Going Back for Our Future : Carrying Forward the Spirit of Pioneers of Science Education [1 ed.] 9781623962555, 9781623962531

This book is the first volume of an attempt to capture and record some of the answers to these questions--either from th

256 112 6MB

English Pages 403 Year 2013

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Going Back for Our Future : Carrying Forward the Spirit of Pioneers of Science Education [1 ed.]
 9781623962555, 9781623962531

Citation preview

Going Back for Our Future Carrying Forward the Spirit of Pioneers of Science Education

A volume in Pioneers in Science Education Jon E. Pedersen, Series Editor

This page intentionally left blank.

Going Back for Our Future Carrying Forward the Spirit of Pioneers of Science Education edited by

Jon E. Pedersen University of Nebraska–Lincoln

Kevin D. Finson Bradley University

Barbara S. Spector University of South Florida

Paul Jablon Lesley University

INFORMATION AGE PUBLISHING, INC. Charlotte, NC • www.infoagepub.com

Library of Congress Cataloging-in-Publication Data A CIP record for this book is available from the Library of Congress   http://www.loc.gov

ISBN: 978-1-62396-253-1 (Paperback) 978-1-62396-254-8 (Hardcover) 978-1-62396-255-5 (ebook)

Copyright © 2013 Information Age Publishing Inc. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, microfilming, recording or otherwise, without written permission from the publisher. Printed in the United States of America

Contents Introduction.......................................................................................... vii 1 Fletcher Guard Watson: Setting an Agenda for Science Education...... 1 Paul Jablon 2 F. James Rutherford: An All Star Hall of Fame Science Educator...... 23 Catherine Lange 3 Paul DeHart Hurd (1905–2001): Science Educator with a Social Vision......................................................................................... 43 Barbara S. Spector 4 A Pioneer............................................................................................... 57 Senta Raizen 5 Willard J. Jacobson: 20th Century Visionary in Science Education............................................................................ 79 Rodney L. Doran and Abby B. Bergman 6 Mary Budd Rowe: What a Researcher Can Say to Science Teachers.............................................................................. 99 Julie A. Bianchini and Nicole I. Holthuis 7 A Career of Opportunities................................................................. 123 Rodger W. Bybee 8 Susan Loucks-Horsley, PhD: Transformational Leader and Spark for Educational Change.......................................................... 155 Susan Mundry

v

vi   Contents

9 Robert Karplus (1927–1990): Science Education Pioneer.............. 177 Robert G. Fuller and Beverly Karplus Hartline 10 The Nature and Development of Scientific Reasoning: My Career in Science Education....................................................... 199 Anton E. Lawson 11 A Half-Century Effort to Create a Theory of Education to Guide the Improvement of Teaching and Learning....................217 Joseph D. Novak 12 Pinchas (Pini) Tamir: A Long-Distance Runner Across and Beyond Science Education............................................. 249 Avi Hofstein, Hanna J. Arzi, and Anat Zohar 13 Shifting Paradigms in Science Education: A Change Agent’s Life on the Edge................................................................................. 269 Barbara S. Spector 14 E. Joseph Piel:Pioneering Technology in Science Education......... 297 Barbara S. Spector and Rene Goytia 15 Pioneers in Science Education: Marvin Druger............................... 309 Marvin Druger 16 Luck—A Defining Element of Success: Or, How a Few Borrowed Innovations, Time, Effort, and Money Combined with Opportunity and Support, Are Creating Success for Some Professors and Many Students........................................................... 327 John E. Penick 17 Why I Became a Teacher.................................................................... 351 J Myron Atkin 18 Don McCurdy: The “Overachiever”.................................................. 367 Peggy Tilgner Epilogue.............................................................................................. 379 About the Editors............................................................................... 383 About the Contributors...................................................................... 387

Introduction Imagine you are in a time hundreds of years ago, standing at the edge of a great expanse of prairie or forest, or on the shores of an ocean, or at the crest of a mountain. You gaze at the sight before you, simultaneously feeling waves of anticipation and anxiety wash over you. You begin to wonder: What lies beyond the horizon? How would I even get there? What perils lie in wait for me? What treasures might I gain by getting there? The thoughts are almost overwhelming as you realize few people, if any, have ever stood in that place, seeing what you are seeing, knowing the limits you know, yet willing—almost driven—to step forward into the unknown to discover what is there and what rewards can be had, to blaze trails that others will surely follow—perhaps later rather than sooner, but nonetheless, knowing (hoping) in your core that they will come. So you set out, on your own, separating yourself from the safe confines of civilization. You are a pioneer, or maybe even a trailblazer! Think about the trailblazers and pioneers. Who were they? Perhaps more importantly, who are they? Certainly, pioneers are not just the mountain men of 1700s North America, or the rugged individuals piloting their Conestoga wagons westward along what would later become the Oregon Trail, or the great seafarers of centuries ago, or those who rocket into outer space. Trailblazers and pioneers are individuals who invent new paradigms, are among the first to venture into new territory, to explore new ways of doing things, to rethink old ideas and fashion new ones—to shift paradigms that redefine and redirect life and living. The life of a pioneer cannot be an easy one. Although there is potentially great reward to be gained, there will indeed be sacrifice and angst and struggles—some certainly life-changing.

Going Back for Our Future, pages vii–xi Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

vii

viii   Introduction

What would possess someone to be a trailblazer or pioneer? What could possibly motivate someone to take on a pioneering persona and subject themselves to the unknown—to step into the vast prairie, or set sail into the uncharted sea, or trek to the next mountain? Do they do it for themselves, for others, or for some higher reason? What do pioneers actually accomplish, and how enduring are those accomplishments? Do they leave descendants, and do those people carry on the work envisioned by their pioneer forebears? In what ways, if any, does all this change the landscape (or seascape or spacescape) that the rest of us will eventually inhabit? This book is about trailblazers and pioneers who had an impact on science education in the United States. Some began their pioneering ventures before “science education” was even a discipline. Phil Schlechty, in his 1993 seminal article about school reform, uses this same analogy. He notes that educational trailblazers take “paradigm-breaking journeys that are not for the timid . . . without maps to places where no one has gone before them. Closely following are the pioneers. Like the trailblazers, pioneers are an adventurous and hardy lot and are willing to take considerable risks” (Schlechty, 1993, pp. 47–48). Who these pioneers and trailblazers were in science education is not only determined by how early in the 20th century they did their work, but more by the unique, unexplored, and sometimesdangerous paths to their careers that they were willing to endure. Although the trailblazers took the lonely and completely uncharted journey, they did not always stay and found institutions where many could live and work together. The pioneers stayed the course. They made the journey and then institutionalized practice so that the rest of the “settlers” could follow. They created new systems and approaches. Schlechty (1993) elegantly states that pioneering effective reform required, “faith, logic, wisdom, and intuition.” We can do web searches and derive lists of publications and presentations made by many of these pioneers and trailblazers. What is more difficult to do, however, is to find out what was the driving force that motivated them to do what they did in the ways they did it. How did they feel about it? What were their joys and fears along the journey? This book is the first volume of an attempt to capture and record some of the answers to these questions—either from the pioneers themselves or from those persons who worked most closely with them. We know there are many pioneers and early trailblazers who are not included in this volume, but there are other volumes to follow. We had intended to include a number of individuals in this first volume, but their chapters were not completed in time for publication. As we have posed questions, rummaged through files and oft-neglected books, and probed the memories of many individuals, we have come to realize our list of true pioneers is ever growing. There are names on the list that most of us readily recognize, and there

Introduction    ix

are names of whom few of us have heard—yet who were significant in their roles as mentors or idea development and teaching. We quickly discovered that the “family tree” showing connections between these people is not a neat, clean simple branching tree, but is more like spaghetti. The connections are many, are intertwined, and all have their significance. The stories in this volume demonstrate how vital this network was in supporting the individual pioneers during their journey in difficult times and continues to be for those of us today in our own enterprise. One of our major goals at the outset of the project leading to this book was to recover and preserve (and share!) the personal histories of the pioneers in science education before the information became lost in time, disintegrating into dust as the pioneers themselves left us—and, as it turns out, before their first-generation descendants (such as their graduate students) also left us. Like someone doing genealogical work on a family member, the documents that are available and accessible become more difficult to obtain the further time moves forward. And few of those documents can really convey to us the true flavor of the personality of ancestors. That is something best done with first-hand contact and dialogue. We, as colleagues in a profession and discipline, are now standing at the critical juncture where we can decide whether to make the attempt to record those personal pioneering histories or lose them forever. Just as when science as a discipline matured, certain universities created specialties in the History of Science, we believe it is now time for certain science education doctoral programs, in conjunction with history programs at their institutions, to create a specialty in the History of Science Education. As editors we have come to realize this historical work needs to be a full-time endeavor for some individuals and their institutions if we are to preserve an accurate and rich history along with the physical documents that can be utilized by scholars in the future. Try to imagine a Biology major who was not familiar with the work and life of Charles Darwin. Unfortunately, as we address in the Epilogue, the parallel is often all too true in our own science education profession where little is known of the work and lives of our own pioneers. How can we create an effective future if not by looking back and building it upon the shoulders of those pioneers? As editors and contributors to this book, we have obviously chosen to do what we can to not lose this heritage. We have chosen to honor our pioneers for their efforts and life contributions, and to come to know them with a more personal and familiar touch. Some of what we’ve found in our endeavors is surprising. Some brings joy, some sadness. All is enlightening. It is like finding our family legacy once again.

x   Introduction

About the Diagrams in the Book At the end of this Introduction and each chapter in this book, you will find a diagram. We thought a little explanation about them would help readers understand why we have included them. Rather than showing exact chronologies, the diagrams at the end of each chapter are intended to primarily show the connectivities or interconnections between the person who is the focus of the chapter and other individuals who influenced him/ her in some way. Pieces included in a diagram are largely drawn from and limited to information provided within the chapter narrative. We acknowledge there are likely many other connections not noted in the chapters, and consequently don’t appear in the diagrams. However, what we have included were those the authors evidently felt were most important to mention. Across the chapters as a group, we noted some names recurred more than others. To that end, we generated the diagram at the end of the Introduction. The more often the collective authors mentioned an individual’s name, the closer it was positioned toward the diagram’s center. Again, we acknowledge there are likely many connections between these individuals that were not counted in this analysis, yet we chose to limit the analysis to just that information provided by the authors because that was the most manageable and least cumbersome route to take. Rather than being a hierarchy of importance, the general purpose of the diagram is to show how the people who are the focus of chapters are networked and connected to others. And that gets us back to the purpose of the book: to show how we are all connected in some way to the pioneers who went before us. References Schlechty, P. (1993). On the frontier of school reform with trailblazers, pioneers, and settlers. Journal of Staff Development, 14(4), 46–51.

Introduction    xi

This page intentionally left blank.

Chapter 1

Fletcher Guard Watson Setting an Agenda for Science Education Paul Jablon Lesley University

Prologue It was terrific working at the Harvard College Observatory for four years despite an interruption of service in the armed forces during the war, but in 1946 it was a bit disconcerting being summoned to President James Bryant Conant’s office. Not only was Conant president of Harvard University, but he was also the man who had taken on the task of transforming Harvard into an increasingly diverse and world-class research university (Conant, 1945). For Fletcher Watson, it had been a long, but interesting road to get to this point in a career. It seemed like such a long time ago deciding on astronomy as a career. Then getting into Pomona College, graduating in 1933 with a degree in astronomy and then a PhD in astronomy at Harvard. But why did Conant want to see Harlow Shapley’s research assistant at the College Observatory and not Shapley himself? The meeting turned out to be inspiring, frightening, and life changing. Here was one of the most powerful educators in the country asking

Going Back for Our Future, pages 1–22 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

1

2   P. JABLON

a research astronomer to leave his astronomy research career and to take on the rather daunting task of leading the university, and because it was Harvard, likely the country, in creating an effective training and research program in pre-college science education. Fletcher Guard Watson took on the challenge and successfully led the enterprise for the next 31 years in the Graduate School of Education and as a national and international consultant. Never on that day could he have imagined that he would be chairing national study groups, leading of one of the most historically significant pre-college physics curriculum projects, pioneering science education research techniques that are still emulated in the profession, and mentoring countless numbers of science education leaders who have had profound effects upon the nature of K–12 science education in our country for the past half a century. The Early Years Becoming an Astronomer Born in Baltimore in 1912, Fletcher Watson attended elementary school in Los Angeles under the shadow of Mount Wilson with its great observatory. I remember when President Harding died, there was an eclipse that went over Catalina Island somewhere along the line and it was right under Mount Wilson. Somehow in a moment of enthusiasm I had decided that I wanted to be an astronomer. [It] was fourth or fifth grade; a ten year old kid. I didn’t know what it was about. It was a glamorous thing. My music teacher—I was taking piano lessons—and the other people around were—they gasped and applauded, and sent me stuff to read, so I started reading about astronomy in a very remote, distant, summarized fashion. (American Institute of Physics, 1990, p. 3)

At the time, astronomy was a seemingly exciting kind of thing because there was a lot of news in the local paper about Mount Wilson. Watson was a sickly teenager and his parents didn’t want him far from home. So after a year at Pasadena Junior College in 1929—the year that Pluto was discovered and the beginning of the Great Depression—he followed his sister-in-law to nearby Pomona College. They had a small department of astronomy, so he became the one astronomy major for the years he was there. Watson spoke about wasting a lot of his early college years and all his high school years because he didn’t make or borrow a telescope. He made no observations; he was an armchair astronomer. This all changed when his college advisor, Professor Walter Whitney, arranged for him to attend the Journal Club at the Observatory where he heard Edwin Hubble with his descriptions of the galaxies and their expansions. Each week he was the only “kid in the room” with the likes of Walter Baade, Edwin Hubble, Adriaan Van Maanan,

Fletcher Guard Watson    3

George Ellery Hale, and Walter Adams (director). As a senior, he had an opportunity to use the observatory to attempt to take spectral readings of meteors, a rather daunting and unsuccessful task (Eisenkraft, 1981). Through these meetings he formed a relationship with Seth Nicholson, who was doing research on the sun. He was a kindly man who not only helped Fletcher with his senior honor’s thesis, but also wrote a letter of recommendation for him to Harlow Shapley at the Harvard Observatory. It was during the Depression, and he was fortunate to receive a teaching fellowship while he was doing his doctorate in astronomy at Harvard. When he was finished, he stayed on as a research assistant to Harlow Shapley, one of the leaders in astrometry in the first half of the twentieth century. In 1938, he joined the Harvard College Observatory staff. Under the wing of Fisher . . . after Peter Millman had just finished his thesis on spectra meteors . . . [I began] a fair amount of observing, which was naked eye observing or running cameras in the middle of the night, usually when it got cold. And, it got colder and colder and colder. [His first time observing it was –18F.] So I had an introduction to meteor observing—naked eye observing. No one had any machinery other than cameras to do anything and I wanted some other kinds of information. (AIP, 1990, p. 7)

He had intended to work with Ernst Opik, who was doing research at the observatory on meteors, but Opik had returned to Estonia before Fletcher arrived. So he became Bart Bok’s teaching assistant for three years. Bok’s work was not on meteors, so Shapley let Watson work mostly on his own. He did interact with Willard Fisher, Dorrit Hoffleit, and Samuel Boothroyd and others gathering a reputation in the field. His astronomy career was going strong. In addition, during this time he met and married his wife Alice, then a Radcliff student who was working for Shapley, and they started a family. In 1941 he published Between the Planets. Written for a general audience, the book summarized the current knowledge of comets, meteors, asteroids, and meteorites, was well illustrated, and was translated into several languages. In a June 2008 article in Nature, Chandler cited this book as “one of the first and most important books written about near-Earth asteroids. And though only a handful of near Earth asteroids had been discovered Watson, along with Opik and Baldwin, came up with the order of magnitude correct understandings about how often a bad thing [collisions with earth] would happen” (Chandler, 2008, p. 262). In addition to this Fletcher had also published over a dozen scientific research articles. His career was well under way.

4   P. JABLON

World War II and Back to the Observatory Shortly after, during World War II, Watson served in the Navy where he assisted in the development of the Long Range Navigation (LORAN) system and achieved the rank of lieutenant commander. He returned after the war to work at the observatory where, unbeknownst to him, Shapley put his name in for a new position to be the “science educator” in the School of Education at Harvard. Simultaneously, he had been co-teaching a course for liberal arts students called “Understanding Science” created by Conant (himself a scientist) and also co-taught by Thomas Kuhn and Leonard Nash. Conant, Watson, and the other instructors of this general education physical science course, aimed it at bright people whom we knew were going to be important but we’re [sic] not going to be scientists. They would end up as possibly the president of the United States or end up in Congress; or as major lawyers, judges, bankers, and businessmen. The people who control the country, which the scientists do not. (Eisenkraft, 1981, p. 370)

They met weekly in Conant’s living room, and this started Watson thinking deeply about teaching science to general students and gave him some feeling of confidence about teaching when Conant approached him about the “science education” position. In a 1981 interview Fletcher stated that he was candid with the education faculty who later interviewed him when he said that he “knew nothing about education.” The Beginning of Science Education at Harvard He realized the incredible significance of the president of Harvard asking him to move from science to education, nevertheless he felt “I had cut off my right arm.” He still had astronomy research papers that were unpublished when he went off to war. After only three years he had been elected to the prestigious International Astronomical Union, and many of his distinguished astronomy colleagues, including Bart Bok, thought making this move would be a grave error. In retrospect, he says he was not dissatisfied with this decision to transition into education, not knowing whether his success would have continued in astronomy research. He soon had an astronomy job offer in Michigan, but Alice was in the hospital about to have their third child. They had just settled into their home in Cambridge. Alice has also said, “I’ll tell you why he did it and that’s because he likes people better than he does mountain-top strange astronomers . . . .The point was that he’s good with people; not everybody is good with people and he is. That was the real turning point” (AIP, 1990, p. 23).

Fletcher Guard Watson    5

However, at the time he felt like a fish out of water. He remembered going to his first educational meeting and thinking it was like an astronomical meeting. “You sit in the first row and you ask bright questions to show your interest and try to provoke the speaker into going further or probing the subject. So I sat in the front row and started sounding off and everybody went ‘Whoa—who is this character?’” Luckily he became acquainted with Ralph Powers, who was head of Science Education at Teacher’s College Columbia and was highly admired. He mentored him and gave him encouragement. Watson felt a little less isolated (Eisenkraft, 1981, p. 368). He began his study of science teaching by examining the effectiveness of undergraduate general education science courses that he and some select colleagues were creating and teaching, but quickly utilized these experiences by having students engage in applications of scientific knowledge to inform his study of pre-college science teaching. As he began to investigate what the situation was in pre-college science instruction and to create programs at the graduate school for both pre-service and in-service science teacher populations, he began to understand and document what these teachers needed. Conant had expected Harvard to define the mission for the country by creating exemplary programs at Harvard. Fletcher not only did that, but also through his published writings gradually began to help define the agenda of science educators in the United States for the next decade. Reading his articles from the 1950s and early 1960s, and then knowing his later accomplishments, it was as if he was creating a checklist for what he needed to accomplish. He then spent the next thirty years collaborating with people to address each item on the list. If one reads his publications in hindsight, it becomes the agenda of science education in the United States for the next sixty years. Although his publications were not the only ones addressing these issues, his were especially prescient in creating an agenda that is still being addressed even as we enter the second decade of the current century. He was probably not aware he was doing this, but rather in his usual fashion he was investigating the real needs of teachers, students, and schools, and unlike many university educators of the times, was candidly and explicitly listing the needs of schools across the country. In a number of cases he was also suggesting possible solutions, including some rather unpopular suggestions of local, state and national government expenditures for teacher salaries and science education staff development and curriculum development programs. He began by locally identifying and addressing needs in his own state. However, he soon began attending meetings of the Association of Educators of Teachers in Science (AETS) at Teacher’s College, Columbia in New York City where a handful of faculty from universities in the northeast who had an interest in training science teachers would have a yearly one-day meeting. He realized that the issues he was encountering

6   P. JABLON

in Massachusetts were universal issues and began writing in journals to a national audience. Within five years, he was organizing multi-day conferences with science education faculty from across the country that created reports to address various problems and to create suggested solutions. He investigated and wrote clearly and succinctly about the needs of pre-service programs, teacher shortages, and the lack of support and culture of professionalism for the K–12 science teacher community. Much of this was before the Sputnik era when there was little or no focus on science education in the United States. He also had an impact upon the way mentors worked with their doctoral students, moving away from the more hierarchical paradigm that was prevalent at the time to a model of being a collaborator on projects of the students’ interest. Rather than the institution defining its research agenda and all the students engaging in that research, the institution’s agenda was at least partially, if not substantially, dictated by the research agenda of its doctoral and master’s level students. As his colleague Thomas Kuhn would say, this was a real paradigm shift. In this case it was not in science, but rather in science education. General Education Multi-Disciplinary Science Courses In 1950, Watson organized the month-long Harvard Workshop “Science in General Education,” which later became a book edited by Watson and I. Bernard Cohen, a Harvard science historian. It was only four years after moving out of research astronomy into the Graduate School of Education at Harvard that Fletcher organized this conference of university science faculty and science historians from across the country to examine the best way to engage non-majors in the study of science. A series of essays that resulted from this conference became a book published in 1952 titled General Education in Science (Cohen & Watson, 1952). Although each essay in this book approached science teaching from a different perspective, each presented something significant in the sculpting of a new approach to college science teaching that utilized case histories to make science usable and humanistic. The two concluding chapters by Watson and Dyer addressed the difficult problem of evaluating the results of this new approach that is described below. This started Fletcher’s career long quest of evaluating science teaching. At this point in time, researching teaching effectiveness was an unheard of concept and an easily dismissed topic by university science faculty (Cohen & Watson, 1952). This conference, the first specific to science teaching, was an outgrowth of the General Education approach to college teaching. Ever since the General Education in a Free Society report (a.k.a. the “red book”) was put to-

