1063.

Huxley, J . 1953. Evolution in Action. Harper & Brothers, New York.

Jacob, F., and Monod, J. 1961. Genetic regulatory mechanisms in the synthesis of proteins. J. Mol. Biol. 3:318-356.

Jennings, H. S. 1939. Genetics of Paramecium bursaria, I. Mating types and groups, their interrelations and distribution. Genetics 24:202-233.

Kevles, D. J. 1985. In the Name of Eugenics: Genetics and the Uses of Human Heredity. Alfred A, Knopf, New York.

Kinsey, A. 1948. Sexual Behavior of the Human Male.

Lederberg, J. 1952, Cell genetics and hereditary symbiosis. Physiol. Rev. 32:403-430.

L’Heritier, Ph. 1948. Sensitivity to CO2 in Drosophila - A review. Heredity 2:325-348.

Luria, S. E. and Delbruck, M. 1943. Mutations of bacteria from virus sensitivity to virus resistance. Genetics 28:491-511.

Lysenko, T.D. 1948
. Heredity and Its Variability. trans. T. Dobzhansky, Kings Crown Press, New York.

Lwoff, A. 1923. Sur la nutrition des infusoires. Compt. Rend. Acad. Sci. 175:928-930.

Lwoff, A. 1950. Problems of Morphogenesis in Ciliates: The Kinetosomes in Development. Reproduction and Evolution. John Wiley & Sons, New York.

Margulis, L. 1981. Symbiosis in Cell Evolution. W. H. Freeman &Co. San Francisco.

Mayr, E. 1957. Species concepts and definitions. In The Species Problem, pp. 1-22, A.A.A.S. Washington D.C.

Mayr, E. 1998. Two empires or three. Proc. Natl Acad. Sci. 95:9720-9723.

Monod, J. and Jacob, F. 1961. General Conclusions: Teleonomic mechanisms in cellular metabolism, growth and differentiation. Cold. Spring Harb. Symp. Quant. Biol. 26:389-401.

Nanney, D. L. 1953. Mating type determination in Paramecium aurelia, a model of nucleocytoplasmic interaction. Proc. Nat. Acad. Sci. USA. 39:113-119.

Nanney, D. L. 1954. X-ray studies on paramecins and kappas of variety 4 of Paramecium aurelia. Physiol. Zool. 27:9-89.

Nanney, D. L. 1957. The role of the cytoplasm in heredity. In W. D. McElroy and H. B. Glass, ed., The Chemical Basis of Heredity. Johns Hopkins Press, Baltimore, pp 291-307.

Nanney, D. L. 1958. Epigenetic control systems. Proc. Natl Acad. Sci., Wash. 44:712-717.

Nanney, D. L. 1981. T. M. Sonneborn: An Interpretation. Ann. Rev. Genetics 15:1-9.

Nanney, D. L. 1983 . The ciliates and the cytoplasm. J. Heredity 74:163-170.

Nanney, D.L. 1997. Frontier Connections. Rev. Soc. Mex. Hist. Nat. 47:201-216.

Pace, N. R. 1992. A molecular view of microbial diversity and the biosphere. Science 276:734-740.

Preer, J. R. 1948. The killer cytoplasmic factor kappa, its rate of reproduction, the number of particles per cell, and its size. Amer. Nat. 82:35-42.

Preer, J. R. 1997. Whatever happened to Paramecium genetics. Genetics 145:217-225.

Roe, Ann 1953. The Making of a Scientist. Dodd, Mead & Co. New York

Sager, R. 1960. Genetic systems in Chlamydomonas. Science 132:1459-1465.

Sapp, J. 1987. Beyond the Gene: Cytoplasmic Inheritance and the Struggle
for Authority in Genetics. Oxford Univ. Press, Oxford & New York.

Schloegel, J. J. 1999. Anomaly, unification and the irony of personal knowledge: Tracy Sonneborn and the species problem in protozoa 1954-1957. J. Hist Biol. 32:93-132.

Schloegel, J. J., and Schmidgen, H. 2002. General Physiology, Experimental Psychology, and Evolutionism: Unicellular organisms as objects of psychophysiological research, 1877-1918. Isis 93:614-645.

Cambridge Univ. Press, Cambridge Schrodinger, E. 1944. What is Life"

Sinnott, E. W., and L. C . Dunn 1939. Principles of Genetics, McGraw Hill, New York.

Siegel, R. W. 1953. A genetic analysis of mate-killing in P. aurelia, variety 8. Genetics 38:550-560.

Sonneborn, T. M. 1937. Sex, sex inheritance, and sex determination in P. aurelia. Proc. Nat. Acad. Sci., Wash. 23:378-395.

Sonneborn, T. M. 1947. recent Advances in the genetics of Paramecium and Euplotes. Adv. Genet. 1:264-358.

Sonneborn, T. M. 1950. Heredity, environment and politics. Science 111:529-539.

Sonneborn, T. M. 1950. The kinetosome in cytoplasmic heredity. J. Heredity 41:222-224.

Sonneborn, T. M. 1954. Patterns of nucleocytoplasmic integration in Paramecium. Proc. 9th Int. Cong. Genetics.

Sonneborn, T. M. 1957. Breeding systems, reproductive methods, and species problemss in protozoa. In E. Mayr. ed., The Species Problem, pp 155-324, Amer. Assoc. Adv. Sci., Washington

Sonneborn, T. M. 1975 H.S. Jennings. 1868-1947. Biog. Memoirs of the Nat'l Acac. Science. V47, 143-223.

Sonneborn, T. M. , and Beale, G. H. 1949. a Influence des genes, des plasmagenes, et du milieu dans le determinisme des caracteres antigenique chez Paramecium aurelia variete 4. In Unites biologiques douees de continuite genetique. CNRS, Paris 7:25-36.

Spiegelman, S. 1946. Nuclear and cytoplasmic facors controlling enzymatic constitution. Cold Spring Harb. Symp. Quant. Biol. 11:256-277.

Sulloway, F. 1996. Born to Rebel: Birth Order, Family Dynamics and Creative Lives. Pantheon Books. New York.

Wagtendonk, W. J. van and Hackett, P. L. 1949. The culture of P. aurelia in the absence of other living organisms. Proc. Nat. Acad. Sci., Wash. 35:155-159.

Wiener, N. 1948. Cybernetics or Control and Communication in the Animal and the Machine. John Wiley & Sons, New York.

Woese, C. R. 1998.
Default taxonomy: Ernst Mayr’s view of the microbial world. Proc. Natl Acad. Sci. 95:11043-11046.