Fletcher Guard Watson    7

gether by the Harvard Committee in 1945, there was much re-planning of courses and curricula in colleges and junior colleges. This report itself was not quite a trailblazing report. A good part of the underlying philosophy of the General Education approach to curriculum problems had been created out of the work of the University of Chicago some years before. Likewise, Columbia University, in planning General Education courses, also didn’t follow the lines of traditional disciplines, but included multiple disciplines in a course where they previously were usually addressed separately. Thus, this General Education multi-disciplinary approach espoused by this conference for the teaching of science was designed to give non-scientists a speaking acquaintance with science that they might not otherwise get on their own, even when studying several fields of science. Its purpose was to have the layperson become appreciative and conversant with some of the elementary science techniques and concepts. The goal would be “giving our scientifically illiterate political leaders the understandings they need to make wise decisions.” They expected this should extend to the general voting public as well. The purpose was to make nonscientists better citizens, historians, and philosophers, and especially it aimed for the cultivation of better thinking. “To have students know the scientific enterprise as a whole . . .  and also to divest science of its priestly robes. [Even in 1950,] one goal was to debunk the fictitious scientific method” (Sutton, 1952, p. 422), something we still have not managed to remove from our textbooks over a half of a century later. So here in 1950, long before the Science for All Americans, the National Science Education Standards or before the existence of the Journal of College Science Teaching is a conference convened by Watson where the proceedings are collected, edited, and summarized and where there is a plea to science educators to “replace coverage by well chosen case histories, across scientific disciplines, where through this study the student is led into scientific thinking, or a reasonable facsimile” (Sutton, 1952, p. 422). The State of Pre-college Science Teaching in the 1940s and 1950s By 1949, Fletcher became concerned enough about the state of science education in Massachusetts that he sent a questionnaire to every science teacher in the commonwealth (Harvard Crimson, 1949). From the data collected it became clear that there was already a shortage of science teachers and that more science majors needed to be attracted into this career path. More Massachusetts science teachers were in their 50s than in their 20s. Inadequate pay was the reason given for more science majors not going into teaching and for many considering leaving teaching and using their

8   P. JABLON

science degrees in more profitable ways. The salaries were so dismal that half worked in the summer and almost forty percent held an extra job during the school year. Fewer than thirteen percent belonged to a science education association, and only about fifteen percent regularly read the most popular scientific journals. Over 63% reported that one of the main difficulties with teaching was that they didn’t have enough time to prepare laboratory equipment (Watson, 1949, 1952). In the summer of 1953, Fletcher organized a month long conference at Harvard on the nationwide problems of science education in secondary schools. This led to a report published by the Carnegie Foundation entitled Critical Years Ahead in Science Teaching (Watson, 1953). This report addressed the national issues of teacher shortages and lack of facilities for the coming baby boom, the inadequate preparation of teachers, and senior science teachers without college degrees or courses in science, among a multitude of other issues. Examples of these issues were contributed by representatives who attended from each section of the country: north, east, south and west. He took his appeal about the crisis to a broader segment of the American public in a 1954 article in Scientific American. Building upon the issues raised in the 1953 report, Fletcher addressed these issues more concisely and pointedly in a 1959 academic article about difficulties in high school science teaching. Watson rued the fact that although having thousands of independent school districts allows flexibility, it is in part why educational change occurs slowly. “Some school systems react to social conditions decades before others do. Only during war do schools react quickly.” Responding to the Sputnik emergency where the United States appeared to be behind the Russians in science and technology, Watson wondered, “is the present peril enough to stimulate significant reaction?” (p. 186). He recognized that with a majority of citizens in the United States this was a larger issue beyond the teaching of science. He noted that there were still many adults who still preserved the image that there is no need for a high school education to be successful in our society. “While lip service may be given to the importance of education, we do not want to pay for what we have, let alone for what we need” (p. 187). He noted that between 1948 and 1958 installment credit buying had increased fourfold. The debt on automobiles, a status symbol at that time, exceeded one full year’s operating costs of all public schools in the nation. He stated that our citizens had their priorities reversed and that their status symbols were more important than the future of their children and of our country’s economy. He saw that few school districts were planning for the forthcoming increase in student population. The buildings were inadequate in size and quality, as were the facilities in them for the teaching of science. He further went on to question how could effective decision-making instruction in science influence these attitudes in the public of the next gen-

Fletcher Guard Watson    9

eration? He stated that the principal purpose of scientific inquiry is to gain concepts by which the behavior of things can be predicted. The creation of such concepts and their testing involves a complex of operations sometimes called “problem solving.” This leads to our future adults having an attitude that is comfortable with predictive systems and making wise judgments. If this were put into practice “a veritable revolution in teaching practices would occur in all subjects” (p. 188). He used the sports analogy that one can recognize sports achievement because one can recognize those who exceed their own. But he noted that if we were to overcome the norm of antiintellectualism in our country, then we needed to also recognize academic achievement and to cultivate it in our youth by recognizing and celebrating those who exceed our own achievement. He continued by calling for what might be interpreted as “tracking.” He noted that there was a diversity of children and what was needed was to have parallel, but distinctly different appropriate programs, not the one approach to teaching science that was used in most schools. It is only later when he began to work on Harvard Project Physics that we see that what he was then espousing was the way we approach the same high level of thinking in the teaching of science. In 1959, he called for a greater range and difficulty where the course would maintain the same basic intent; the intent being the making and testing of judgments, but with lesser and greater degrees of complexity. He suggested that in order to accomplish this, teachers needed originality, versatility, and creativity. Although he was addressing high school science courses and asked for changes in the way that students were instructed, he recognized also the need to have students better prepared and more interested in science before entering high school. He asked that K–8 teachers have better pre-service and in-service staff development so they are better informed and more secure. Likewise, he noted a need for science teaching materials in junior high schools and elementary schools. He also called for a certification of junior high school teachers who teach science. He spoke of the place of professional development in industry and how comprehensive staff development programs in science are needed in school districts for teachers, and that the teachers needed to be compensated for this additional time. Much of this was not very popular with government officials at local, state, or national levels. All of it translated into a cost for providing teachers with expertise in effective science teaching methods and knowledge. He ended by asking, “Can the United States in this grim, protracted race for survival continue to afford the insincerity and hypocrisy toward education that we have practiced in the past?” (Watson, 1953, p. 191). In 1958, Watson collaborated on a methods of teaching high school science text that was widely used across the country (Brandwein, Watson, &

10   P. JABLON

Blackwood). In 1966 he was named Harvard’s first Henry Lee Shattuck Professor of Education. Harvard Project Physics—The Curriculum, the Staff Development and the Evaluation of Effectiveness Watson served as a co-director of Harvard Project Physics, a nationwide course development effort funded by the Office of Education, the National Science Foundation, the Ford Foundation, the Alfred P. Sloan Foundation, the Carnegie Foundation, and Harvard. The project, which began in 1964, created a new physics course with a humanistic emphasis on an effort to attract all high school seniors, particularly including girls, to physics. First published in 1968, the course integrated texts, readers, measurement film loops, 16mm films, tests and new laboratory equipment, and raised the standards for what physics instruction should be for the general school population nationwide. F. James Rutherford, who was a high school physics and chemistry teacher from California with an interest in the history of science, received a scholarship to come and study with I. Bernard Cohen in the History of Science department at Harvard. Rutherford had been incorporating the history and philosophy of science in his teaching of physics to give a personalized or human side to a subject perceived as “cold” by many students. This could be accomplished by having them look at the concepts through the eyes of the scientists doing the physics. Having used his book as a reference, Rutherford sought out Gerald Holton, who was teaching Introduction to Physical Science with Ted Campbell of the physics department. It was through Holton that Rutherford became acquainted with Fletcher and the work he was doing in teaching general education physical science courses at the undergrad level, as well as the work he had initiated with high school science teachers (Rutherford, personal correspondence, March 19, 2012). Fletcher argued that all of these approaches could be adapted for pre-college teaching and thus have high school students gain the same reasoning skills and understandings of physics concepts at an appropriate level for them. Holton, Watson, and Rutherford became the triumvirate that went to work creating Harvard Project Physics (HPP), although Watson has said that, “Ted Campbell could be considered the grandpapa of Harvard Project Physics;” others might argue with this statement. Watson saw HPP as a humanistic reaction to the Physical Science Study Committee (PSSC) created by Jerrold Zacharias and others at MIT. PSSC physics was the first of the “alphabet curricula” produced in response to the Russian’s launching of Sputnik and our country’s perceived need to create

Fletcher Guard Watson    11

a massive cadre of scientists, starting in high school physics. Following the ideas in his earlier writings about the necessity of parallel approaches to the same subject matter, Watson saw PSSC physics “as an interesting and academic course for a very special clientele,” whereas HPP was created to truly engage the general high school population in understanding physics concepts and the thinking involved with creating those concepts. Watson was quoted as saying, “I was emotionally very disturbed by the arrogance of Zacharias and other people’s attitudes towards high school science teachers” (Eisenkraft, 1981, p. 369). For not only did high school students find PSSC physics a most difficult course, but a good number of teachers found PSSC difficult to teach. In 1974, Fletcher collaborated with Uri Zoller, a doctoral student from Israel, on work that was a precursor to what in the 1980s became known as the Science, Technology, and Society (STS) approach to science teaching. He extended this enterprise with international partners by creating an instrument to evaluate the STS skills (Watson, Billeh, & Malik, 1977). However, it was in the creation of Project Physics that some of the more seminal work was done. Fletcher thought that science teaching was in real trouble because science teaching was addressing only what happened in the classroom, not how that relates to their lives outside the classroom. “Not only do we wish the students to know the fact, and to know how we know that, but also what to do after we know.” This third central concern to science teaching later became known as STS. All three thought it important to look at the kinds of images, impressions, and attitudes of the general public. Before HPP, the problems in the text had nice clean numbers as answers, but they were not socially relevant. “We also had a great concern about the transformation of scientific knowledge through the swinging doors of technology into the society. What are the consequences of this?” (Eisenkraft, 1981, p. 370). Watson saw HPP as a way to generate a genuine human interest in the people responsible for doing the science and the societal context in which their knowledge was put to both beneficial and detrimental use. He saw this drawing some to becoming scientists. But more importantly, most were not going to become scientists, but rather become leaders in their own professions and communities. With HPP they would as high school students have an opportunity to practice the process skills in creating the science concepts and then a chance to practice another set of decision-making skills when addressing societal issues using scientific understandings. “Fletcher was always focused on the impact of science, not knowledge, not content, (but rather science) in the Science, Technology, Engineering and Society context” (Zoller, personal communication, April 25, 2012). HPP pioneered many effective approaches to teaching science. First, there was the integral use of original films throughout the curriculum. Then, teams worked for years to design special equipment to match newly

12   P. JABLON

developed investigations that were field tested to allow general students to uncover physics understandings. Evaluation materials for teachers that diagnosed both students’ deep understandings of concepts as well as their understanding of how to uncover these concepts accompanied these investigations. Field tests also demonstrated that thinner paperback materials were less threatening to students than one huge tome. Of course, threaded throughout all of these materials were the human struggles of the scientists who discovered how the universe worked—both their failures and successes. There are few, if any, current curricula that can match this multifaceted approach to teaching physics even in the twenty first century. These materials, except the films, have been made available at the Project Physics Course web site (Project Physics Collection, 2012). Evaluating the Effectiveness of a Curriculum Although all three co-directors worked on the curriculum materials, with Jim Rutherford leading this effort, Fletcher took on the task of field-testing the materials and then evaluating the effects of the materials on both teachers and students after the curriculum was completed. Gerald Holton recalls, “One of Dr. Watson’s proudest accomplishments was to organize the lengthy and exemplary method of testing the national program, Project Physics. [With the assistance of Wayne Welch, Herb Walberg, Jim Minstral, and Richard Brinckerhoff] he tested the materials in schools throughout the U.S. in order to have the final edition fit the needs of the various students the course tried to reach, including women students who were avoiding science” (Fletcher Watson, astronomer, science educator dies at 85, 1997, n.p.). This was a daunting task—one that Watson sculpted in a comprehensive manner not previously attempted and is still guiding present day evaluations. At each step in the development process, field test teachers and students from a broad spectrum of the American population gave feedback that allowed the developers to recraft each and every aspect of the curriculum. The summative evaluation utilized data from 54 teachers, 103 classes, and 3085 students. Fletcher made these data available to researchers throughout the country. In addition to the traditional factors of content knowledge understanding, inclination to pursue science as a career, and levels of engagement in the curriculum, Fletcher gathered a wide range of data on characteristics of teachers, students, and classrooms that were not normally addressed in science education at the time. Unfortunately, in many cases they are still not considered. Some examples include intimacy, formality, goal direction, and disorganization.

Fletcher Guard Watson    13

One study undertaken by Gary Bates, a graduate student at Teacher’s College Columbia University, and Fletcher (1976) can be used as a case study to illustrate both the unique, across-university collaboration on data from a curriculum project, as well as the other aspect of non-traditional factors considered in an evaluation. This study utilized some of the summative evaluation data to see if the interaction of teacher, student, and classroom characteristics had a significant effect on classroom climate. When researchers had observed classrooms and when students reported their observations of their classrooms, it was clear that classroom climates differed dramatically not only between different teachers’ classrooms, but also between the classrooms of the same teachers. They questioned if it was possible to predict the classroom climate based on the different pairings of characteristics of the teachers, the characteristics of a particular class, and the structure of classroom activities. As it turned out there was not much correlation beyond the obvious (e.g., content-oriented teachers scheduled less student interaction), except for what allowed HPP the possibility of being inviting to females. This was one of the first, if not the first, investigation of how to make physics inviting to girls. There were more females engaged in classes that were more intimate, satisfying, and democratic. The girls found the classes to have less friction, and everyone was less apathetic. The question that remained unanswered was to determine if the climate was caused by the presence of the girls in the class or if they were attracted to taking the class because that atmosphere existed. This was certainly groundbreaking work on attitudes, social interactions, teacher beliefs, student beliefs, and gender issues in science. Pre-service Preparation of Science Teachers and the Nature of Science Although Watson had wrestled with this throughout the previous thirty years (Watson 1957, 1965), in the 1970s he began to focus on and write a great deal about the characteristics of effective science teachers, especially physics teachers, and the pre-service preparation necessary to allow them to thrive in their own classrooms. This followed an in-depth, 133 page national study of all the science teacher preparation programs in the country that was funded by the U.S. Office of Education (Newton & Watson, 1968). They studied the characteristics of the practice teaching experience, the science methods courses, and the instructors of the methods courses. They realized that almost all of these programs reported having one methods course and most didn’t even reference all the “new curricula,” including HPP, as part of the course. After seeing that most students they interviewed in these programs thought the approach of these courses made it unlikely that they

14   P. JABLON

would ever be used in “real” schools, Fletcher began to write numerous articles and make presentations at conferences throughout the country providing in-depth explanations about why one method course and a brief student teaching experience was not going to make for effective science teaching. He went further and examined the major characteristics of the proposed second generation of science curricula for secondary schools; outlined the qualities, competencies, and teaching strategies of the “curriculum-proof” teacher; and made suggestions for future pre-service and in-service teacher training programs (Zoller & Watson, 1974). A few years later in 1980, he and Zoller would address the affective as well as cognitive needs of students in science, technology, and science education, a topic rarely addressed in science education circles (Zoller & Watson, 1980). Fletcher began by clearly addressing the course content and instructional methods desirable in teaching modern physics in the secondary school, arguing that physics must be taught in a humanistic manner to replace the desiccated, technological, dehumanized, and generally irrelevant manner in which it had been commonly taught (Watson, 1973). Amazingly, for someone who taught university science courses and was then focused on secondary science teaching, one can see Fletcher addressing how physics teachers should work with their high school students in almost the same way that Eleanor Duckworth would have early childhood teachers work with their students. Here the focus is upon the learner, the process of learning, and the conditions under which learning is enhanced. Thus the teacher’s role in interacting with students is quite different. The premise recognizes the individuality of the student and the possibility of different learning styles. Because students differ, they require a diversity of learning materials and the freedom to select among them. Direct interaction of students with materials is central. No longer does the teacher stand between the student and the phenomena. Instead, the teacher’s role is that of the mature, wise helper, or guide who is aiding the student to formulate, test, and justify his own conclusions. That learning is occurring is evidenced by a variety of changing behaviors which include student initiative, self-reliance, joy, independence, and creativeness, as well as confidence in manipulating the major concepts of the subject. (Watson, 1971, p. 504)

These ideas are still radical in the second decade of the 21st century. Secondary science educators in the forefront of their profession are still attempting to engage their colleagues in creating classrooms that model these behaviors clearly defined by Watson over forty years ago. Fletcher then followed this description of quality science teaching with an analysis of where various pre-service candidates begin when they enter, why they arrive with these perspectives, and what it would take to move

Fletcher Guard Watson    15

them to a place that would allow them to teach effectively. He noted that the behavior of a science teacher depends upon at least four factors and how s/he applies them to the classroom: orientations to science, orientations to technology, orientations to children, and orientations to the learning process. At least the first three are part of this heritage and involve systems of values that evolve throughout her/his entire life and, like all value systems, are resistant to easy change. He reasoned that pre-service teachers were unaware of their value systems. If they arrived having experienced science as a set of rules and theorems, then they would become authoritarian dogmatic teachers. He contrasted this to those who understood what science educators have now understood for the past two decades: what is now called the Nature of Science. He expected colleges to engage students in doing science so that pre-service teachers entered with this understanding. Forty years ago, he clearly defined how understanding the scientific enterprise would affect science teaching. If, however, science is viewed as a continual quest to make sense out of selected phenomena of the world, stress will be upon the interweaving of limited empirical evidence and temporary trial explanations. Emphasis will be not only on what do you know, but also how do you know it, and how well do you know it. The difficulties and errors of the scientific greats, as well as their successes, will be considered to develop the idea that no accepted concepts were easily achieved and that each man had his personal limitations. The classes of a teacher who views science in this mode will stress the inter-relations between evidence and theory, the limitations of each, and the process by which any generalizations are formulated and appraised by the scientific community. (Watson, 1971, p. 502)

Once this understanding is achieved, then many terms of observations and practice teaching in effective science classrooms in public schools supervised by clinical professors are necessary. These need to occur both in whole class and “micro” teaching situations and will lead to success once they have their own classrooms. With astounding foresight, he carefully, and in an extremely controversial manner not only for the times, but even for current practice, described the effective practice that is expected: The teacher is no longer dominant. Much of the planning is done by students, individually or in small groups. Responsibility for accomplishments rest in the students—not in the teacher. Evaluation is not primarily based on tests, and numerical problems, but also includes observed student initiative, creativeness, commitment. Achievement is considered more in terms of individual growth than in uniform standards of academic performance. There is no set syllabus, no single course, no external examination. With such an approach, the classroom becomes primarily a learning laboratory. Various investigations are preceding concurrently, a variety of instructional media—books, films,

16   P. JABLON loops, transparencies, programmed instruction as well as apparatus—are being used. Discussions, even arguments, are encouraged. Peer pressure and task orientation, rather than teacher dominance and threats, control the behavior of the students. (Watson, 1971, p. 504)

This explanation clearly delineated the foundation for the approach that is taken by some universities in the twenty first century who realized that we needed to change the format of science courses taught by science faculty, teach directly and overtly about the true nature of the scientific enterprise, and then have students experience observing and teaching in public school classrooms where science teachers are engaging in effective, cutting edge practices. After Leaving Harvard—Setting Continuing Goals for the next Century In 1977, having reached the mandatory retirement age of 65, Fletcher retired from Harvard. At this time, James Rutherford was professor of science education at New York University and was director of Project City Science, a school change project funded by the National Science Foundation and the United States Office of Education. The project involved collaboration with three inner-city school districts to change their approach to junior high school science teaching. In 1976, Rutherford received a presidential appointment to serve as the assistant director of the National Science Foundation and head of the Education Directorate. He asked Watson to take over leadership of Project City Science (Rutherford, personal communication, March 19, 2012). He accepted and was confronted on a daily basis with how difficult it was to change science teacher practice, especially in a large urban area, and how few teachers were utilizing effective methods or curricula. Watson began to write about the need for some official body to define appropriate method and content for effective science teaching. So, in 1982 in an article in the Science Teacher, he asked that the National Science Teachers Association to take responsibility for setting goals for science teaching (Watson, 1982). A year later in another Science Teacher article, he further defined the task by saying that these need to be far reaching enough to be useful to students in the twenty-first century and outlined seven suggestions in order to accomplish this (Watson, 1983a). Not coincidentally, it was Jim Rutherford who, when he became educational director of AAAS a few years later, would undertake this task with the creation of Project 2061. Despite his background in statistics in scientific research, and his transformation to the use of inferential statistics in thirty years of science education research (Watson, 1962), Watson became disillusioned with the way

Fletcher Guard Watson    17

science education research was being conducted (Watson, 1983b). Perhaps the influence of working with the inner-city teachers and seeing the irrelevance of much of the research to their daily plight led him to critique the research in a number of ways. When asked about university science educators in a 1981 interview, he commented that, Science educators are solving 1960 problems in 1980—solving them in a piddling way which is cute, but not socially significant . . . the whole problem of the use of numbers and research in the education of children is that often after the “covariant adjustments” some residual effects are looked at and maybe they are statistically significant, but not intellectually of interest. (Eisenkraft, 1981, p. 371)

Perhaps this was an early plea for ethnographic research that at the time was certainly not acceptable among science education researchers. “After the data has been massaged I can’t go back and find a real child. We’re ignoring the whole child. It may be that 5% of the invariants is due to the factor studied, but 95% of the variance is due to other things.” He had become “disinterested in the numerical research . . . .[He more appreciated] the papers when there are ideas, lines of approach; a good clean approach that will lead to explorations and judgments about possible actions” (Eisenkraft, 1981, p. 372) Mentoring and International Science Education Bart Bok remembers Fletcher as a naïve graduate student when he entered Harvard at the observatory. Fletcher reflected that Shapley, by virtue of accepting him and others as colleagues and then encouraging them to pursue their research “made them into important people.” Fletcher continued this process throughout the rest of his career with his own graduate students. Over the years, Fletcher supervised eighty-five graduate students’ theses. “These are my academic children. I hope that doesn’t sound arrogant, supercilious or some naughty word like that.” He took each of them seriously, took their interests seriously, and made them important both by supporting them and collaborating with them. As a graduate student, he had studied meteors that nobody else was much interested in and felt alone and extremely isolated. He refused to let this happen to his students and spent much personal attention on each of them. I think it is fair to say that no other scientist in this country has produced so many outstanding advocates for science literacy. He recruited doctoral students with both a strong grounding in science and an abiding commitment to education, and he prepared them for leadership. They can now be found

18   P. JABLON in this country and many others, carrying on his traditions in key positions in school districts, state departments of education, professional societies, museums, Federal agencies, colleges and universities and national and international committees. He treated all his students with great respect—which in his eyes included holding them to high standards—and closely followed their careers as long as he lived, helping them find good positions, criticizing their papers and books (when requested, which was often) and seeing that their work was known to others. (Rutherford, 1997, pp. 86–87)

Since a number of his students came from other countries, this led him into collaborating with them on work that aided the teaching of science beyond the United States. For example, in 1968 he worked alongside of one of his students creating a high school science curriculum for the Lagos schools in Nigeria (Watson & Roberts, 1968) and continued this work for over a decade with other Nigerians (Watson, Adejumo, & Ehindero, 1980). Likewise, working with another graduate student, they published a study of Arab science education (Watson & Saber Selim, 1978). In addition to these collaborations, once his work became known in the international arena, he was invited to consult on science teaching in Thailand, Australia, Turkey, France, England, Japan, and Korea. He began this work in the 1960s (Watson, 1969) and continued it through the 1980s. Legacy He was the founding chairman and a long time member of the Planetarium Advisory Committee of the Boston Museum of Science. Watson was a member of the American Academy of Arts and Sciences, the American Association for the Advancement of Science, the National Science Teachers Association, and the American Association of Physics Teachers. His numerous awards include the National Science Teachers Association distinguished service citation in 1972, the Outstanding Science Teacher Award from the Massachusetts Association of Science Teachers in 1977, and the Robert J. Carlson award for his “outstanding national leadership in science education” from the National Science Foundation in 1985. Despite all of these accolades, at the age of 78 Fletcher reminisced about his career in science education. He contrasted it with his possible career in astronomy if he had continued with that: Looking back on about 30 years in education, I feel very much like things went wrong. I had responsibilities and students and all of this, but I wasn’t creating something; continually creating. There was no center; there were individual activities here and there and everywhere like fireworks going on. I was isolated; I was the only one [science educator at Harvard]. I never did

Fletcher Guard Watson    19 find an intellectual center for what I was trying to do. That is an awful thing to have to be saying after all this time mucking around at it, but I was always doing things but they never piled up. I don’t know—a contrast between two kinds of subjects. In astronomy you had a feeling of reality and continuity. You did something and you did; you published a paper and people criticized and reacted to them, or they didn’t—but you were doing something that was adding up. You were going to get somewhere. In education I haven’t felt that . . . .I got lots of awards—which was very kind—from other people—all those things all behind you. I guess I am basically a nice guy; at least they thought so. I don’t feel—to me—an accomplishment about the—I have a long biography of papers in education but they are argumentative mostly, suggestive commentaries but they don’t add up to anything. (AIP, 1990, pp. 25–26)

Uri Zoller disagrees, “Being perhaps the last graduate student of Fletcher (1971–1973) and [presently still] an active researcher in both science education and organic chemistry, I strongly believe my science education perspective is a research-based continuum, as was Fletcher’s (Zoller, personal correspondence, April 12, 2012). In writing this chapter, I have identified more than 75 books, journal articles, book chapters, national and international reports and studies, and many curriculum materials that he authored or co-authored. I only wish Fletcher could see his work as I see it writing twenty to sixty years later. It is such irony that despite his inability to see it as a substantive, coherent body of work, it was his profound thinking within these writings and in his presentations that helped set the storyboard for science education for decades afterwards. We are still rewriting and playing out the script he unwittingly has supplied for us. In so many ways, he was a visionary—one who was candid in his thoughts that he knew would be considered radical by some, and would ruffle not a few feathers. I only met him briefly once or twice about 35 years ago when I was a graduate student. If I had only known then what I know now, I would have made it a goal to try and get to know him and work with him on some project sometime during my career. His insights were brilliant, his writing clear and concise, his energy boundless, and his goals were lofty. They are best expressed through one of his statements: My manifesto—every child should study science every day of every year at school.

References American Institute of Physics (AIP). (1990). Interview of Fletcher Watson by Ron Doel on November 20, 1990. Niels Bohr Library & Archives, American Insti-

20   P. JABLON tute of Physics, College Park, MD. Retrieved from http://www.aip.org/history/ohilist/28591.html Bates, G., & Watson, F. G. (1976). Predicting learning environments from teacher and student personality. Washington, DC: ERIC Clearinghouse. Brandwein, P. F., Watson, F. G., & Blackwood, P. (1958). Teaching high school science: A book of methods. New York, NY: Harcourt Brace. Chandler, D. (2008). The burger bar that saved the world. Nature, 453(26), 1165– 1168. Cohen, I. B., & Watson, F. G. (1952). General education in science. Cambridge, MA: Harvard University Press. Conant, J. B. (1945). General education in a free society. Cambridge, MA: Harvard University Press. Eisenkraft, A. (1981). Talks with great teachers: Fletcher Watson. Physics Teacher, 19(6), 366–373. Fletcher Watson, astronomer, science educator dies at 85. (1997, May 15). Harvard University Gazette. Retrieved from http://news.harvard.edu/gazette/1997/05.15/FletcherWatsonA.html Harvard Crimson. (1949). State needs new science teachers, Watson reports. Harvard Crimson, May 27, 1949, retrieved from http://www.thecrimson.com/article/1949/5/27/state-needs-new-science-teachers-watson/ Newton D. E., & Watson, F. G. (1968). The research on science education survey: The status of teacher preparation programs in the sciences, 1965–1967. Cambridge, MA: Harvard University. ERIC Document ED 025 435. Project Physics Collection. (2012). An electronic collection of the Harvard Project Physics curriculum materials curated by F. James Rutherford. Retrieved from http://www.archive.org/details/projectphysicscollection Rutherford, F. J. (1997). Obituary: Fletcher Guard Watson. Physics Today, 50(11), 86–87. Sutton, R. M. (1952). Book review: General education in science. Science, 115(2990), 422. Watson, F. G. (1941). Between the planets. Philadelphia, PA: Blakiston & Co. Watson, F. G. (1952). Science teaching in the secondary schools. Science, 116(3010), 261–263. Watson, F. G. (1953, July/August). Critical years ahead in science teaching. Report of conference on nationwide problems of science teaching in the secondary schools held at Harvard University, Cambridge, Massachusetts, July, 15 to August 12, 1953. Watson, F .G. (1954). A crisis in science teaching. Scientific American, 190(2), 27–29 Watson, F. G. (1957). Course requirements for future science teachers. The Scientific Monthly, 85(6), 320–323. Watson, F. G. (1959). Basic difficulties in present high school science teaching. Daedalus, 88(1), 186–191. Watson, F. G. (1962). Toward effective research in science education. Theory into Practice, 1I, 277–283. Watson, F. G. (1965). To the Council on the Education of Teachers of Science. The American Biology Teacher, 27(1), 11–17.

Fletcher Guard Watson    21 Watson, F. G. (1969). Is humanistic science appropriate for a developing country? Paper presented at the Rehovot Conference on Science and Education in Developing States, Jerusalem, Israel. Watson, F. G. (1971). Pre-service pedagogical formation of physics teachers. Physics Teacher, 9(9), 501–505. Watson, F. G. (1973). The teaching of physics . . . Science Activities, 10(3), 35–37, 45. Watson, F. G. (1982). The road to take. Science Teacher, 49(8), 31–33. Watson, F. G. (1983a). On the drawing board: A 21st century curriculum. Science Teacher, 50(3), 62–63. Watson, F. G. (1983b). Science education: A discipline? Journal of Research in Science Teaching, 20(3), 263–264. Watson, F. G., Adejumo, D., & Ehindero, S. (1980). Facilitating learning of scienceoriented textual material in a developing country: Study in the use of organizers. Science Education, 64(3), 397–403. Watson, F. G., Billeh, V. Y., & Malik, M. H. (1977). Development and application of a test on understanding the nature of science. Science Education, 61(4), 559–571. Watson, F. G., & Roberts, D. (1968). Planning a basic science curriculum for the Comprehensive Secondary School at Aiyetoro, Western Nigeria. Lagos: Ford Foundation Program in Pre-vocational and Technical Education in Nigeria. Watson, F. G., & Saber Selim, M. (1978). Recent developments in Arab science education. Science Education, 62(1), 119–124. Zoller, U., & Watson, F. G. (1974). Technology education for non-science students in the secondary school. Science Education, 58(1), 105–116. Zoller, U., & Watson, F. G. (1980). Evaluation of affective and cognitive-affective domains in socially-oriented innovative science and technology curricula. European Journal of Science Education. 2(4), 339–351.

22   P. JABLON

Chapter 2

F. James Rutherford An All Star Hall of Fame Science Educator Catherine Lange Buffalo State College

Introduction It is not easy to trace a person’s life story in such a manner that will accurately capture the essence of the person, along with the driving ambitions that collectively become absorbed into the intellectual history of science education. There are many ways that individual contributors have earned their way into the All Star Hall of Fame of Science Education. Some have contributed directly in explicit ways to the body of theoretical knowledge that now is the framework and/or body of considered acceptable “best practice” or methodologies of learning and instruction. Some have been pioneers in the creation of exemplary curricula that blend theory and practice or are innovative. For others, their notoriety is associated with pivotal research in the learning and teaching of science. Some, such as F. James Rutherford, have accomplishments that cross over several of the above-mentioned categories. For Rutherford, the story of his accomplishments includes many leadership roles, including but not limited to: Going Back for Our Future, pages 23–42 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

23

24    C. LANGE

• the design of a well-known curriculum, Harvard Project Physics; • organizational Leadership (The United States Department of Education, National Science Foundation, National Science Teachers Association, National Association of Research in Science Teaching, and the American Association for the Advancement of Science); and • the development and guidance of science education policy. The sum total and range of his activities over the course of his lifetime are impressive and considerable in scope and influence and deservingly place him among the most influential science educators of the twentieth century. The development of this chapter on F. James Rutherford was guided by Edel’s criteria: 1. Understanding your subject’s way of thinking and dreaming; 2. Remaining as a participant-observer; staying disengaged with the subject; 3. Analyzing the materials associated with the subject to discover deeper truths; and 4. Finding the right literary form to capture the subject properly (Edel, 1984, pp. 28–30 ). In order to understand Rutherford’s “way of thinking and dreaming,” traditional qualitative methods were followed. Several tape-recorded interviews, personal communications, and analysis of some of the materials associated with Rutherford were used, including, but not limited to: 1. Harvard Project Physics (HPP) curriculum and associated research publications analyzing the effectiveness of the project; 2. The American Association for the Advancement of Science (AAAS) publications (Science for All Americans and Benchmarks for Science Literacy and associated supplemental materials); 3. Peer reviewed and invited publications by Rutherford; and 4. Other sources such as AAAS newsletters, news releases, a website hosted by Rutherford, and an unpublished autobiography. Foundational Years F. James Rutherford was born in Stockton, California on July 11, 1924. Like many other families of the Great Depression, Rutherford’s parents were struggling to survive. Rutherford described the hardships:

F. James Rutherford    25 Watching my father nearly destroyed by not being able to get work, standing embarrassed with my mother in a bread line, myself working for penny tips by washing car windows in gasoline stations, my parents moving from place to place trying to land where work opportunities were better (which for me meant changing schools frequently), living in houses crowded with relatives and friends of my parents in the same situation, and so on, had to make a difference in my life. But in the Depression, that was a common story. (J. Rutherford, personal communication, 8/21/2008)

As a consequence of difficult economic times and a migratory life, Rutherford attended five elementary schools in three cities and three high schools, finally graduating from Kern County Union High School in Bakersfield, California, in 1942. Rutherford did not have one key event that triggered him to consciously pursue a teaching career, but instead there were a series of events that drew him into the field. A first experience he recalled was in the 7th and 8th grades at St. Francis de Sales elementary school in Oakland, California. His math teacher, who was also the principal, was frequently out of the classroom. During those times she appointed student assistants to grade papers and to provide tutoring to those having difficulties. Rutherford liked being an assistant because he “found it a challenge both to figure out why students were getting the problems wrong and how to help them learn to get it right” (J. Rutherford, personal communication, May 31, 2012). This experience stayed with him throughout his career because he realized that understanding something was not sufficient for teaching it well. Another poignant memory includes two very good, but different, math teachers that Rutherford had. The first one (Monterey High School) was traditional and extremely rigid. The second teacher (Kern County Union High School) had a much different approach. He elaborates; The trig teacher by contrast, formed us into groups for actually surveying the campus (which was huge and uneven). We had real surveyor’s notebooks and actual professional surveying instruments, and spent 3–4 days each week out of the classroom. We were all graded at the end as a team on the quality of our work (we did well), plus each individual’s final trig test score. The point is that while I knew from experience that some teachers were better than others (at least for me), I came to believe that there is no one best approach to teaching. (J. Rutherford, personal communication, May 31, 2012)

Rutherford also learned a great deal from a public speaking class he took at Kern. For his speech final he gave a 15 minute speech on the origin, discovery, and structure of the living cell. He received a “B” in the course and was told he did two things wrong: crammed too much material into the speech and used too many technical terms. He recalls the meaningful lessons of the experience, “I first realized that to just tell people things did

26    C. LANGE

not necessarily lead to understanding—their knowledge and capabilities much be taken into account. Also, I developed confidence in my ability as a speaker” (J. Rutherford, personal communication, May 31, 2012). It was at Kern that he met Mr. Harris, who taught him chemistry and physics. This influential teacher recognized Rutherford’s interest and aptitude in science and helped him to get a summer job at the University of California Radiation Laboratory. During his senior year of high school, the Japanese attacked Pearl Harbor and within weeks of the event, recruiters came to Kern. Rutherford joined the Navy, even though he had only been on a boat once—a ferry in San Francisco! His mother was determined that he would become a medical doctor, and so he began studies at the University of Berkeley as a pre-med major. The Navy allowed him to begin studies at U.C. Berkeley in an accelerated program, which required taking 21–23 credits per semester. He recalled that “it was not a time to think about career, but a time just to get through it with decent grades” (J. Rutherford, personal communication, May 31, 2012). Two years later, in 1944, he was sent to midshipman’s school in Chicago and then to Florida to learn how to use radar in naval combat. He also went to Pearl Harbor to practice interception skills with night fighter pilots. Following this preparation, he was placed in charge of training and advancement of enlisted radar personnel on an aircraft carrier in the Pacific Theater. Rutherford described his first experiences with the training process: I quickly found out, to my considerable embarrassment, that teaching them “textbook” physics and testing them in the college fashion had little impact on their ability to operate radar. Indeed some of the most adroit among them scored lowest in their knowledge of the pertinent physics and mathematics of radar. Privately, a senior enlisted man (11 years older than me with more than two years of combat duty) took me aside and coached me on how better to carry out my educational responsibilities. (Rutherford, 2007, n.p.)

The pedagogical modifications Rutherford applied had a significantly positive impact on his effectiveness as an instructor. He also became wary of the notion that “scientific knowledge is a necessary precursor to the effective use of science-based technologies” and he “began to wonder if it was not the case that studying science and its relevant technological applications simultaneously was often not better than studying them in isolation from each other” (Rutherford, 2007, n.p.) This experience in teaching radar helped Rutherford to see himself as a science teacher rather than a science researcher or physician. Nonetheless, he returned to U.C. Berkeley after the war and completed a bachelor’s degree in biochemistry in June of 1947. Shortly after graduation he worked for a large chemical company that produced and sold pharmaceuticals. He first worked in the laboratory, then as a writer and as a specialist who

F. James Rutherford    27

worked with physicians. He was not happy and still at odds with his career and recalls a book that profoundly changed his direction: One day . . . I came across General Education in a Free Society, a book destined to point me toward my life’s work. James Bryant Conant, the distinguished scientist and president of Harvard University, established a faculty committee to inquire not primarily on collegiate education but “into the problems of general education in both school and college . . .” (p. 156). The resulting report became widely known simply as the red book. (Rutherford, 2007, n.p.)

His physics advisor at Berkeley did not share the same enthusiasm for education and claimed there was no good reason for Rutherford to settle on becoming just a science teacher when he could be a physicist. Still struggling with the next career step, his former commanding officer at Pearl Harbor encouraged him not give up his desire to teach and suggested that he investigate Stanford. This proved to be good advice. At Stanford, Rutherford experienced a satisfying and fertile foundational environment where he thrived, growing in many ways including his inclinations to think deeply about humanistic science education approaches. He graduated in 1949 with a master’s degree in science education. His thesis was entitled, Status of Teaching Atomic Energy in California. By this time, he was married to his wife Barbara (Webster) and had a daughter, Constance (born in 1946). With a wife and a growing family, he was ready to seek employment as a science teacher. In 1949, Rutherford began teaching algebra, general science, chemistry, and physics at South San Francisco High School. The school was attended by children of the workers in the nearby steel, railroad, and meat packing industries. The parents of the children were mostly Greek, Italian, and Mexican second-generation immigrants. They were very grateful that their children had an opportunity to receive an American education. Rutherford recounts fondly of the experience: The families were different then, the Italian parents would send you presents to school, including bottles of wine, and they would invite you on graduation day to a party at their house, grandparent, parents, it was fun! (J. Rutherford, personal communication, October 11, 2001)

District policies were another matter. Rutherford became upset when he found out that some teachers in the district earned less because they taught elementary school or because they were female, or both. Rutherford recounted the situation with conviction: The teachers were poorly paid and elementary were paid less than high school teachers for the same degrees and years’ experience and the women,

28    C. LANGE somehow even in high school, who otherwise had equal degrees and were not paid as much. (J. Rutherford, personal communication, October 11, 2001)

Much to the chagrin of the superintendent, he organized a chapter of the American Federation of Teachers (then AFL-CIO), and was elected president. The union, under his leadership, succeeded in obtaining a unified salary schedule for all of the teachers of the district. His colleagues warned him against such activism, because he did not have tenure and would lose his job. Later though, he recalled the outcome: You know it was just the opposite, I was too visible and it would look vindictive and sure enough, the superintendent made sure, called me in and hated to have to offer me a job for my third year. (J. Rutherford, personal communication, October 11, 2001)

He declined the offer from South San Francisco in order to join the San Mateo School District as part of an innovative educational experience called Capuchino in San Bruno, California. This position appealed to him because he was part of an 11-member team that was charged with designing a curriculum, developing innovative courses, working with an architect in the design of the facilities, and participating in the of hiring new faculty as the school grew. It was quite unusual for a school to use a shared-decision making leadership model during the 1950s. He spent the next twelve years teaching high school science at Capuchino. There were several accomplishments and events during his public school career that stand out as exemplary and include: 1. He was hired as department chair of Science (had only 3 years of teaching experience in 1951), 2. he received a Ford Foundation Grant to study as a doctoral student in the History of Science at Harvard (1954), 3. he began work on Harvard Project Physics (1960), 4. he received his doctorate from Harvard (1961), and 5. he received a Carnegie Foundation Grant to develop a program that would foster interactions between science and the humanities in the high school (1962). Rutherford felt that he had to develop his own materials to effectively teach courses in South San Francisco and at Capuchino. With his wife Barbara’s assistance (and the mimeograph machine), he produced materials that were, in essence, a curriculum for his courses. These included the history and philosophy of science because he believed that lessons blending science content and its history are at the core of understanding the nature of scientific achievement. He also felt that he needed to make his courses more interest-

F. James Rutherford    29

ing and humanistic to appeal to a wider range of students. Through his own readings and creativity, he was building interesting and meaningful stories and examples that would substantiate the existing content. He could only do so much on his own, however, and in 1954 he felt the need to return to school for a deeper understanding. He explained: I began inserting this stuff, into the physics. I decided that I needed to know more history. I didn’t know enough history, and applied for and to my surprise, was awarded a Ford Foundation Scholarship to Harvard University to study the History of Science in the Graduate School of the History of Science. (J. Rutherford, personal communication, October 11, 2001)

Even though Rutherford had come to Harvard with experience as a public school science teacher, his advisor, I. B. Cohen, in the history of science department, could not understand his dual interest in science education and science history. Recognizing this problem, Rutherford began to gravitate to the education department where he met his new advisor, the wellknown science educator, Fletcher Watson. Watson assisted Rutherford in his efforts to apply theory to the curriculum project that was driving him forward with a strong purpose. In 1956, Gerald Holton, a physics professor and science historian, published Experimental Physics: A Laboratory Manual for the Introductory Physics Course, which was similar to the work that Watson and Rutherford were doing, and subsequently, by 1960, the three men began working together on Harvard Project Physics. This dynamic group formed a strong triumvirate. This alliance profoundly shaped the lives of each contributor as the project gained interest from science educators, physics teachers, professors, and science historians around the world. Rutherford completed his doctorate in science education in 1961, and his dissertation was titled An Analysis and Evaluation of Policies and Practices in the Selection, Training and Employment of Science Teachers. The research and learning that was involved in completing the dissertation deepened his knowledge and understanding of the issues that science teachers face. Rutherford has put great value on the dissertation, claiming, “That publication, now 44 years old, still influences my thinking about graduate education in science education” (Rutherford, 2007, n.p.). He was the first to be awarded the Distinguished Dissertation Award from the Harvard Graduate School of Education, a most notable distinction for a high school science teacher. Rutherford returned to the classroom at Capuchino in 1961 to fulfill his obligation to the district for the time he had taken during the sabbatical. Nevertheless, he found it difficult to work on the innovative materials, to teach, and to be a father to his four children. Consequently, he applied for and was awarded a $54,000.00 grant from the Carnegie Foundation of New York, for the Science Humanities Project. This gave him the freedom to

30    C. LANGE

direct a district-wide, seven high school testing site from 1961 to 1964 for the newly written curriculum materials. The amount of the grant was a fortune for a teacher in 1961 and allowed Rutherford to hire an assistant who worked in a makeshift office in the attic of the school. The grant funded collaboration between science and history teachers to increase teacher effectiveness in the delivery of lessons that connected science and technology to society. Rutherford was disappointed when the funding ended along with teacher interest in the project. Reflecting on the program, he stated, “although many school and university educators favor the idea of crossdiscipline studies, at least in principle, they are extremely difficult to pull off” (Rutherford, 2007, n.p.). Harvard and Harvard Project Physics Harvard Project Physics (HPP) was one of the alphabet (or ABC) curricula produced as a national response to the Russian launch of Sputnik. (Curricula were labeled such because of the many acronyms that identified the Science, Technology, Engineering and Mathematics [STEM] projects.) Welch (1979) estimates that by 1977 there were “more than 500 different projects” used in the United States high schools, and by 1972 Holton estimated that “300,000 students per year were using Harvard Project Physics” (Holton, 2001). Overall about twenty of these projects were widely used, and only one, Biological Sciences Curriculum Study (BSCS) has survived and is still published and still in wide use as of this writing. While Rutherford was a graduate student at Harvard, he approached Holton and asked him if he would write a version of the book Experimental Physics; A Laboratory Manual for the Introductory Physics Course for high school students. Holton responded, “Why don’t you do it?” (Holton, 2001). They agreed that Rutherford would attempt a revision with Holton monitoring the process, “and that would have been the end of it” (Holton, 2001). But Holton, called to an emergency meeting at the National Science Foundation shortly after the release of the Soviet spacecraft, Sputnik, now had a fortuitous opportunity. He explained: We were implored by the NSF officials to throw ourselves, individually or in groups, into the awesome task of designing, writing, testing, re-editing and finally publishing a national high-school physics course. Everyone at the meeting was sensible enough to say “no.” Except one. That’s how I became the principal investigator of what we first called Harvard Project Physics. (Holton, 2001, n.p.)

Holton saw an opportunity to develop a “humanistic, historically orientated course” in physics that was not just “one damned thing after another,

F. James Rutherford    31

but a coherent story” made from the “thoughts and work of living beings” (Holton, 2001, n.p.). Holton pursued Rutherford and Fletcher Watson to join the project. The resulting “affluent” and “ambitious” curriculum was incredibly varied and extensive, even by today’s standards (Holton, 2001). A supporting staff of over one hundred and fifty people ran the project with one hundred and twenty advisors. Rutherford served as the executive director and senior author and editor of the project until 1971. HPP was so massive that a 42-page catalogue—as thick as a Sears and Roebuck catalogue—was needed to order HPP materials (Holton, 2001). Eventually large publishing companies took over the marketing and production of some of the ABC curricula. Today, teachers and districts have grown to expect the large-scale packaging of textbooks and supporting paraphernalia, but at the time, the HPP production was definitely groundbreaking. HPP stood as one of the most notable projects of the post-Sputnik era because of its widespread use and because of the many products that were available for teacher and student use. HPP was also well known for the excellent summer institutes that Rutherford organized. Holton elaborated: But the key for most such teachers was to take a paid-for leave to go to a six- to eight-week summer institute at one of the many teacher-training sites in various parts of the country, which we organized for many years. Thousands of teachers went through those—great for them, but as you can imagine, an additional burden above all on Jim Rutherford, who acted tirelessly throughout the project as its Executive Director. (Holton, 2001, n.p.)

The high quality of the curriculum was also part of the appeal. Rutherford hired designer Albert Gregory to incorporate historic depictions, photos, and artwork to the pages of the materials. No expense was spared to include authentic historic documents such as Copernicus’s De Revolutionibus, obtained directly from the library at the University of Krakow. Teacher participants attended fully paid six to eight week summer workshops that showed them how to use the materials. As the scope of the project grew, it soon became necessary to increase the inner circle of professionals needed to drive the curricular engine. One of the most important associates that joined the team was Andrew “Chick” Ahlgren, a high school physics teacher. Ahlgren, who became a lifelong friend and colleague of Rutherford, was “incredibly inventive and productive and he deeply understood, agreed with and acted on the humanistic philosophy” of HPP (J. Rutherford, personal communication, 6/2/2012). Twice during Rutherford’s tenure as project director, he faced pressure by project team members to remove Chick from the team he was working on because “he had unusual work habits, and he frequently seemed to offend some of the others on

32    C. LANGE

the project” (J. Rutherford, personal communication, 6/2/2012). Despite the pressure, Rutherford recognized Ahlgren’s outstanding contributions as the more important issue. As the project continued over the years of its development, an important feature of the development was feedback that science teachers offered, which often resulted in modifications to the materials. Participating teachers felt distinguished, and in many schools, things were noticeably different. Peter Dow, curriculum historian and contributor to Man: A Course of Study (MACOS) a cross-disciplinary and cross-cultural social studies project for middle school and upper elementary grade describes what he noticed in his school: As a young teacher just beginning my career at the time, I vividly remember the impact of these innovative programs on the classroom. I first learned about the NSF’s efforts when physics students in my school began building ripple tanks to examine wave motion and swinging pendulums from the rafters of the gymnasium to study the rotation of the Earth. This was very different from the textbook-based learning I had encountered in my own education. (Dow, 1991, p. 3)

In realistic terms, Dow’s experience supports what McCormick indicates were “pockets of excellence where well-trained teachers conducted wonderful new NSF science programs” (McCormick, 1992, p. 18). Teachers should be the decision-makers and implementers of new ideas, and they “determine how much value to attach to what they already do, how much changes will help their students, and how much energy and time they can invest to make the changes, given the organizational and personal constraints that they face daily” (Cuban, 1993, p. 239). Studies conducted by Glanz (1979) and Kyle (1985) are consistent with many other findings that suggest teachers involved in the curricular surge had changed their philosophies and approaches remarkably little. Rutherford understood what teachers faced in the classroom from his own experiences and focused a great deal of attention to the construction of materials in HPP so teachers would be able to easily add what they wanted or needed to their existing classroom lessons. Science education researchers differ on the impact of the reform era and specific effectiveness of individual projects such as HPP. There is agreement that the most reviewed and successful high school science projects were BSCS, Physical Science Study Committee (PSSC) and HPP. Despite his crucial involvement, Rutherford is critical of the overall success of HPP and the many other ABC projects of the era. He claims that science education did not mature in the post-war years to become all that it could have. Why? “Indeed much of what passed for research were short-term assessment of courses, materials, approaches, and projects having more to do with jus-

F. James Rutherford    33

tification than with the advancement of knowledge” (Rutherford, 2005, p. 375). Rutherford points to a lack of persistence that would have created a steady and thus enduring pace for successful reform and indicates that there were immature “exuberant ups and disconcerting downs” throughout the past fifty years of science education (Rutherford, 2005, p. 375). Notwithstanding such critical assessment, the unprecedented scope and scale of the post-Sputnik projects are in themselves worthy of historical note. Unfortunately, many of the grant-funded projects lacked effective summative evaluation processes and procedures, making it difficult to determine how much students learned from the curricula. There was little or no attention paid to teacher attitudes and/or pedagogical implications of the new materials during the development and evaluation processes of the projects despite the disproportional amount of research and development resources directed towards teacher training. Following the release of the ABC curricula, throughout the late 1970s and early 1980s, many researchers sought out ways to assess overall changes in science classroom instruction. Myriad qualitative and quantitative data emerged. One of the compelling attempts to determine the overall impact of the ABC curricula was the 1977 NSF survey that collected data from 7,000 teachers, principals, central office personnel and state supervisors and officials (Stake & Easley, 1978). The NSF 1977 National Survey of Science, Mathematics and Social Science Education used the percentage of program utilization by districts as a marker for determining the success of the programs. This report found that HPP was used by 12% of the school districts in America. Over sixty studies of the effectiveness of HPP were evaluated by Wayne Welch, and he determined that the most salient features of HPP were the contributions it made to “retention in science, participation of women. . . . [Performance] on critical thinking tests and understanding of subject matter all showed improvement where the Project Physics curriculum was adopted” (Matthews, 1994, p. 6). Fletcher Watson published over sixty articles, research papers, and reports and directed fifteen doctoral theses on HPP (Holton, 2001). Rutherford added many examples of pure science into a collection of HPP readers that contained relevant stories written by famous scientists. The immense supporting materials served the single purpose of deemphasizing the textbook by offering teachers laboratory ideas and activities to promote discovery and inquiry learning. Rutherford directed the development of all HPP materials, and by the project’s completion in 1970, it was a massively expansive and comprehensive program of textbooks, worksheets, overhead transparencies, tests, readers, full-feature films, and teacher and student resources too numerous to mention. The program offered teachers content and resources that they would otherwise not have easy access to such as historic experiments, literature, assessment tools, and depictions.

34    C. LANGE

Additionally it provided student-friendly information about cutting edge scientific research projects that average kids could read and understand. Its success was partly due to the comprehensive and interdisciplinary nature that connected science to a wide range of domains. Part II Mid Career New York University In 1970, Rutherford accepted a position as chair of the Science and Mathematics Department in the Graduate School of Nursing and Education at New York University. He taught science education seminars, history of science, sociology of science, public understanding of science, and science and technology to graduate students. There were several important lessons that Rutherford learned about systemic reform and university politics in this phase of his career. When Rutherford took over as chair of the Graduate School of Education of NYU, he found it troubling that the future teachers were receiving their science subject matter preparation from education faculty. He set out to restructure the program so that scholars in the fields of chemistry, physics and other sciences would teach the students. It was not an easy sell, and his plan fell short of fulfillment. That was not going to be the only challenge Rutherford faced during this point in his career. He put his energy into a systemic reform effort of the New York City School District, called Project City Science. A plan to transform science teaching in inner city junior high schools, along with a clinical PhD program, became problematic on several fronts. The first big setback occurred when awarded funds were frozen due to budgetary restrictions applied by the Nixon administration. The grant team rewrote the grant, resubmitted it under a different program and, after a year’s delay, were ready to implement the project. Its long-term goal was to bring significant and improved science education to the junior high schools of New York City. Further difficulties were encountered when one of the school districts dropped out of the program following a misunderstanding with one of the graduate students from NYU. There were internal issues with administrative and teacher buy-in and trust. Paul Jablon, a young science teacher who had a small amount of prior experience redesigning science instruction and curriculum at a New York City high school, was recruited to join the project as a doctoral student and research assistant. Jablon had a good working relationship with the science teachers who were targeted to participate, and

F. James Rutherford    35

he understood the infrastructure of the schools involved. The junior high schools in Brooklyn in the early 1970s were a mess, Jablon explains, These schools were seen as the most difficult of all the schools in the city, and in many cases the science teachers had the least training of all teachers. Many of them were teaching out of license. The schools did not have a lot of science supplies, and there was very little leadership within the school building or within the science faculty. In addition, New York City had a historical tradition of having its own curriculum. Even separate and different from the rest of New York State. So to try to create a program that would change the way that teachers would be teaching in junior high schools, in New York City with many out of licensed teachers, was certainly a heroic undertaking by Rutherford. (personal communication, 03/02/2012)

The lack of commitment of school administrators to Project City Science struck Jablon during the first meetings with the principals of the junior high schools. In fact, he states that “they [principals] had no intention of buying into what we were trying to do, nor did they really understand it, and hadn’t any intention of going through upsetting what their teachers were doing” (Jablon, personal communication, 03/02/2012). Jablon was frustrated because from his view of the situation, he could see that this was going to be a big problem, and since he wanted the project to be successful, he expressed his concerns to Rutherford and other senior members. Rutherford had extensive experience as a project director and as a school teacher, but he had little knowledge of the insurmountable urban issues that are pervasive and widespread in such a system. But in the end, these cited problems and several others could not be overcome by Rutherford and/ or his dream team of advisors that included Mary Budd Rowe, Leo Klopfer, Audrey Champagne, and others, and the project failed to accomplish its goals. Jablon indicates it was the lack of understanding of the uniqueness of the large and complex system such as New York City as well as a lack of any other team members with experience in inner city urban science teaching or administration that was at the heart of the failure of the project. During this period Rutherford was elected president of the National Science Teachers Association and held that office from 1974 through 1976. Rutherford’s attempt to facilitate a systemic change might not have gone the way he wanted it to, but he used the experience to deepen his understanding of the complexities and interrelatedness accompanying largescale reform. Perhaps his testimony to the U.S. House of Representatives Committee on Science on October 29, 1997 reflected some of the lessons of Project City Science; It simply is not possible, with or without federal support, to quickly change a system as huge and decentralized as American K–12 education. In our system

36    C. LANGE no federal agency can or should be able to command reform; but working together, the federal agencies can be and ought to be important partners in the reform for as many decades as it takes to create high-quality science education nationwide. (n.p.)

In 1977, Rutherford received a presidential nomination to serve as the assistant director of the National Science Foundation (NSF) and was responsible for the K–12 through post-doctoral science, mathematics and engineering education division. Following this appointment, the Department of Education cleaved from the Department of Health, Education and Welfare. It was during this time that Rutherford was given a different job where he was charged with heading a new department to try to establish interrelated and mutually supporting roles between the NSF and the Department of Education. The official title of this position was the Assistant Secretary for Research and Improvement in the United States Department of Education. This was a daunting position that included overseeing the National Institute of Education, National Center for Educational Statistics, Fund for the Improvement of Post-Secondary Education, and all the federal programs that support libraries and the development of educational technologies. This new role as a government administrator deepened Rutherford’s personal understanding of the nature of the science education reform process from a pragmatic standpoint. Not only did it provide him with a view of the agencies that surrounded the Washington policy-making process, but it further established Rutherford’s presence nationally as a member of the coalition of national reform leaders of all disciplines, but most importantly, those that were currently working on science, technology, and mathematics issues. During this period of his career, Rutherford’s involvement with international science education (specifically China, Japan, Russia and East and West Germany) gave him valuable knowledge that would influence his ideas about science education that he would eventually apply in the United States. Part III The Culmination of his Career American Association for the Advancement of Science In 1979, Rutherford was appointed as the deputy leader of the Pre-college Science Education Delegate to the Peoples’ Republic of China. In 1981 Rutherford was invited to join the American Association for the Advancement of Science (AAAS) to develop “plans and actions to give life to the AAAS’s desire to engage the scientific community energetically and knowledgeably in a sustained K–12 science education reform effort” (Rutherford, 2007, n.p.). Rutherford was qualified for such an appointment with his diverse background as

F. James Rutherford    37

a science teacher, curriculum developer, professional development facilitator, university professor of science methods, grant writer, researcher, author, government education officer, and reform leader. He fully appreciated the value and meaning of this new position and stated: It is from those experiences that I acquired a set of beliefs about science education reform—some fortifying my existing convictions, others changing them—that eventually led me to my role in the creation of Project 2061 and its first product, Science for All Americans. (Rutherford, 2007, n.p.)

He believed that the AAAS was the right organization for this work as well. Its large membership cuts across all scientific areas, and it was held in high regard globally, and by science policy makers worldwide. Because Rutherford had experience working in many different settings, he carefully weighted the association of this project to the AAAS. Carefully strategizing the team of professionals to assist him with the next project, one of his first decisions was to bring his friend Andrew Ahlgen on board. Ahlgren at that time was a tenured full professor at the University of Minnesota. After Rutherford assumed the role of education director of the AAAS, he “persuaded him [Ahlgren] to give up his tenured position to work at a place offering no such thing, to take a cut in salary and become dependent on our ability to secure grants, to leave his beautiful home to buy or rent in expensive Washington, DC, and to risk his professional future on the gamble that this new project was” (J. Rutherford, personal communication, 6/2/2012). Ahlgren’s role as the associate project director was a critical part of the production of Science for All Americans, Benchmarks for Science Literacy and the Atlas of Science Literacy (AAAS, 1989, 1993, 2006). The Atlas was the work of Ahlgren, who designed the benchmark strand maps that “stand as a monument to the importance of Chick Ahlgren in science education” (J. Rutherford, personal communication, 6/2/2012). Rutherford and Ahlgren complemented each other with their ability to throw ideas back and forth to one another. Rutherford explains how they volleyed ideas off one another: For nearly three years, Chick would meet with me in my office after working hours to wrestle with the question, how might engineering thinking, coupled with modern computers, contribute to the modernization of curriculum design? We read and argued about books on engineering design, marked up my white board with ideas and diagrams, and eventually refined the idea sufficiently to secure financial support. (J. Rutherford, personal communication, 6/2/2012)

The publication, Designs for Science Literacy (AAAS, 2001b), was an innovative project that included the goal to create a model of how computer assisted design could be applied to science curriculum design, which was never fulfilled.

38    C. LANGE

To achieve a scientifically literate body of K–12 students and future American citizens, Project 2061 created explicit learning goals that were compelling and challenging but attainable. The explicit learning goals were published in Science for All Americans and were “final rather than accumulating grade-band learning goals and without reference to teaching methods or materials” (Rutherford, 2007, n.p.). Most strongly stated was the determination that this accomplishment was to “be generated without involvement of or financial support from the federal government, in order that their authority will derive from the scientific community rather than from any agency of government” (Rutherford, 2007, n.p.). In order to create a systemic reform effort, Rutherford relied on three valuable lessons from his other endeavors and career experiences and so was seeking to address the following problems that had stopped reform in the past: 1. Sustainable science education reform needs to be steady rather than fragmented; 2. There needs to be a national consensus on the direction that science education should take; and 3. There are no short term or impatient solutions—effective reform requires long-term commitment from all involved. With a consensus of what science literacy means, Rutherford began to pull educators, scientists, administrators, mathematicians, engineers, historians, and learning specialists to produce the book that would allow districts and teachers the tools they needed for “fashioning their own curricula” (AAAS, 1989). This idea of setting a baseline, from which teachers could then individualize their programs to suit their needs and styles, is similar to the premise that Rutherford had in HPP. Several years after the release of Science for All Americans, and greatly influenced by the publication, the National Science Teachers Association and the National Academy of Sciences asked the National Research Council to coordinate the development of the national science education standards (NRC, 1996, p. 14). The National Science Education Standards was a specific directive for what should be taught, and its sister companion, Inquiry and Learning: A Guide for Teaching and Learning Science (NRC, 2000) focused on how science should be taught with Project 2061, thus providing the tools to guide all states to uniformity in K–12 science. Rutherford, as director and leader of Project 2061, had fulfilled the planned objectives upon joining the AAAS in 1985 in the “sustained K–12 science education reform effort” that would engage a scientific community (Rutherford, 2007). At the outset, Project 2061 was an American undertaking to foster science literacy in its own people. With time, however, it became widely known outside of the United States. The Europe-based Organization for Economic

F. James Rutherford    39

Cooperation and Development studied Project 2061 in 1989 and characterized it as the “single most visible attempt at science education reform in American history” (Allman, 1993). Translations of Project 2061 publications have been made in Asia, Europe, and Latin America. New Zealand researcher John Clark used the theoretical premises of Project 2061 to support a call for reform in teacher preparation programs in New Zealand that he believes are currently: overstuffed and undernourished with too much emphasis on a detailed understanding of the nuts and bolts of individual components of the mandated curriculum and far too little emphasis on the more fundamental contextual aspects of the curriculum such as historical determinants, sociological influences, philosophical justifications and political motivations. (Clark, 2005, p. 520)

Rutherford stepped down as education director of the AAAS and Project 2061 in 1998 and retired from the AAAS in 2001. Conclusion Of those in academia who rise to star status, most reach that level through repeated and consistent research publications. Their work, evidenced through their writing, becomes well read and well known, and, eventually, they are known throughout research circles. This research on F. James Rutherford highlights a very different pathway of that a high school teacher who was motivated by a fierce determination to create ambitious, creative, and immense projects with national scope and influence. During his lifetime, he has been honored with many prestigious awards including the highest award from the NSTA, the Robert H. Carleton Award; the Oersted Medal of the American Association of Physics Teachers; the University of California Lawrence Hall of Science Award for Lifelong Commitment to Science Education; the Paul F-Brandwein Lecture of the National Science Teachers Association (for “Is Our Past Our Future? Thoughts on the Next 50 Year of Science Education Reform in the Light of Judgments on the Past 50 Years”); the Prakken Professional Cooperation Award of the International Technology and Engineering Educators Association; the Distinguished Service Medal of the National Science Foundation; the Joseph H. Hazen Education Prize of the History of Science Society, shared with Gerald Holton, Harvard University; and the Ciné Award for distinguished educational filmmaking. He did not always succeed and openly admits, in a reflective manner, what he learned from his mistakes. Change is particularly tenuous during current times when educators of every kind are fearful of new trends emerging that are taking high stakes testing and job performance to new levels

40    C. LANGE

that Rutherford refers to as “intrusive,” such as No Child Left Behind. Time will tell if reform directives such as Project 2061 live up to their intended goals. Rutherford, in his unpublished manuscript, admits his doubts about the future: “I am well aware, indeed painfully aware, that much of the ‘fruit of my experience’ did not, alas, turn out to be as ‘durable as stone’” (2011, n.p.). Nonetheless, this prolific science educator has left an indelible mark on many areas of science education in America through his strong visionary leadership and keen political skills. References Allman, W. F. (1993, November 8). Back to first principles. U.S. News and World Report. Retrieved from http://www.usnews.com/usnews/culture/articles/931108/ archive_016072.htm American Association for the Advancement of Science (AAAS). (1989). Science for all Americans. New York, NY: Oxford University Press. American Association for the Advancement of Science (AAAS). (1993). Benchmarks for science literacy. New York, NY: Oxford University Press. American Association for the Advancement of Science (AAAS). (2001a). Atlas of scientific literacy. New York, NY: Oxford University Press. American Association for the Advancement of Science (AAAS). (2001b). Designs for science literacy. New York, NY: Oxford University Press. Clark, J. (2005). Curriculum studies in initial teacher education: The importance of holism and Project 2061. Curriculum Journal, 16(4), 509–521. Cuban, L. (1993). How teachers taught: Constancy and change in American classrooms 1890–1990. New York, NY: Teachers College Press. Dow, P. (1991). Schoolhouse politics: Lessons from the Sputnik era. Cambridge, MA: Harvard University Press. Edel, E. (1984). Writing lives: Principia biographica. New York, NY: W.W. Norton & Company. Glanz, E. (1979). What are you doing here? Washington, DC: Council for Basic Education. Holton, G. (2001, November). The project physics course, then and now. Keynote speaker address, Annual History of Science Meeting, Denver, CO. Retrieved January 6, 2012 from http://www.scienceeducationencore.org/library/origins/ the-origin-of-project-physics/ Kyle, W. C. (1985). What became of the curriculum development projects of the 1960s?. In D. Holdzkum & P. B. Lutz (Eds.), Research within reach: Science education (pp. 3–24). Washington, DC: National Institute of Education. Matthews, M. (1994). Science teaching: The role of history and philosophy of science. New York, NY: Routledge. McCormick, A. (1992). Trends and issues in science curriculum. In D. Cheek (Ed.). Science curriculum resource handbook: A practical guide for K–12 science curriculum (p. 18). Millwood, NY: Kraus International Publication. National Research Council (NRC). (1996). National science education standards. Washington, DC: National Academy Press.

F. James Rutherford    41 National Research Council (NRC). (2000). Inquiry and the national science education standards: A guide for teaching and learning. Washington, DC: National Academy Press. Rutherford, F. J. (2005). The 2005 Paul F. Brandwein lecture: Is our past our future? Thoughts on the next 50 years of science education reform in the light of judgments on the past 50 years. Journal of Science Education and Technology, 14(4), 367–386. Rutherford, F. J. (2007). F. J. Rutherford. Unpublished manuscript. Stake, R. & Easley, J. (1978). Case studies in science education (Vol. 1d). Urbana, IL: Center for Instructional Research and Curriculum Evaluation. Welch, W. (1979). Twenty years of science curriculum development: A look back. In D. C. Berliner (Ed.), Review of research in education 7. Washington, DC: American Educational Research Association.

42    C. LANGE

Chapter 3

Paul DeHart Hurd (1905–2001) Science Educator with a Social Vision Barbara S. Spector University of South Florida

He will be remembered as someone who succeeded in articulating a vision that shaped much of science education during the middle of the 20th century. —Elliot Eisner, May 27, 2002

Introduction Upon his death at age 95, Paul DeHart Hurd had been science education’s eminent elder statesman and eloquent spokesperson for decades. In his generation, reaching retirement at 65 was an accomplishment. One would have expected him to retire, not just from his university position (which he did), but also from work in science education (which he did not). Our enterprise had the gift of his contributions for seventy-two years, thirty years past retirement age. His final manuscript was submitted to BSCS: Innova-

Going Back for Our Future, pages 43–55 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

43

44    B. S. SPECTOR

tive Science Education (initially Biological Science Curriculum Study) organization one week before he died from pneumonia (Spector, 2007). Leading The Way I think of this intellectually gifted indomitable man who influenced science education world-wide as the “grandfather of science education” in the United States. “What he wanted, perhaps above all, was the creation of new ways of thinking for a new age” (Eisner, 2004). Inventing Science Education for the New Millennium (Hurd, 1997) synthesized what Dr. Hurd believed about science education in ninety pages filled with crisp quotable quotes (Roeder, 1998). Where there are postmodern approaches to pre-college education in the sciences that have had meaningful dramatic impacts, one finds Paul’s academic children and grandchildren on the forefront, for example, respectively, F. James Rutherford with Project 2061, and Roger Bybee with the National Science Education Standards (National Research Council, 1996). Rutherford completed his master’s degree with Paul while teaching in California and went on to Harvard and a distinguished career. Bybee was Rutherford’s student. Robert B. Sund earned his doctorate with Paul before going to the University of Northern Colorado and making several book contributions—a methods book, a book on Jean Piaget, one on questioning in science, and several activities books. When one looks at other colleagues who have led the charge for reform in the enterprise, one again finds Paul’s influence. For example, Robert Yager (and subsequently his academic lineage), the driving force of the science/technology/society (STS) movement, was influenced by his work with Paul on many initiatives, including setting the foundation for our current enterprise with Project Synthesis (Harms & Kahle, 1978). BSCS, the organization at the forefront of curriculum development for science education, was influenced by Paul’s thinking, direct involvement, and guidance for 35 years beginning in 1959. He encouraged them to develop a “lived curriculum,” meaning curricula in which students perceived they were involved with their own development and recognized they could use what they learned. BSCS used Hurd’s mantra, “science for life and living,” as the title of their 1990 elementary program in his honor. Paul used his elegant persuasiveness as the primary consultant to the National Academy of Sciences and consultant to the National Science Foundation (NSF), guiding the development of Requests for Proposals (RFPs) for many years forwarding his agenda. Starting in 1997, awards for science research proposals to NSF had to include projected ways the research findings were expected to be used to benefit social progress, the economy and science education (Mervis, 1997). Wherever there was

Paul DeHart Hurd (1905–2001)     45

change making learning science meaningful to individuals and society, it connects back to Dr. Hurd. Paul’s legacy is visible at every turn when one looks around science education currently. Many of the things we take for granted began with this visionary and his concern for the human dimension and impact of the natural sciences. He introduced “scientific literacy” in 1958; not just handson, but minds-on and a variety of other aphorisms. Scientific literacy was discussed by Hurd in 1998 p. 410: A concept of scientific literacy must recognize the range of changing forces in our society. Examples include such changes as the emergence of an information age, the birth of a global economy, and new ways of communication (the cyberworld). These changes provide a basic frame-work for constructing a meaning for scientific literacy that includes the ability of optimal use of science/technology knowledge (Boulding & Senish, 1983). Efforts to effect a supporting curriculum have been in progress for over 350 years (Longino, 1990; Hurd, 1970, pp. 13–20; Husén and Keeves, 1991). The idea that the emphasis on knowing should be on the utilization of knowledge was recognized by Confucius some 2500 years ago when he wrote that “the essence of knowledge is, once obtained, is to use it.” (RANN 2, 1976, p. 23)

Hands-on, minds-on was introduced to counter the assumption of many teachers that just experiencing a science activity enabled a person to learn. He emphasized the need to mentally interact with the materials and intentionally reflect on the events observed in order to learn. Middle school science and the way characteristics of early adolescence required teachers to provide learning opportunities in sync with these characteristics was unheard of before Paul turned a spotlight on it and explained how it could be done. Transforming Middle School Science Education, published in 2000, synthesized all his thoughts on education in these vital developmental years (Hurd, 2000). Paul’s uppermost interest was eliminating the gap between science and technology in society (Eisner, 2004). The science enterprise was separated from technology and the rest of society when Paul began his work. As a result, biology teachers were explicitly told to attend to basic science concepts to the exclusion of applied science or technology, which frustrated many teachers, me included. Paul embarked on this challenge of closing the gap between academic science and science for citizenry, which today is expressed as the science/technology/society reform movement, in spite of that challenge having three hundred and fifty years of history. He summarized that history in Scientific Literacy: New Minds for a Changing World (Hurd, 1998). When Paul encouraged inquiry as a way for students to learn science and for teachers to teach, he described “social inquiry,” meaning systematic

46    B. S. SPECTOR

investigation beyond the traditional discipline-bound notions of scientific inquiry (Hurd, 1998). His point was that learning science needed to incorporate logical thinking, problem solving, and decision-making skills to meet students’ social, economic, biological, and intellectual needs, while leading to responsible citizenship in this science and technology driven knowledge-based society. In doing so, science education would correlate with the shift in science itself from basic, theory driven research to socially driven, strategic research. Establishing new theories and laws is a procedure formerly recognized as basic research. Investigation into the functional aspects of science/technology as it relates to human welfare, economic development, social progress, and the quality of life is strategic research (Hurd, 1998). He sometimes used the expression, science enlightenment, when talking about this phenomenon. Amnesia He read voraciously with an analytical eye, staying current in science, in technology, and in the condition of society. He often pointed out the paradox between the fast paced changes in the natural sciences and the stagnation in science teaching and the science education enterprise. Many years ago Paul wrote about what the current generation of senior science educators in the last few years discussed at the Association for Science Teacher Education (ASTE) Seniors Forum that was in fact the stimulus for this book: “Our enterprise suffers from amnesia. In science education we do not learn from either our past successes or failures. We just constantly reinvent the wheel and give it a new name” (Personal conversation, January 20, 2000). Historian He was the unofficial historian and archivist for the science education enterprise. He talked knowledgeably about science education from the 1600s, the time of Francis Bacon, to 2000. Regardless of what aspect of the enterprise one wanted to discuss, Hurd would do so supporting his comments with hard data from his studies and his life experience. He had lived so much of science education’s history and interacted with the great thinkers of the past, such as John Dewey. Paul’s continuous analysis of the science education enterprise began in 1947 with his dissertation at Stanford University titled, A Critical Analysis of the Trends in Secondary School Science Teaching from 1895–1948. The landmark report he later wrote titled, Biological Education in American Secondary Schools 1892–1960 (Hurd, 1961/1984) updated his analysis. He continued

Paul DeHart Hurd (1905–2001)     47

updating his accounting in the 1970s when he chaired the Project Synthesis biology committee made up of Jane Butler Kahle, Robert Yager, and Roger Bybee. The committee’s mandate was to ascertain the status and future of biology education. At the National Science Teachers Association conference in 1992, his address summarized the past century of science education in America. Paul’s collection and all his works are now housed in Cubberley Education Library at Stanford University in California (https://www.stanford. edu/group/cubberley/collections/hurd). It is described by the library’s website this way: The Paul DeHart Hurd Collection on Science Education includes materials on the teaching of science, science education policy, curriculum, and related topics. Professor Hurd collected these materials throughout his career as a science education reformer, and they represent milestones in science curriculum design, government policy toward the teaching of science, and innovations in science pedagogy.

The Person Dr. Paul DeHart Hurd, had no need for self-aggrandizement even though it was common for him to be in the limelight, since public media often sought him out for his clear and direct communications about science education policy. This wise, kind and gentle man with a subtle sense of humor and a twinkle in his eye never spoke a negative word about another person (Bybee, 2002). Whether you were a middle school dropout sitting on the steps of a school with Paul, a classroom teacher at a conference, someone else’s graduate student telephoning him from across the country, or a professional colleague, you could be sure he would be respectful, exhibit interest, express empathy, and be generous with his time and extensive knowledge. Early Years Hurd was the grandson of a western “gold country” boomtown newspaper publisher. As a young boy in Colorado, he wandered around the prairies searching for Native American artifacts and loved to lie in the prairie grasses and watch the insects. He learned the native lore of the plains, the teachings of The Great Circle, the Medicine Wheel, and the Vision Quest as a means of developing one’s identity in a culture. This set the stage for his professional posture emphasizing the need for all youngsters to learn science as a part of their culture, not something isolated from it.

48    B. S. SPECTOR

An accident damaging his ankle in first grade left Paul with a significant life-long limp exacerbated by age. When the injury got infected, swelled, and did not heal properly after considerable time, his mother had the wisdom to tell him he would have to “use your head instead of your feet to make a living.” And such a creative, dynamic, brilliant brain was in that head! Fortunately for us, he ultimately chose to use it to the betterment of society through science education. In 1989, after extensive interviews with Paul DeHart Hurd, science educators Emmett Wright and Seliesa Pembleton wrote a biographical sketch describing Paul’s life experiences and included extensive quotations from their interviews with Paul. My understanding of Paul’s early years emerged from that sketch. The quotations from Paul that follow are excerpted from that article. Paul had his first taste of teaching science when he was a student in a Denver junior high school. The teacher of a newly instituted general biology course in which he was enrolled had to leave the classroom to attend many curriculum-planning meetings. Paul had previously completed a botany course. He offered to share what he had learned with his classmates by filling in for his teacher when the botany section of the course was to be studied. The principal agreed, as long as a substitute teacher sat in the back of the room. It was not a big leap from this experience to an AB degree in botany in 1929 from State teachers College at Greeley, Colorado (now, University of Northern Colorado) where he had been supported by a four-year scholarship. Paul reflected: I ended up there because I didn’t have much money and needed a scholarship. Greeley, at that time was about 80% women and 20% men and had a new president who wanted to increase the number of men who were attending the college and going into teaching. At Greeley, I decided to major in zoology. In high school botany I had my best teacher. In college zoology, I had my worst teacher. In fact, I went to the president of the college and asked why the teacher was hired. The president counseled, “You plan to be a teacher, and this course is required, so why don’t you just study his methods as an example of what you will never do as a teacher?” I must say that was the best methods course I ever had—even though it was self-taught. (P. D. Hurd, personal communication, July 13, 1987)

Teaching was clearly something he liked. In fact, he voluntarily did extra practice teaching while earning his degree, even though it did not count in his program. His exquisite powers of persuasion were in evidence early when he graduated from college and applied for a biology teaching position in Greeley High School: He convinced them they should waive their required two years of teaching experience, because he was qualified for the position by virtue of his four years of laboratory teaching assistant experience in bacteriology, zoology and biology and his extensive practice teaching experience.

Paul DeHart Hurd (1905–2001)     49

Paul’s experience earning a living as a teacher in Greeley High School during the Depression is echoed by financial uncertainties many teachers face in today’s economic downturn that began in 2007. He reminisced: I started out in 1929 within the minimum salary range in Colorado, which was $125 per month. The next year, I got a 15% cut in pay. The next year salaries were cut again. After 3 years teaching experience, I was earning $68 per month. (P. D. Hurd, personal communication, July 13, 1987)

This was particularly noteworthy when one considers that Hurd’s undergraduate research on intercellular leaf space volume in sugar beets gained him an offer of a research position at Great Western Sugar Company upon graduation. His mentor, Frank C. Jean, convinced him not to accept it, because he believed scientific research should be done purely for the sake of science, and should not be done for a commercial company. Paul’s ten-year high school teaching career began during the Depression. Basic survival was a concern for many of his students. He believed they needed to know about proper nutrition and the wellbeing of their own bodies. They lived in rural areas and knew about animals and plants being raised on the land, but not about themselves. He created his own biology courses to help them understand the workings of their own bodies and their interaction with the environment—human biology and human ecology. He included human genetics in his high school biology course and was thought to be the first to do so. Connecting human science with investigations into human culture led him to even touch on evolution, a topic forbidden by the superintendent. Fortunately, no one complained. His classroom was filled with treasures, such as mastodon teeth, bones of prehistoric camels and horses, and arrowheads collected from local land that had been scoured by the dustbowl thus revealing artifacts, and artifacts from Alaskan Eskimos. (He had traded artifacts with a teacher in Alaska.) All of this was a far cry from the strictly basic science concepts in a typical biology course at that time (and in some places even now). His students did learn the basic science concepts, but always in the context of themselves as humans and the ecology of the world in which they and others lived. He was obviously an intellectual risk taker willing to say and do what he believed was important. Paul’s career-long quest to integrate science, technology and society in K–12 schools was a manifestation of his concern for cultural validity that was first expressed in the biology courses he developed at Greeley High School. The culture of modern and postmodern society in our democratic country is determined primarily by innovations in technology and findings from science research that drive our lives. Thus the STS movement was a promulgation of Paul’s concern for science teaching in a cultural context.

50    B. S. SPECTOR

Paul’s interest expanded from his own classroom to encompass the lives of younger children in elementary grades. Nature study was declining. Paul convinced the school board to allow him to develop an elementary science curriculum for the entire school system. His contemporary, Gerald Craig, had done a dissertation in 1927 that provided a teaching model for elementary science of interest to Hurd, that lasted until the reform movement in the 1960s. Paul’s career as a pioneer in science education was on its way. Creating A Vision Hurd developed his first vision of science education for people living in the 20th century while he was attending a series of graduate seminars organized by the president of the college at Greeley. The topic of study in one seminar was the manuscript, A Program for Teaching Science, published by the National Society for the Study of Education (Tildsley, 1932) written by a committee chaired by S. Ralph Powers of Columbia University. Wilbur L. Beauchamp, University of Chicago; Francis D. Curtis, University of Michigan; and V. T. Thayer, Chairman, Commission on Secondary School Curriculum of the Progressive Education Association (PEA) were other notables conducting seminars Paul attended while earning his master’s degree, which he completed in 1932. The University of Iowa offered a $300 per year scholarship for Paul to work toward his doctorate continuing his investigation of sugar beets in 1939. About halfway through what could be his dissertation, he was notified the scholarship was being reduced for all graduate students to $150. In response to the cut, Hurd moved to Los Angeles in 1940. But, this Great Plains explorer quickly concluded it “was no place for a human being to live.” He moved on and found a job in Menlo School and Junior College near Stanford University. He completed his doctorate in 1947 at Stanford and joined the faculty at Stanford. He described the evolution of his perspective on his science education research this way: In general, my research interests were shaped back in the progressive education days of the 1930s. My association with members of the P.E.A. writing team for Science In General Education, the writing of a former high school science teacher, John Dewey, and the work of J. D. Bernal (University of London) on The Social Function of Science, strongly influenced the course of my future views and study. Though all my training for educational research had been in statistics and methodology, reflecting the “scientific movement” in education current at that time, I felt that something was missing. It seemed to me that empirical research would never mean all that it could unless housed in a theoretical framework, much in the way that theory gives meaning to scientific research. It was a deliberate decision on my part to take the lesser trav-

Paul DeHart Hurd (1905–2001)     51 eled and rougher road in science education and concentrate on the meaning of an education in the sciences for an age of science. I interpreted the social contract between schools and society to mean science for everyone, not just for the college-bound and the aspiring scientists. Over the years I have found that many of my peers do not understand or accept the rationality and objectivity of philosophical, historical, or policy research. I feel today, as I observed fifty years ago, that too much of the research in science education is to remain forever trivial for the lack of a conceptual framework to give it meaning. (P. D. Hurd, personal communication, July 13, 1987)

Policy Maker Many policy-making groups benefited from Paul’s vision, as governmental and non-governmental groups appointed him to membership. Paul was appointed to the first national committee to develop a public education program for National Aeronautics and Space Agency (NASA). His work was acknowledged with the Apollo award in 1970. Terrance Bell, Secretary of Education, asked Paul to write a position paper on science education for the National Commission on Excellence in Education in 1982 (Hurd, 1982a). The National Science Board Commission on Precollege Education in Mathematics, Science and Technology asked him for a similar document. He then prepared a position paper to present at the National Academy of Sciences Convocation and Forum in May, 1982 (Hurd, 1982b). The Forum, choreographed by Senta Raizen, was accompanied by a well-orchestrated media blitz declaring to the public there was a crisis in science and mathematics education (Spector &Yager, 2011). A summary of that statement follows: Among developed and developing countries throughout the world there is widespread recognition of the importance of science and technology in fostering human well-being, and in meeting the economic and political demands of living in the 20th century. For the United States to meet similar goals will require a knowledgeable and concerned public as well as specialists in the generation of scientific knowledge and technological innovations. The committee of pre-college education in the sciences and technology must be to the general public. There is a vast amount of scientific information that should be part of human experience, not only for the intellectual insights it provides to the individual, but for its usefulness in advancing human welfare and the nation. To achieve a science education with these ends will require the cooperative endeavors of parents, scientists, engineers, government officials, school administrators, media specialists, and experts in the social sciences, and all others with a positive vision of America’s future. These people need to lay the groundwork for school curriculum in the sciences and technology that has both scientific and cultural validity.

52    B. S. SPECTOR

This call for a systemic approach to change was later heeded by NSF with RFPs to fund science education systemic initiatives, and ultimately state, rural and urban systemic initiative projects. Paul DeHart Hurd has been honored for his contributions to science education by many organizations: elected to the American Academy of achievement in 1962; the National Science Teachers Association’s 1969 Distinguished Service to Science Education Citation; the National Aeronautics and Space Agency 1970 Apollo Award; the National Science Teachers Association’s 1979 Robert H. Carleton Award for National Leadership in Science Education; the National Association for Research in Science Teaching 1987 Distinguished Contribution to Science Education Research Award; the American Association for the Advancement of Science Fellow; awarded honorary life memberships and emeritus status in eight professional societies; honorary doctorate of science degrees were conferred by the University of Northern Colorado, Ball State University, and Drake University; and he was recognized in American Men and Women in Science, Who’s Who in the West, and Leaders in Education. Epilogue Historian Hurd was stimulated to summarize the 350-year history philosophically underpinning the post-modern STS movement, the longest lived reform initiative in science education in the U.S., by Aristotle’s (384– 322 bc) words written after surveying the schools of Rome: There is no general agreement about what a young person should learn, either in relation to more virtue or to success in life. The existing practice is perplexing; no one knows on what we should proceed—should the useful in life, or should virtue, or should the higher knowledge, be the aim of our training; all three opinions have been entertained. Again, about the means there is no agreement; for different persons, starting with different ideas about the nature of virtue, naturally disagree about the practice of it. There can be no doubt that children should be taught those useful things which are really necessary . . . (Jowett & Hutcher, 1964, p. 268) For modern times it is science/technology literacy, a lived curriculum, and an understanding of the current practice of research in science/technology that are needed to make science useful in our lives . . . [Learners must understand the] socialization of science and its relevance to how science impacts our culture, our lives, and the course of our democracy . . . The level of one’s ability to use advances in science/technology for improving various aspects of one’s life is now viewed as building “human capital.” In this context, scientific literacy represents cognitive capacities for utilizing science/technology information in human affairs and for social and economic progress. (Hurd, 1998, p. 411)

Paul DeHart Hurd (1905–2001)     53

Hurd had some doubt whether the NSES and Project 2061 goals that were expressed within the context of academic science rather than strategic science were leading our nation in this direction. We still do not know the answer. “Business as usual” is the dominant action in K–12 schools in spite of the past three decades of public proclamations of the urgent need for change. Most science in schools is descriptive, focused on laws, theories and concepts of discrete science disciplines. It seems the old adage “The more things change the more they stay the same” often describes K–12 science education. Politics at all levels, national, local, university and K–12 school, continuously raise barriers to progress in one form or another. As much as BSCS led the way for curriculum change in keeping with Paul’s vision, Roger Bybee, when he was executive director, acknowledged “. . . To be truthful, funding priorities publishing demands, and other realities left us with less than the ideals Paul espoused” (Bybee, 2002, p. 10). Science education has been presented with a well-articulated vision, which has great promise, and a model of a person with incredible persistence. Paul’s long term dedication to make science education consistent with current science and technology is an inspiration to me and others who feel discouraged at times, because the process of change is so slow in K–12 science education. The dramatic restructuring of K–12 schooling required to attain Paul’s vision of science education has not yet come to pass. In the United States to date, the entire scope of Hurd’s vision has not been implemented. We have yet to experience the full fruits of the genius of Paul DeHart Hurd. References Bybee, R. W. (2002, Spring). Remembering Paul DeHart Hurd: A clear vision and strong voice for science education. The Natural Selection, 9–10. Eisner, E. (2004, June 2). Memorial resolution: Paul DeHart Hurd. Stanford Report. Stanford University, Stanford, CA. Retrieved from http://www.stanford.edu/ group/cubberley/collections/hurd Harms, N. C. & Kahle, J. (1978). The status and needs of pre-college science education: Report of project synthesis. (Final report to NSF for Grant, SED 77-19001). Washington, DC: National Science Foundation. Hurd, P. D. (1961). Biological education in American secondary schools, 1890–1960. Reston, VA: American Institute of Biological Sciences. Hurd. P. D. (1982a). An overview of science education in the U.S. and selected foreign countries. Prepared for the National Commission on Excellence in Education. U. S. Department of Education. Washington, D.C. ERIC No. ED 227–076 Hurd, P. D. (1982b). “What is the problem?” in science and mathematics in the schools: report of a convocation. Washington, DC: National Academy of Sciences. Hurd, P. D. (1997). Inventing science education for the new millennium. New York, NY: Teachers College Columbia University.

54    B. S. SPECTOR Hurd, P. D. (1998). Scientific literacy: New minds for a changing world. Science Education, 82(3), 407–416. Hurd, P. D. (2000). Transforming middle school science education. New York, NY: Teachers College Columbia University. Jowett, B. & Hutcher, S. H. (Trans.). (1964). Aristotle: Politics and poetics (book eight). New York, NY: The Heritage Press. Mervis, J. (1997). NSF adopts new guidelines. Science, 276, 26. National Research Council. (1996). National Science Education Standards. Washington, DC: National Academy Press. RANN 2. (1976). Realizing knowledge as a resource (Vol. I). Washington, DC: National Science Foundation. Spector, B. S. (2007). Paul DeHart Hurd: Staying the course for 72 years. In S. Totten & J. Pedersen, (Eds.), Addressing social issues in the classroom and beyond: The pedagogical efforts of pioneers in the field (pp. 253–266). Charlotte, NC: Information Publishing. Spector, B. S. &Yager, R. E. (2011). The many faces of STS: Social issues in science education. In S. Totten & J. E. Pedersen (Eds.), Teaching and studying social issues: Major programs and approaches (pp. 277–312). Charlotte, NC: Information Publishing. Tildsley, J. L. (1932). A program for teaching science. The thirty-first yearbook of the National Society for the Study of Education. Journal of Chemical Education, 9(5), 962–966. Wright, E. L. & Pembleton, S. M. (1989). Science as a personal experience: Paul DeHart Hurd. Science Education, 73(2), 195–205.

Paul DeHart Hurd (1905–2001)     55

This page intentionally left blank.

Chapter 4

A Pioneer Senta Raizen Formerly at the National Science Foundation and at the National Academy of Sciences/National Research Council

Introduction A phone call from the National Science Foundation (NSF) in 1962 catapulted me into science education. At the time I was working at the National Academy of Sciences as a literature chemist developing an encyclopedia of chemical terms. My previous career included being a research chemist for the Sun Oil Company, developing new processes for separating hydrocarbon fractions. Throughout my college years, graduate school, and my work as a chemist I had had little interest in education other than my own. Working at the National Science Foundation The call from NSF came as a result of the country’s investment in science education, responding to the shock of Sputnik, sent aloft by the Soviet Union in October 1957. By 1962, the country was in the depths of the Cold War,

Going Back for Our Future, pages 57–78 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

57

58    S. RAIZEN

including the Cuban Missile Crisis in October. The most dramatic response to Sputnik had come in 1961, with President Kennedy’s announcement of the ambitious goal to send an American to the moon. The scientific community, with support from the National Science Foundation (NSF), had already taken initiatives to stir the education system into developing manpower able to respond to the Soviet challenge and the race to the moon. This included the expansion of students enrolled in technical fields as well as reform of high school science and mathematics curricula. The NSF caller asked if I would be interested in working on the curriculum development projects the agency was supporting to recruit high school students into science and mathematics and ensure an up-to-date curriculum. NSF recruited only people with a scientific rather than an education background. I lacked any formal training in education, but I had become somewhat interested in the field because I had school-age children.1 I accepted with alacrity the job ­­­­­­ offered to me and became the cognizant program officer for the physical science curriculum projects. This included independent handling of proposals from initiation to grant action and subsequent monitoring and evaluation of projects. These early years as NSF assistant/associate program officer were probably the most exhilarating years I spent in science education. I had wonderful mentoring from my two NSF supervisors—Charles Whitmer and John Mays—from whom I learned how to do my job. Moreover, I had the opportunity to interact with extraordinary scientists of the time devoting their efforts to improve the science (and later mathematics) that was being taught in schools. Also, as I learned later when I was myself soliciting grants from various government and private agencies, it was a lot more fun to have money to give away than to try to obtain money. There were three assumptions behind the initial NSF curriculum development program. First, that involvement of working scientists was critical. This was completely consonant with the culture of NSF, which had been created as a balance wheel for basic research to support scientists in innovative work not related to a particular mission agency, such as the National Institutes of Health or the U.S. Department of Energy.2 Second, the high school population, and specifically the science- and mathematics-able students, needed to be encouraged towards science-related and technical fields. Third, some high school science curricula were outdated and sometimes incorrect and needed to be revised if not totally re-fashioned. The last assumption came from the examination of the current curricula by scientists who found the texts and associated materials wanting with respect to focus, accuracy, and currency. In sum, the aim of the NSF program was to create better curricula for elite students, which would both increase the size of the potential technical workforce and the level of its knowledge before entering college.

A Pioneer    59

During this period, I profited greatly from the opportunity to observe the scientists’ summer curriculum-writing sessions, where they identified the key concepts to be taught within the relevant science. Attending numerous field trials made me aware of some of the implementation difficulties that lay ahead. These included the need for better teacher preparation in science, particularly at the elementary school level, and the need to provide instructional strategies to teachers. Confronting the simplistic thinking that just having good content rich curriculum materials was enough to change practice helped us develop the program for the National Center for Improving Science Education (NCISE), which we founded subsequently.3 Implementation difficulties also raised our awareness of the need to move beyond the “one teacher at a time” approach to implementation. NSF’s Curriculum Projects The basic model of developing rigorous curricula for science-talented students was exemplified by the Physical Science Study Committee (PSSC), headed by the eminent MIT physicist Jerrold Zacharias. The projects were very costly—we joked at the time of the unit of one Zach (one million dollars, but more like ten million today) being the minimum investment for PSSC.4 The high cost included not only the time of the scientists writing curricula but also careful piloting of the draft curricula with high school students around the country and subsequent revisions. PSSC also set up a network of PSSC teachers, which proved a powerful component in implementation of the curriculum as envisaged by the scientists who created it. This provided one example for reaching groups of teachers with similar concerns rather than working with one teacher at a time. This approach was feasible in physics because of the relatively small number of physics teachers, as few students enrolled in physics compared to the first level of high school science, which usually was biology. As the field trials and implementation of PSSC proceeded, the teachers themselves contributed their experiences and developed effective teaching strategies, which were shared among the teachers and scientists. The original conception of PSSC, now a straightforward high school physics course, was to do an integrated physics/chemistry course, but this idea foundered on the disagreement between physicists and chemists on the description of the hydrogen atom to be included in the text. The physicists wanted to focus on the hydrogen atom as the simplest atom to explain the nature of atomic properties, including weak and strong forces. The chemists, on the other hand, wanted to focus on the strong bonding property of the hydrogen atom due to the single electron in its electron shell. Thus the chemists developed their own curriculum proj-

60    S. RAIZEN

ects, including the most influential, Chem Study, still influencing high school chemistry texts today. The biologists developed Biological Sciences Curriculum Study (BSCS).5 BSCS developed three alternative high school biology curricula based on Molecular Biology, Ecology, and Cell Biology. The organization still exists under its original name and has created a number of additional curricula as well as a newsletter focused on BSCS activities and other science education news. BSCS was led for a long time by Rodger Bybee, who became highly influential in science education policy throughout the decades that followed the curriculum development effort of the sixties. Two flaws of the original approach soon emerged. First, the curricula were being implemented by untrained and sometimes ill-prepared teachers. Second, the high school students lacked a foundation in science required to deal with the rigorous high school curricula developed by the study groups. To address the first flaw, NSF developed a program of summer institutes for teachers that offered both courses in the science required to teach rigorous high school curricula and also instruction in implementing the specific curricula such as PSSC Chem Study and BSCS. Eventually NSF also realized the importance of involving science educators as well as working scientists in the development of the curricula. To address the second flaw, during the 1970s and 1980s NSF extended the curriculum development program to junior high school and eventually elementary school. A consequence of moving down into the lower grades was the shift from addressing only the most capable students (the science elite) to broadening the goal to science understanding for all. It was deemed important for everyone to understand the natural world that surrounds us. Approaches to Science Literacy Several approaches developed as a result of NSF’s program goal evolving to “science for all.” One of the projects for elementary school was the Astronomy Project, developed by Myron Atkin. Another influential project was the Science Curriculum Improvement Study (SCIS), led by Robert Karplus, which introduced Mr. O (an outside observer of events) to children. Karplus was a highly creative and deep thinker about the fundamental concepts he wanted children to understand and how to introduce these concepts to them. He once remarked wistfully to me that this work was as difficult as doing physics but not recognized as such. His brother, also an eminent physicist, had been elected to the National Academy of Sciences for achievements in physics, whereas Bob Karplus received no such recognition from any scientific body for his work in education.

A Pioneer    61

High school science course development was the focus of additional projects such as Harvard Project Physics, and later Chemistry in the Community (Chem Com) and Project 2061. Harvard Project Physics was led by Gerald Holton, James Rutherford and Fletcher Watson from 1962 to 1972. Harvard Project Physics took a different approach from PSSC in the hope of attracting more girls to the study of physics. The supposition was that a humanistic historical treatment of the development of key concepts in physics would be more appealing to girls than PSSC’s mathematical approach. Unfortunately this strategy did not bear fruit, as physics enrollment of females did not increase substantially and is still low today. On the other hand, female enrollment in chemistry grew steadily to nearly two thirds of all females enrolled in high school and today surpasses that of males. Chem Com was launched in the 1980s by the American Chemical Society (ACS). Chem Com tried to interest students in chemistry through an issuesfocused curriculum that would introduce chemical concepts as needed. The course was designed to take advantage of interest in environmental concerns as they developed in the 1980s (e.g., pollution of air and water). Chem Com as a high school course had an interesting history. Although intended to be as rigorous as Chem Study, it was considered by many school administrators as a weaker alternative for less able students. The influence of college entrance requirements became evident since many tertiary institutions did not accept Chem Com as a full chemistry credit. Through ACS’s lobbying of the influential California higher education system, the California Regents decided to grant Chem Com equal status to the more traditional chemistry courses. This influenced many colleges and universities in the rest of the country to also accept Chem Com as a legitimate chemistry course with respect to entrance requirements. The project with the widest time horizon was Project 2061, also started in the 1980s under the leadership of the visionary Rutherford.6 He wrote the influential report, Science for All Americans (originally published in 1989), which formed the basis for the succeeding work of the project, including developing Benchmarks for Science Literacy and an Atlas of Science Literacy.7 The Atlas shows how the individual benchmarks for science education connect to each other and appear across the K–12th grade to develop understanding of important ideas in science. The Role of Physicists in Science Education Reform The Conceptual Physics movement represented a more radical approach to curriculum reform. Rather than just revising individual courses, it sought to make more logical the entire structure of the high school science curriculum by presenting physics first and using it as a foundation for chem-

62    S. RAIZEN

istry and subsequently biology. Physics in its traditional placement within the curriculum followed biology and chemistry and relied on advanced mathematics, making it virtually inaccessible to most students. Conceptual Physics, on the other hand, “de-mathematized” physics while still retaining the study of fundamental physics concepts, and placed physics first as fundamental to the understanding of chemistry, and chemistry as fundamental to the understanding of biology. In the late 1980s, Paul Hewitt developed a textbook for the proposed high school course based on his popular college text for freshman, Conceptual Physics. Another notable advocate for reversing the usual high school science sequence was Nobel Laureate Leon Lederman, representing the continued strong involvement of physicists in science education reform. Lederman retired from Fermi Lab in 1989 to join the faculty of the University of Chicago where, among other things, he became Science Adviser to the governor of Illinois. In response to a request from the Department of Energy, which was becoming involved in science education to help carry out its mission, Lederman undertook to reform science and math education in the Chicago Public Schools (CPS). He was instrumental in founding two institutions within CPS. One was the Teacher’s Academy for Math and Science (TAMS), partially funded by the state and partially by CPS, and the other was an advanced science and mathematics high school for very able students. While TAMS flourished for some years, it came to a close after a decade or so for lack of sufficient state funding. The difficulty that researchers coming from the natural sciences have had in dealing with education is transposing their worldviews from a hard science, such as physics, to education—for example Leon Lederman stated that he would transform Chicago Public Schools in five years. This goal, of course, was not achieved. Nevertheless, through such ambitious projects, the field learnt a great deal, myself included. Since the 1960s, all of us have learned how difficult it is to make changes in public schools, transforming teachers’ ideas about who should study science and how students learn. Indeed, students have selected themselves out on the basis of expecting science and mathematics to be too difficult for them, particularly females and students belonging to some minority groups. Systems and policies surrounding science education often present particular obstacles, not only in K–12 but also in higher education. For example, the earlier struggles faced by Chem Com were replicated by the controversies over Conceptual Physics. The college admission requirements did not allow this less-mathematically rigorous course to be counted as a science credit and therefore made it less popular among high school administrators than Hewitt, Lederman, and other physicists had anticipated.

A Pioneer    63

The “New Math” As NSF’s curriculum improvement program evolved, I was also given responsibility for K–12 mathematics, including the School Mathematics Study Group (SMSG) led by Ed Begle of Yale and Stanford Universities and the University of Illinois Committee on School Mathematics (UICSM) led by Max Beberman. Both intended to have children understand the number system instead of just regurgitating arithmetic facts they had learned without understanding. It is worth noting here that the “new math” was deemed a failure. This, however, is a complete misunderstanding of the initial mathematics curriculum programs. For example, SMSG introduced bases other than base ten (e.g., base 2, critical for understanding computer programming) and set theory and set notation to high school students. Because SMSG never turned over their materials for commercial publication, commercial publishers picked up the externalities of the SMSG program and actually pushed some of these notions down to middle and elementary school—an obviously mistaken idea. Failure of SMSG’s original approach for high schools at lower levels could have been anticipated and gave the “new math” a bad name.8 Project Adoption or Adaptation The various NSF supported curriculum projects took different approaches to getting their materials to students. Some, for example Chem Study, thought it best to have a formal publisher already active in the textbook field, partly in the hope of influencing the publisher’s related science textbooks. This proved a successful strategy, as today’s chemistry textbooks still show the influence of Chem Study. Others, such as BSCS, established their own centers for publication, which allowed them total control and continuity. This method also proved quite successful, as BSCS continues to this day to develop and publish biology and health related curricula under their own auspices as well as through one of the major commercial textbook publishers. Still others, notably SMSG, published only in paperback experimental editions, in this way hoping to influence commercial publishers into picking up and adapting their ideas. Part of the difference in approaches was a philosophical one. Namely, whether curricula should be “adopted” exactly as written and envisaged by the developers, or would work more effectively if “adapted” by educators to fit the local context, or publishers to gain market share. Adaptation, however, carried the risk of changing the curriculum so that the fundamental concepts might not be treated adequately. As noted, this risk induced NSF to support summer workshops for

64    S. RAIZEN

teachers that focused on the implementation of the curriculum projects the agency had funded.9 Evolution of NSF’s Curriculum Improvement Program Toward the end of my tenure at NSF, I dealt with myriad programs that NSF had created to support greater involvement and improved performance of students in science and mathematics, including fellowships and traineeships in science, institutional development to strengthen science and mathematics education in schools and universities, development of curricula and educational technology, upgrading of physical facilities, research programs for high school and college students, and training for teachers. Unfortunately, after I left NSF in 1972, the agency’s support of curriculum development projects became much more limited. This decrease was triggered by controversy surrounding the implementation in schools of Jerome Bruner’s social science curriculum, “Man: a Course of Study” (MACOS) in 1970. Because this curriculum dealt with the adaptation of humans to various conditions, similar to animal adaptation, it clashed with some religious values and created a backlash not only against MACOS but against federal support of curriculum development in general. Sexism in Science and Science Education From 1969 to 1971, during my last couple of years at NSF, I was asked, together with a female colleague, to investigate the status of women employees at NSF. This study was very revealing to me regarding the institutional barriers faced by women in science and science education. I knew discrimination against women occurred in general, but I had not worried overly much about these barriers in my own career. My first brush with possible sexism was in college when, at the end of my freshman year, I asked to be a lab assistant for freshman chemistry in the following year. The chair of the department looked at me and said, “I have never had girl lab assistants.” I asked how the lab assistants were selected, and he said it was on the basis of their grades in freshman chemistry. As I got the highest grade that year, the chair was true to his word and appointed me in my sophomore year as a lab assistant for freshman chemistry. I went to graduate school at a women’s college (Bryn Mawr) and had no difficulty in obtaining employment as an industrial research chemist for Sun Oil. I was probably fortunate in my timing because in 1945, when I obtained this job, men had not as yet returned from World War II, and positions were open. It may be significant that in my job interview I responded negatively to the

A Pioneer    65

question of whether I could type and take short-hand—a question that never would have been asked of a man of equal qualifications. Although I could type and take shorthand, I had the feeling that if I revealed this, I might end up as a girl who did quasi-secretarial work. I wanted to do the research, not be the person who simply wrote up the notes and typed professional papers resulting from the research. As it turned out, the research was both interesting and significant and eventually led to an important industrial process for separating hydro-carbons through adsorption. After getting married and during the raising of three children through infancy, I worked at home as a literature chemist, translating German articles and abstracting both German and English scientific articles for Chem Abstracts. (German was my mother tongue.) During this time, I took some education courses required by the state of Virginia for teacher certification. My plan was to teach chemistry, a job with a schedule coinciding with my children’s schedules—a typical concern for a mother at the time. I acquired provisional certification as a chemistry teacher, to be fully licensed after a year of practice teaching. Perhaps fortunately for my future high school students, when I was ready to go back to work, I had no difficulty obtaining a position at the National Academy of Sciences on my own terms, limiting my hours so I would be home when the children got home from school. I was able to negotiate these same arrangements—unprecedented, I believe, at the time—when I joined NSF in 1962. My experiences led to my attitude that hard work of high quality would always be acknowledged regardless of gender. In other words; “If I could make it, anyone of similar ability and drive could also.” I had to change this attitude after our study of NSF’s women employees. It turned out that they remained at lower levels, never achieving the rank of program or division director. This was especially striking because our report found that they were the ones who did the work of the agency and made the grant programs effective, but they never got promoted after they reached a certain level. Again, I was somewhat of an exception, as I had moved out of the grant-awarding track to a more general policy planning and oversight position as the assistant to the assistant director for science education. Nevertheless, because of my work on the report, I recognized the institutional barriers to women’s advancement at NSF.10 (This has changed over time, with women now serving not only as division directors but also assistant directors of various branches of NSF.) Expanding Beyond Science Education After leaving NSF in 1971, I got involved with planning for the National Institute of Education (NIE), a rather quixotic attempt to create a coun-

66    S. RAIZEN

terpart in education to the highly successful National Institutes of Health (NIH), which supported basic research to improve medical practices. This activity led me to accept a position with the Rand Corporation the following year. At Rand, the inner sanctum of systems analysis, one of my assignments was to design an evaluation for the Head Start program. I remonstrated, saying I knew nothing about early childhood education, but was told that as a systems analyst, I had the tools to address any program regardless of familiarity with the field. My work at Rand indeed gave me the confidence to think about indicators of the quality and effectiveness of science education programs in general. Partly because I had been involved in its planning, in 1974 I accepted an associate director position at NIE where I was in charge of building the infrastructure for research and development in education, including original research, development, and dissemination. One of my group’s tasks was to administer the Education Resources Information Center (ERIC), which pre-dated the foundation of NIE and became our responsibility. At the time ERIC had several centers specializing in disseminating education research in different areas, including a very active Science and Mathematics Education Assistance Center (SMEAC). A continuing challenge for all the ERIC centers was quality control, that is, the decision about what research was valid and deserved inclusion in the database.11 The model of NIH, however, proved not a good one for education. Despite detailed plans, NIE was soon overwhelmed by the politics of education and, rather than being an independent agency, became part of the newly created Department of Education. The department inherited a number of ongoing programs such as ERIC, as well as a considerable staff invested in these programs. NIE’s survival as an independent agency was further impeded by the fact that its first director was a brilliant researcher who had, however, little experience with communicating either with Congress or the educational establishment. The research functions originally envisaged for NIE are now the responsibility of the Institute of Education Sciences (IES), in the Department of Education.12 The main responsibility for promoting educational programs that contribute to producing a capable and scientifically literate citizenry, however, continues to reside with NSF. My disenchantment with NIE led me back to the National Academy of Sciences in 1980. The Academy had just created the Commission on Behavioral and Social Sciences and Education (CBASSE), in recognition of the importance of the social sciences with education as somewhat of an afterthought. At CBASSE, I helped to develop a research and development agenda and indicators of the condition of mathematics and science education (National Research Council, Committee on Indicators of Precollege Science and Mathematics Education, 1985). I was also responsible for organizing a national convocation on mathematics and science education and

A Pioneer    67

writing the resulting report distributed to 30,000 individuals in education and the sciences (AAAS, 1987; Levinsky, 1982; National Research Council, Committee on Research in Mathematics, Science, and Technology Education, 1985). My work at the NAS/NRC taught me a great deal about program evaluation, including programs for teachers and students and assessment of what students know and can do in science. This learning served me well in our later NCISE evaluation work. NCISE My work at the academy, and particularly the convocation, persuaded me that there needed to be a center for science education that would regularly report on the health of science education in the U.S. and stimulate improvements. Together with Susan Loucks-Horsley, I founded the National Center for Improving Science Education (NCISE). The center was initially funded by the Department of Education through a competitively awarded grant and later received support from several foundations. The center currently is based within the Science-Technology-Engineering-Mathematics (STEM) Program at WestEd.13 This affiliation has proved a happy and productive one, allowing the center and me to undertake work of both importance and interest with collaboration from universities as appropriate. NCISE was created to synthesize research and practice knowledge and develop analytical reports and recommendations. NCISE was intended to bridge the gaps among research, practice, and policy, and promote cooperation and collaboration among organizations, institutions, and individuals committed to improving science education. The center developed reports at the elementary, middle, and high school levels, including “The High Stakes of High School Science.” The reports dealt with the science curriculum at the given level, program assessment, and teacher development. NCISE also returned to the issue of adequate science education for prospective elementary school teachers (Raizen & Michelsohn, 1994). A continuing focus for NCISE has been program evaluation, with particular attention to teacher induction, the need for students to understand the “designed world,” and the importance of developing good measures or “indicators” of the state of science education, specifically assessments of what students know and are able to do in science. Program Evaluation One focus of my work was the overall evaluation of programs designed to improve the quality of science and mathematics education. Observing a

68    S. RAIZEN

number of programs made it possible for us to develop templates spelling out characteristics of successful elements leading to effective implementation of innovative curricula. Program evaluation generally included observation of the particular program in action, as well as analyses of written documents pertaining to the program, such as results of teacher development components and student achievement data and then comparing our findings to the elements in the template. Thus, for instance, for several years we evaluated the effectiveness of TAMS in providing teachers with the science and mathematics knowledge they needed as well as pedagogic strategies for helping their students become proficient in these subjects. Another example was the ten year evaluation of the Mid-Atlantic Center in Mathematics Teaching and Learning (MAC-MTL), one of the first centers established by NSF in the 1990s to improve mathematics and science education, and the one of only two that received a second five-year award after the initial five-year funding. Strangely, in education, NSF did not follow accepted good practice in allowing at least a ten-year period for development and effective impact of these centers, as was the case with centers NSF supported in other areas of its responsibility. Instead, possibly under pressures from the education community, the funds available to NSF for the purpose of the education centers were spent on creating new centers instead of allowing full development and functioning of just a few selected ones. The history of MAC-MTL that we documented clearly established that five-year funding was inadequate to create fully functioning centers; nevertheless, a new administration took over the education component of NSF, the center program was abandoned. This stands in stark contrast to NSF’s practice in its support of centers in basic and applied research, for example, NSF’s support for over twenty years of centers to develop materials science. Teacher Induction As we became familiar with educational practices in many countries through our work in TIMSS (Trends in International Mathematics and Science Study) and PISA (Program for International Student Assessment), it became evident that a number of countries where students scored high had very explicit programs for introducing new teachers to the classroom and to their teaching responsibilities, unlike the common U.S. practice of “sink or swim.” This led us to study a series of induction practices in five countries, selected on the basis of their students’ high scores in PISA and the comprehensiveness of their induction programs. These countries were France, New Zealand, the German-speaking cantons of Switzerland, China (Shanghai), and Japan. It was striking to find that teaching in all these venues was considered a public and collaborative act, contrasted to the U.S.,

A Pioneer    69

where teaching commonly was and still is considered a private act in control of the individual classroom teacher.14 Our research resulted in several publications, including More Swimming, Less Sinking (Britton, Raizen, Paine, & Huntley, 2000), written for the National Commission on Mathematics and Science Teaching in the 21st Century, and Comprehensive Teacher Induction (Britton, Paine, Pimm, & Raizen, 2003). These publications added to the movement within the U.S. of providing more mentors to beginning teachers but generally did not introduce more comprehensive induction programs such as characterized in the five countries we had studied. In order to provide good examples of U.S. practice, we researched exemplars in this country, resulting in the publication of Bold Ventures (Britton, Raizen, Atkin, Huberman, & Rowe, 1996). Large Scale Assessments Three assessments, two international and one national, occupied much of my intellectual work in the last few years. My role was to help develop the framework and test specifications that would guide the development of questions and tasks for each assessment, and to review the resulting assessment before it was administered for fidelity to our specifications. In each case the concern has been to provide maximum information in a very short testing time as school systems have become more and more reluctant to take time away from instruction for testing, especially in the U.S., where the “No Child Left Behind” requirements already levy a heavy testing burden on students, teachers, and instructional time.15 For all three of these large-scale assessments, the challenge was similar: how to develop assessment tasks that would mirror the goals of education in the given domain (e.g., science) without distorting them too much. For science, there clearly are limits for this type of on-demand, limited time assessment. For example, one cannot assess the crucial characteristic of perseverance applied to the pursuit of a research question when the original approach proves unsuccessful. Nor is it possible to track students’ ideas about how to change the research approach appropriately, based on results obtained initially. Throughout the development of these assessments, Audrey B. Champagne played a critical role in the shaping of the framework and specifications of the intended tests.16 She became a trusted colleague on whose judgment I learned to rely. TIMSS In the early 1990s, I was asked to work on the preparation for the Third International Mathematics and Science Study (TIMSS), later renamed

70    S. RAIZEN

Trends in International Mathematics and Science Study. This was first conceived as a repeat of an earlier mathematics assessment, but fortunately, science was added as an additional focus. Eventually, the study involved over 40 countries and prepared tests at the fourth grade, eighth grade, and high school levels. To ensure that the assessment would be fair to all the participating countries, it was necessary to analyze their standards, frameworks, and textbooks to establish what was included in each country’s curriculum. In order to accomplish this analysis we needed to identify specific pieces of each framework and textbook and code them for the mathematics and science content. This laborious endeavor established that both in mathematics and science the U.S. was more inclusive than most other countries—“the mile wide and inch deep curriculum.” It also demonstrated the repetitiveness of the U.S. curricula as students advanced from grade to grade—particularly in mathematics. Our examination of curricula over the participating countries resulted in several influential publications (Schmidt, Raizen, Britton, Bianchi, & Wolfe, 1996; Schmidt, McKnight, & Raizen, 1996). The assessment administered in 1995 showed that while our elementary students ranked among the highest achieving countries, the ranking decreased as U.S. students progressed to eighth grade and high school. By twelfth grade U.S. students placed no better than average among the participating industrialized countries. These results have been quite consistent over several administrations of TIMSS. My work with developing the science part of TIMSS assessments continued throughout the following decade, including framework development and reviewing reports on U.S. and international results. NAEP At the same time, starting in the 1970s and continuing to the present, I co-chaired several of the committees charged with developing the framework and assessment specifications for the science National Assessment of Educational Progress (NAEP) assessments and served on the overall guiding committee for test development. Over the several science assessments (1972, 1976, 1986, 1990, 1996, 2009), the nature of NAEP changed considerably. Ralph Tyler, instrumental in the original conceptualization of NAEP, envisaged it as a true indicator of the state of education in the various fields tested (including not only science, but mathematics, reading, social studies, etc.) with absolute interdictions to report anything except at the national, and later regional, level. The assessment was matrixed, assessing nationally representative samples of students. Although the testing period was only 50 minutes, including a short background questionnaire, the content was spiraled so that more of the domain could be assessed. Results were only to be reported at the national level for the three grades tested (4, 8, and 12). As interest in NAEP results increased, some states sought to have results re-

A Pioneer    71

ported at the state level. This required state-level representative samples of students to be tested, as well as nationally representative samples. Eventually, almost all states participated in NAEP assessments, and state NAEP rankings were reported in NAEP publications and the media. This transformed NAEP from a low-level indicator to a high-stakes assessment, as individual states appeared to be ranked against each other in NAEP reports. Another factor that called attention to the NAEP assessment results for individual states was that, in some cases, there existed a disparity in results reported on substantially easier state tests versus those reported in more difficult NAEP tests within the same domain. Thus, the results of some state exams showed much higher proficiency in the tested domains at given grade levels than appeared in NAEP results for the same grades. More recently, NAEP has also reported results for large urban systems, but generally NAEP never reports on individual districts, schools, or students. Simultaneously with this transformation was a change in the guiding frameworks for each assessment—from a small booklet for the assessments in the 1970s which merely listed the topics to be included in the assessment to an 8½" × 11" detailed book for the 2009 and 2011 assessments.17 The current frameworks and test specifications include examples of test questions that illustrate the meaning of specific components of the framework. For example, in the 2009 science framework one component discussed under Science Practices is “Using Science Principles.” One of the examples in this section provides a food web starting with sunlight to corn and oak trees and progressing to mice, robins, snakes, and hawks. The question asked of students is “If the corn crop failed one year, what would most likely happen to the robin population? Explain your answer.” The scoring guide allows for several correct responses, provided a reasonable explanation is given for the response. What this example illustrates is the potential of probing more deeply students’ thinking than is possible with a multiple choice question. Also, greater attention was paid to having the test specifications match the frameworks, both commissioned by the national assessment governing board and developed by specially appointed committees. One means of ensuring consistency between framework test specifications and the actual test being developed by an outside contractor was to have oversight of the whole process by an advisory group that included several members of the original framework committees, including myself. PISA Another international assessment that I also worked on from its initiation was the Programme for International Student Assessment (PISA). Conceived by the Organization for Economic Cooperation and Development (OECD), it was intended to assess reading, mathematics and science every three years, focusing on students’ ability to apply their knowledge and com-

72    S. RAIZEN

petence in these three fields to novel problems. While all three domains were to be assessed in each of the three-year cycles, one was given prominence in each cycle, thus focusing on reading, mathematics, or science, and heavily weighted toward that specific domain. The original notion was to prepare an assessment only for OECD member countries (i.e., the industrially developed world); however, PISA became so popular that developing countries asked to be able to participate as well. PISA assesses student knowledge and competence at age 15, the common ending of compulsory education and preparation for majority citizenship in many countries. Like TIMSS and NAEP, PISA too was based on nationally representative samples rather than a full census assessment testing all students. It therefore only reports at the country level. Again, U.S. students did not do well in this international comparison. PISA in particular has been criticized for not probing the science domain adequately since it focuses on problem solving and application of knowledge rather than the amount of knowledge retained by students.18 Comparing Assessments TIMSS, NAEP and PISA have been analyzed to compare their content. It turns out that NAEP and TIMSS are similar, but NAEP is a somewhat more rigorous assessment. PISA does not fit into the same paradigm, as its focus is on problem solving and application of knowledge rather than the amount of knowledge retained by students. Because of the careful development of open-ended questions, including clear scoring rubrics (see the food web example above), NAEP goes beyond testing declarative knowledge to probe procedural and even strategic knowledge. For the U.S., NAEP has become the “Gold Standard” of h­­­­­ow assessments should be developed and what they should assess, against which state assessments can be judged. Engineering Concepts The need to have students understand the designed world was already recognized during the period of NSF’s original curriculum support, resulting in the Engineering Concepts Curriculum Project. They produced The ManMade World (ECCP, 1971), which dealt with engineering concepts and design principles. As a result of this interest, NCISE produced the book Technology Education in the Classroom: Understanding the Designed World (Raizen, Sellwood, Todd, & Vickers, 1995). It is interesting to note that in the mid1990s, when the book was published, few educators were even considering teaching engineering concepts in K–12 education. Yet by 2011, several states

A Pioneer    73

have made this content a part of their standards for student learning and the National Assessment of Educational Progress (NAEP) decided to administer an assessment of these concepts in 2014. To this day, I have pursued my interest in having students understand the designed world and engineering concepts by co-chairing the committee that developed the framework for the 2014 assessment in technology and engineering literacy.19 Assessing 21st Century Skills In the last few years there has been a movement to specifying more generalizable skills across domains such as creativity, problem solving, use of information, and so on. This is quite controversial with some science educators, who doubt that skills such as problem solving are generalizable across fields, but believe that such a skill has to be lodged in context and content. I was involved in leading the writing of “Chapter 2: Defining 21st Century Skills” (Binkley et al., 2012) for the book Assessment and Teaching of 21st Century Skills, developed under the auspices of the University of Melbourne, AU. The chapter synthesizes the work of a number of countries in defining these generalizable skills and gives some examples of how they can be assessed. Ten skills were identified, organized into four categories: Ways of Thinking (e.g., problem solving), Ways of Working (e.g., collaboration), Tools for Working (e.g., information literacy), and Living in the World (e.g., personal and social responsibility) (Binkley et al., 2012). The work continues with the design of actual assessments in the area of problem solving in a collaborative context, which are being field tested in several countries at this time. WestEd continues to be involved in this work, although I retired in 2010 at age 85. Lessons Learned It may be useful to summarize the various lessons about improving science education learned over the course of my career. Lesson 1: Education is an applied field and requires the contribution of researchers in relevant areas as well as practitioners (teachers, principals and administrators) and the interested public. In other words, as important as their contributions are, it takes more than the talent of highly qualified scientists to make lasting improvements in K–12 science education. Lesson 2: “Education is everybody’s business.” Because everyone has been to school, everyone has an opinion about education. Moreover, because it is deemed so significant for children’s development and the nation’s future, education is extremely politicized. Much more so than with politicized sci-

74    S. RAIZEN

entific topics (e.g., women’s fertility, climate change), this makes a logical process of inquiry and improvement following the scientific model highly questionable. Majority opinion may not allow one to ask the pertinent question, formulate appropriate hypotheses, nor publicize polarizing results. Lesson 3: Another way in which the scientific model is a poor fit for educational improvement is the “scalability” problem. Replication holding independent variables constant is intrinsic to the scientific process, but this is impossible in a field where people are the variables. Statistics is a useful tool in social science research where large sample sizes are available, but this does not help in the wider application of reform initiatives developed in small settings. Lesson 4: The U.S. assumptions of what students can master at a given age or level have been undermined by PISA and TIMMS, which demonstrated more advanced learning in some other countries than expected of U.S. students. To date, U.S. expectations have been based largely on long ingrained traditional curricula and instructional strategies. The current advances of cognitive science should provide the education community with better guideposts for what children at a given age and level are able to learn, and the proper sequencing of science concepts. In individual classrooms, the “wisdom of practice” (the educator’s knowledge and understanding of students), should continue to inform the education process in concert with the application of cognitive science findings. Lesson 5: Society keeps raising its expectations, the knowledge base keeps expanding, and the technological advancements keep multiplying. For example, the current emphasis on 21st century skills, including effective use of information technology, as contrasted to narrower competencies, reflects the changing views of what education needs to provide to students. More broadly, education must prepare students for the fast-changing technological and economic environments and the globalization of work. There has ­­­ never been any shortage of challenges in trying to improve science education in the schools. Shortly before I retired a friend and colleague said half-jokingly to me, “You have been at this for 50 years now, and you haven’t fixed science education yet?” The answer is that we keep enlarging our goals, for example from improving science education for the elite to achieving science literacy for all. And we continue to meet new challenges. One of these challenges lies immediately ahead of us, namely the resource constraints—including the loss of teachers and the lack of appropriate facilities—induced by current economic conditions. And yet, to overcome these unfavorable economic conditions surely will require the support of a science-literate public and the effective use of information technology in science education. In order to maintain the U.S.’s position as a global leader in science and technology in the face of ever-stiffer international competition, we must ensure that a core of highly qualified high

A Pioneer    75

school graduates will be prepared to become our scientists and engineers of the future. These challenges guarantee that science education reform will continue to occupy the attention of people like me. ACKNOWLEDGEMENT The author wishes to thank her granddaughters, Feonix Fawkes, Trude Raizen and Claire Raizen, for their invaluable help in producing this chapter. I am legally blind and they served as my eyes—typing the manuscript and doing some editing and limited research. Notes 1. As my children grew older, I prepared to be a high school chemistry teacher so that my working schedule would fit theirs. See section “Sexism in Science and Science Education.” 2. See Vannevar Bush, “Science, the Endless Frontier” (1945). The creation of NSF grew out of the important contribution that basic research had made to winning World War II. It took five years before NSF actually came into being in 1950. 3. See section titled “NCISE.” 4. One million dollars was the approximate cost of PSSC. 5. http://www.bscs.org/ 6. The AAAS website explains the name of the project as follows: “Project 2061 began its work in 1985-the year Halley’s Comet was last visible from earth. Children starting school now will see the return of the Comet in 2061—a reminder that today’s education will shape the quality of their lives as they come of age in the 21st century amid profound scientific and technological change.” http://www.project2061.org/about/default.htm 7. See http://www.project2061.org/ 8. Anecdotally, on a personal level, when my daughter switched from chemistry to computer science for her graduate work, she fell back on the SMSG mathematics she had studied in high school. Interestingly, the pendulum has swung back to having even elementary school children understand the number system rather than rote memorization of arithmetic facts, e.g., the new Montgomery County mathematics curriculum about to be published by Pearson. 9. After several years, the Office of Management and Budget ordered the agency not to provide curriculum specific summer institutes anymore, but to concentrate instead on more general teacher training in the relevant domains. This decision was taken because of concerns of commercial publishers that this government-funded training might give the NSF curricula an unfair advantage in the marketplace.

76    S. RAIZEN 10. Even the general language conveys sexism, as the English language does not have a single word meaning “both male and female,” such as the German “Mensch.” At the time, I, like everybody else, was not aware of the implicit sexism in the NSF curriculum project titles such as “The Man-Made World” and “Man: A Course of Study.” 11. Currently ERIC has been transmuted into an online digital library. According to the website “ERIC provides ready access to education literature to support the use of educational research and information to improve practice in learning, teaching, educational decision-making, and research.” http://www.eric. ed.gov/ERICWebPortal/resources/html/about/about_eric.html 12. IES supports research and development across all subject areas focused on measuring what works and what doesn’t to inform and improve educational outcomes. This includes research centers for science and mathematics and many research studies of discrete science and mathematics curricula and interventions. See http://ies.ed.gov/ 13. WestEd’s mission statement reads, “WestEd, a research, development, and service agency, works with education and other communities to promote excellence, achieve equity, and improve learning for children, youth, and adults.” From http://www.wested.org/cs/we/print/docs/we/agency.htm. 14. Consider, for example, Japan’s practice of “lesson study,” which has intrigued U.S. researchers, who have studied this practice in depth and are now introducing it to several venues in the U.S. 15. The baleful effects of narrowly conceived testing have been demonstrated by some of the No Child Left Behind testing, instituted by a number of states which have resulted in narrowing the curriculum and emphasizing memorization rather than teaching for understanding, particularly in science. Further, since NCLB focused on mathematics and reading, science dropped out of the curriculum altogether in many elementary schools, or received just a few minutes per week—totally inadequate to the notion of involving children in hands-on experimentation. NCLB has been reshaped by the Obama administration by granting states more flexibility, and also making funds for improvement available through the new program, “Race to the Top.” These reforms, however, do not deal with the problems created by current US testing practices. 16. Audrey B. Champagne, now retired, was a Professor, Department of Educational Theory and Practice and Department of Chemistry, University at Albany, State University of New York (SUNY). 17. See the NAEP frameworks available at the NAEP website, http://nces.ed.gov/ nationsreportcard/ 18. Development of the PISA assessment originated with identifying problems that would be important in the next twenty years, such as fossil fuels and alternative fuels, resource consumption, availability of water, etc., rather than the topics traditionally included in specific science courses, such as Physics, Chemistry, and Biology. Hence the criticism stemmed from a traditional view of science assessments and failed to consider the innovative PISA approach. 19. See http://nces.ed.gov/nationsreportcard/

A Pioneer    77

References AAAS (American Association for the Advancement of Science). (1987). Assessing the quality of the science curriculum. Presented at the AAAS Forum on Science Education and published in This Year in School Science, 1986. Washington, DC: Author. Binkley, M., Erstad, O., Herman, J., Raizen, S., Ripley, M., Miller-Ricci, M., & Rumble, M. (2012). Defining twenty-first century skills. In P. Griffin, B. McGaw, & E. Care, Assessment and teaching of 21st century skills (pp. 17–66). New York, NY: Springer. Britton, E., Paine, L., Pimm, D., & Raizen, S. (2003). Comprehensive teacher induction: Systems for early career learning. Boston, MA: Kluwer Academic Publishers. Britton, E., Raizen, S., Atkin, J. M., Huberman, M., & Rowe, M. B. (1996). Bold ventures: Case studies of U.S. innovations in science and mathematics education (3 Vols.). Dordrecht, the Netherlands: Kluwer Academic Publishers. Britton, E., Raizen, S., Paine, L., & Huntley, M. A. (2000). More swimming, less sinking: Perspectives on teacher induction from the U.S. and abroad. Commissioned paper for the National Commission on Mathematics and Science Teaching in the 21st Century, 2000. Bush, V. (1945). Science, the endless frontier. A Report to the President by Vannevar Bush, Director of the Office of Scientific Research and Development, July 1945. Washington, DC: United States Government Printing Office. ECCP (Engineering Concepts Curriculum Project). (1971). The man-made world. New York, NY: McGraw Hill. Levinsky, N. G. (1982). Science and mathematics in the schools: Report of a convocation. Washington, DC: The National Academies Press. National Research Council, Committee on Research in Mathematics, Science, and Technology Education. (1985). Mathematics, science, and technology education: A research agenda. Washington, DC: The National Academies Press. National Research Council, Committee on Indicators of Precollege Science and Mathematics Education. (1985). Indicators of precollege education in science and mathematics. Washington, DC: The National Academies Press. Raizen, S., & Michelsohn, A. M. (1994). The future of science in elementary schools: Educating prospective teachers. San Francisco, CA: Jossey-Bass. Raizen, S., Sellwood, P., Todd, R., & Vickers, M. (1995). Technology education in the classroom: Understanding the designed world. San Francisco, CA: Jossey-Bass. Rutherford, F. J. (1991). Science for all Americans. New York, NY: Oxford University Press. Schmidt, W. H., McKnight, C. C., & Raizen, S. (1996). A splintered vision: An investigation of U.S. science and mathematics education. Boston, MA: Kluwer Academic Publishers. Schmidt, W. H., Raizen, S., Britton, E. D., Bianchi, L. J., & Wolfe, R. G. (1996). Many visions, many aims: A cross-national investigation of curricular intentions in school science. Boston, MA: Kluwer Academic Publishers.

78    S. RAIZEN

Chapter 5

Willard J. Jacobson 20th Century Visionary in Science Education Rodney L. Doran University at Buffalo, Emeritus Abby B. Bergman Independent Consultant

Few individuals have had such a profound influence on the advancement of science education in the United States and in many other regions of the world as Willard Jacobson. His impact upon individuals, his mentorship, and his international leadership all serve to leave a stamp of influence, humanity, and colleagueship. His stature, both literally and figuratively, was profound and will most assuredly affect the outcome of science education in the most positive ways for countless years to come. Not having the chance to interview Willard about this chapter, the content will be the authors’ best interpretations from his writings and our best recollections from his actions and discussions. We invited many of his graduates and colleagues to share with us recollections and stories of him and

Going Back for Our Future, pages 79–98 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

79

80   R. L. DORAN and A. B. BERGMAN

his impact on their professional and personal lives. Also, we have shared our interpretations with Willard’s wife, Carol, and daughter Ellen. Willard—A Background with Rural Roots Willard J. Jacobson was born in 1923 and attended elementary and high school in Hixton, a farm community in west central Wisconsin. Growing up on a farm provided direct experience with nature. In Willard’s era, agriculture also had a significant place in the rural school curriculum. Attending college was not a given for young men from these small towns, so Willard must have had family and/or teachers who witnessed his intellectual capabilities and encouraged him to pursue a college education. He enrolled in the nearest state teachers college, the University of Wisconsin at River Falls. As this was in the middle of World War II, Willard chose to join the U.S. Army Air Corps (later called the Air Force) and completed flight training, received his wings and second lieutenant bars. He served in the European theater with a unit that won recognition for its work in Belgium. Having served as a co-pilot of a B-26 bomber, he was later decorated with an Air Medal with seven oak leaf clusters. After completing his military service, Willard returned to college and completed his undergraduate degree program, earning his BS degree in chemistry in 1946. In addition to working on his chemistry program, Willard was active in the Honor Society, the Science Club, the Mathematics Club, and he also played on the varsity baseball team. He was recognized by the University of Wisconsin at River Falls as one of its outstanding alumni in 1976. His first teaching assignment, lasting just one year, was as a science teacher at Palmyra, Wisconsin. Willard then moved to New York City and taught at the Horace Mann-Lincoln School in Manhattan, near the Columbia University campus. With benefits from the GI Bill, Willard was able to quickly earn his MS degree in science education in 1948 from Teachers College, Columbia University and his EdD degree in science education in 1951, also from Teachers College, Columbia. The title of his dissertation was Science Education and the Development of Abilities to Cope with Problematic Life Situations. Willard’s family recalled his great appreciation for the G.I. Bill, which facilitated the quick, efficient completion of his master’s and doctorate at Teachers College, Columbia University. Willard was always pro-active in helping others to complete their programs of study and life tasks. Some have described Willard as a “relentless optimist.” This optimism helped him to view any obstacles that he may have encountered, such as his agricultural roots, a lack of tradition for higher learning in his family, or his participation in the war, as minor “speed bumps” on the path to eventual success.

Willard J. Jacobson    81

Establishing a Distinguished Career at Teachers College, Columbia University This bright, young, (28 year-old) doctoral graduate was immediately hired by Teachers College, Columbia University and began his long, illustrious career there. At that time (the 1950s), graduate programs in science education were largely housed in the large private institutions (e.g., Columbia, Harvard, and Stanford). Willard began his career at one of the handful of elite institutions with doctoral programs in science education, but he maintained his humble, helpful approach to his career of research, teaching, and service. Willard’s doctoral committee at Teachers College was chaired by S. Ralph Powers (1897–1970), a giant in science education during the early 1900s. In most cases, doctoral advisors have a significant influence not only on the dissertation, but often in shaping the outlook and perspective of the student. Clearly, Dr. Powers became a prime influence in his development as a scholar and he also left a mark on Willard’s continued work and research. In addition to Dr. Powers, another science education faculty member at Teachers College was Gerald S Craig. Like Dr. Powers, Professor Craig was a prolific scholar in elementary science education and a charismatic leader in the field. If nothing else, the traditions and expectations for Teachers College science education faculty set a high bar. It is also quite likely that there was considerable mentoring available for young Willard from these giants in the field. During Willard’s tenure at Teachers College, there were a number of other world-class scholars with whom Willard worked. He collaborated with Richard Wolf, a renowned scholar in the area of tests and measurements, on the Second International Association for the Evaluation and Educational Achievement (IEA) Science Study and other projects, such as his studies of population growth. Other notable Teachers College faculty that worked on IEA studies included Robert Thorndike, Harold Noah, and Harry Passow. In the 1960s, O. Roger Anderson joined the science education group at Teachers College. For many years, Willard and Roger shared leadership of the Department of Science Education at Teachers College. Being located in New York City, one of the largest and most internationally-oriented cities of the world, facilitated interaction with visitors from many countries and universities. The administrators at Teachers College, Columbia consistently recruited and appointed world class scholars. Indeed, a premier institution with world-class scholars would also attract the finest graduate students. So, Willard benefited from a college of esteemed colleagues and a continuous stream of graduate students and assistants to help in his work and writing.

82   R. L. DORAN and A. B. BERGMAN

The 1950s—The Afghanistan Project Willard was a consultant or visiting scholar to many groups, colleges, and nations, including a Fulbright scholar at London University, a visiting professor at the University of Hawaii, and a consultant to UNESCO. To be sure, these involvements were all very productive and impressive. One very special activity—a two-year position in Kabul as a science consultant to the Royal Afghan Ministry of Education (1954 to 1956)—clearly catapulted Willard to international stature. This position/adventure came after Willard had been at Teachers College as a faculty member for only three years. He was accompanied on this adventure by his wife, Carol, and family (at that time) of two young daughters, Ellen and Susan. In an April 1955 letter, Willard reported that Suzy is in first grade in a school that has just been started. Since the kindergarten is being held in our home, Ellie takes part in that. Carol is teaching the kindergarten for the first two hours of the morning, and during the last two hours of the morning, she works in the lower grades of the elementary school. (Jacobson, personal communication, April 7, 1955)

As if not already busy enough, Willard commented further: We have just finished planting our garden. Willard and Carol were both taking Persian lessons, and Carol was doing better. Too many of our friends had infectious hepatitis or yellow jaundice and there has been considerable polio among the people here. Both of our girls have had measles along with a large fraction of the other children in Kabul. Health, of course, is one of the great concerns of people on missions such as this. If a safe source of drinking water can be developed (we, of course, boil our water) Kabul could be one of the healthiest places in the world.

Willard was working with one of the laboratory schools just outside of the city. He continued: The boys in the fifth grade want to learn something about electricity. There is no electricity in the village, but they have hopes. Late last summer, we began thinking about and planning for a series of workshops during the three months of the winter when the schools of Afghanistan are closed.

As one of the workshops was to be in the Helmand Valley, they traveled there by Jeep. They encountered bands of nomads along the way that tried to disrupt their progress. Willard asked the guide/interpreter what would happen if they stopped. In his cautious way, he said, “It would mean our lives.” Willard reported that they did not stop, adding that the nomads were

Willard J. Jacobson    83

an “independent lot” and that they must lead an unbelievably harsh, rugged life. Those who survive are indeed a hardy people. From another letter sent by one of Willard’s daughters, it was revealed that at the same time as he was working with Afghan officials and teachers to plan workshops and activities and helping his young family acclimate to this difficult country, Willard was drafting outlines for books. His plans for a three-year series of general science textbooks for the junior high school years were sent to the American Book Company (Jacobson, personal communication, April 10, 1955). It is clear how Willard managed to get so much accomplished—he was a master at multi-tasking, balancing the needs of several projects simultaneously—in addition to being bright and a fine writer! A Clear Focus Emerging for His Mission— Willard’s Commitment to Elementary Science Education Willard’s dedication to science was initially demonstrated by his decision to major in chemistry in college. Having enrolled in a state teachers college also immersed him in the world of education. When you combine these elements, the ingredients for why Willard might consider science and science education as a lifelong career were clearly set. While Willard was a thinker and leader in many facets of science education, much of his writing was focused on the teaching of science in the elementary schools. As recalled by one of Willard’s graduates, Ira Kanis: ”Willard Jacobson, Professor of Natural Sciences” was how he would introduce himself to the audience at professional conferences or organizations. Professor Jacobson would often say that through effective elementary school science programs, children would develop an interest and an appreciation of the world in which they lived. He was a champion of elementary science education often challenging his audience of educators, researchers and school administrators by asking, “Why do children come to us as question marks and leave as periods?” (Kanis, personal communication, Feb 5, 2012)

Willard believed that school age children (the adults of tomorrow) lived and played in the natural world, and they were constantly confronted by and questioned the phenomena of the natural/physical world. He believed that it was essential for elementary school age children to be taught science by integrating it with their experiences in the environment. He asserted that children are interested in science, and their natural curiosity should be developed and encouraged in the elementary school. He affirmed that we should not wait until the child was of high school age to develop these interests. Willard believed that experiences in science contributed to the

84   R. L. DORAN and A. B. BERGMAN

growth and development of children and could help children learn how to think. He helped establish the idea that through science education in the elementary schools, we could instill in children a method of working and solving problems. This would, he argued, contribute to the general education of children. From his early writings, it is clear that Willard believed science should be a central part of all levels of education, beginning in the elementary school. He also believed that engaging, student-initiated experiences should replace the book learning in science classes. Much of his writing was to produce material that teachers could easily use to facilitate high quality experiences in science. He believed that books should be viewed as just one resource. Willard affirmed that science should not be viewed as an island (or silo), but that student experiences in science should be connected to the skills and experiences in mathematics, reading, and other disciplines. Even from his early books, it became clear that Willard believed that studying science was a powerful resource of information and experimental procedures critical for all citizens to understand issues and problems that affect individuals and society. Making decisions should be based on comprehensive background knowledge and a thorough analytical lens to examine proposals, plans, and claims. Clearly, for Willard, science was a fundamental way of knowing and a tool to be used by citizens to make sense of the increasingly technological world in which live. Building upon the work of John Dewey, Willard valued education (and particularly science education) as necessary for participation in a democratic society. If our citizens are to participate in the decision making process that can propel society, then its citizens need to have developed a scientific literacy to understand some of the major issues confronting a modern, technological society. Such topics as nuclear waste disposal, population control, and proper nutrition should be addressed by an active, educated populace. Willard Begins a Career of Writing Willard’s first professional writing was his dissertation, entitled Science Education and the Development of Abilities to Cope with Problematic Life Situations (EdD, Teachers College, 1951). In the introduction of his dissertation, Willard stated that the development of abilities to cope with life situations had been “advocated by almost all policy-making groups who have made a study of what should be the nature of our educational system.” He argued that such abilities are necessary because: 1. The welfare of our society depends on such abilities, 2. Individuals need these skills to face obstacles, and

Willard J. Jacobson    85

3. Our ways of life are becoming more and more interdependent (Jacobson, 1951). Already, there were emerging themes and foci that drove so much of Willard’s future work—the need for education to help every student and the need to address societal issues and problems. The sheer number of his publications is simply startling. This career of writing began in 1951 and continued into the 1990s. Most of the obituaries of Willard stated that he had written over 50 books, while one of his graduates claims he wrote 73 books (June Miller, personal communication, Feb 3, 2012). The list of publications just from the Second IEA Science Study was considerable (10 monographs, 15 articles. 4 chapters, and one special issue of a journal). By any perspective or approach, his published works rate Willard as one of the most prolific science education scholars of all time. In reflecting upon Willard’s own development through his published work, Elizabeth Meng, his doctoral graduate and Teachers College adjunct professor offers the following: Willard’s curriculum materials reveal a gradual shift not so much in the content that was used, but more in how children were involved with topics. The early textbooks were just that, textbooks. They involved illustrations and discussions of interesting topics of a scientific nature such as Rocks, Forces, Nature, and so on. Unusual, perhaps, for the time was the use of simple materials, usually with an explanation of what to do with them and then, related questions. The illustrations were numerous and were of children doing the activity or simple, attractive cartoon-like figures. One of the problems involved in the development and implementation of any science program is that the teachers must be willing, and want, to use them. They, of course, must also have to be able to use them. Therefore, looking back through text books and programs, it would be useful to recall the times they were written for before some judgments are made. (Elizabeth Meng, personal communication, Feb 15, 2012)

In the 1960s, Willard and some colleagues published books intended for somewhat older children, with more content, but similar in approach; for example, a set of books entitled Thinking Ahead in Science (Jacobson, Lauby, & Konicek, 1965) and another called Field Research in Science (Jacobson, Kleinman, Hiack, Carr, & Sugarbaker, 1969). Willard was senior author for all these books. Several of the co-authors were or had been his doctoral students. The 1960s was the time of the “alphabet soup programs.” Some of the bestknown programs developed during this period for the elementary grades were Science—A Process Approach (SAPA), the Science Curriculum Improvement Study (SCIS), and Elementary School Science (ESS). They all were materials-rich

86   R. L. DORAN and A. B. BERGMAN

and activity-based. The equipment for two of these projects came as extensive kits. The three landmark programs had varied kinds and amounts of structure to them. SAPA was organized around specific scientific science skills process skills such as observe, classify, predict, and so on, while SCIS used broad organizations of explore, invent, discover, and the like. They were both sequenced for grade levels. Their instructional approaches varied in structure from prescribed to very minimally structured. ESS was less structured and more dependent on teacher intervention. It was the time of “child centered” activity. There was no scope and sequence for the ESS units/activities, and they were, in the opinion of many, more dependent on teacher understanding of the processes, content, and nature of “finding out” in science. With some guidance from the teacher, the activities could be adapted to “How could we find out?” then planning for a fair test and then following through. Therefore, the activities would involve the concepts, processes, and methods of science on an elementary school level. According to Meng, Willard’s approach to his subsequent textbooks was affected by his experience with these “alphabet soup” programs. How remarkable that he could work on multiple projects and be able to “switch gears” and help develop such varied projects while still being true to the best of science education (Elizabeth Meng, personal communication, Feb 5, 2012). Willard was a major contributor to the Science Curriculum Improvement Study (SCIS). This program was based on the structure of science, as seen by modern scientists. The research upon which the program was developed was consistent with the view of the nature of intellectual development of children and reflected the experience of elementary teachers working with trial versions of instructional units. Willard was one of the first to put forth the notion of “scientific literacy,” which he viewed as an awareness of the modes of inquiry in science and an understanding of the conceptual structure of science. Not unlike John Dewey, Willard felt that scientific literacy was essential for active participation in a democratic society, and, of course, an increasingly technological society. Willard described his idea of scientific inquiry simply “as the quest for a better understanding of the world in which we live.” In the SCIS Sourcebook (Jacobson & Kondo, 1968), Willard defined the experiences or principles in which children needed to be engaged in order to develop scientific literacy as: • • • • • •

Children need direct experiences with phenomena. Children should engage in investigations. Children develop their own conceptual structure of science. Guidance and discussion are integral parts of this process. Science activities lead children into additional science experiences. Scientific statements are considered to be tentative in nature.

Willard J. Jacobson    87

According to Herb Thier, a former administrator of the SCIS program: I first met Willard in 1958 when I was a doctoral student in School Administration at Teachers College. I transferred to NYU, but we remained friends and colleagues from then on. During the development of the Science Curriculum Improvement Study (SCIS) Willard was a valued member of our Advisory Board and author with Alan Kondo of the SCIS Elementary Science Sourcebook first published in 1968. This was revised and expanded in 1974 and became the SCIS Teachers Handbook. When we wanted to establish a SCIS Field Test Center in New York City, Willard suggested we and Teachers College cooperatively hire Mary Budd Rowe to run the center and teach at the College. A year or so later Mary asked me for an extra $500 to go to a learning and psychology conference, not usual for science education in those days. I ran it by Willard, and since she was rather new at the time we agreed it would probably be good for her professional growth, and so SCIS supported it. A very good investment, since out of it came the beginning of Mary’s work on “Wait Time.” When we wanted to open a center in Hawaii to expand SCIS westward, Willard was again very helpful in putting us in touch with Professor Albert Carr from the University of Hawaii, who was a former doctoral student of Willard’s. He also helped us convince NSF it would not be a “surfing center” since at the time there were no NSF centers in Hawaii. (Herb Thier, personal communication, Feb 5, 2012)

Based upon this work with SCIS, Willard went on to expand his ideas about teaching and learning science in a methods book entitled The New Elementary School Science (Jacobson, 1970). In this book, Willard proposed how to help children develop a “world view” by exploring the major concepts in the area of the universe and the solar system, the Earth on which we live, the air and the atmosphere, water and the hydrosphere, and the world of living organisms. He further defined the science processes that need to be promoted in an elementary school program, as well as science and its relationship to society. He expounded the importance of employing a variety of approaches in the classroom. No one approach suited all topics, all teachers, or all students. The Breadth and Depth of a Career of Publication Any comments about the more than 50 books written by Willard will merely be a sampling. Many of Willard’s books, intended for students, were published by the American Book Company (ABC), beginning in the late 1950s and continuing into the 1970s. There were some for the elementary school students and some for the junior high science program. Unfortunately, to

88   R. L. DORAN and A. B. BERGMAN

the uninitiated, each of these series of books was considered as an ABC Science Series, as they all were from the American Book Company, were about science, and had Willard as the lead, senior author. In 1959, a set of three books for the junior high school was published. The three books were entitled, Adventures in Science, Broadening Worlds of Science, and Challenges in Science (Jacobson, King, & Killie, 1959). In the Teacher’s Guide to the Adventures book, the authors wrote that “if possible, every science student should be given the opportunity to make systematic observations and to work with a definite problem—preferably a problem of his own choosing.” Students should receive guidance in “defining his problem, finding and evaluating sources of information, formulating his hypothesis, designing his experimental procedure, and evaluating his final results.” The authors cite the emphasis on science activities, stressing that “the study of science should lead to actual activities on the part of students.” All units contained suggestions for activities to be carried out by students. Specifically, for the unit on Our Solar System and the Universe, the Teacher’s guide reminds teachers that “direct observation of phenomena of the daytime and nighttime sky should be an important part of the student’s experiences in this area.” The Teachers Guide describes the textbook as “primarily a source book. It is set up to encourage self-direction in study.” An elementary school science book series, entitled the Thinking Ahead in Science series, was written by a team of authors, with Willard as the lead author. (Jacobson, Lauby, & Konicek, 1965). The titles of these six books were: Looking into Science, Learning in Science, Probing into Science, Searching in Science, Investigating in Science, and Inquiring into Science. For each grade level, a hard cover book was published, as preferred by some schools and teachers. In addition, a set of “topic books,” which were the individual units or sections of the grade level book, were also made available. For example, in the first grade book, Looking into Science, the topic books were entitled Rocks, Animals, Machines, and Fire and Temperature. After some information was presented via text and visuals, students were provided with several activities, labeled “Find Out!” In the Rocks topic or unit, the students were asked to make a collection of rocks and then sort these by color and texture. In the Inquiring in Science book (Grade 6), students were provided with an activity on the effect of grass on soil erosion and the testing of a “home made glacier” (soil, pebbles, and water frozen in a refrigerator tray) on some soft, sedimentary rocks. In 1975, Willard and co-authors published a six volume series of elementary science books entitled Investigations in Science (Jacobson, Victor, Bullett, Konicek, & Wong, 1975). The titles of the grade level books were Building Ideas, Comparing Ideas, Discovering Ideas, Exploring Ideas, Formulating Ideas, and Generating Ideas. As with the 1965 Thinking Ahead in Science series, there were hard cover, grade level books and “topic books,” which were a chapter

Willard J. Jacobson    89

or unit from the grade level book. To illustrate the change in instructional format from the earlier series, the Rocks, Minerals, and Soil topic book contained 14 investigations or activities for students to conduct. Each of these investigations was introduced by a question and followed by simple directions and photos to guide the student investigators to safe conclusions. In this section, the investigations included streak and scratch testing, making a conglomerate, a rock hunt near the school or home, and extended investigations in which students grew a salt crystal. There was a dominant theme of doing science to learn science. Willard was always eager to engage his students in collaborating with him in a variety of publications. For many, this would be an entrée into the world of publishing from which they might not otherwise benefit. In 1977, Willard asked Abby Bergman to co-author a science methods textbook with him for use in teacher education programs. Aptly titled, Science for Children: A Book for Teachers (Jacobson & Bergman, 1980), this volume saw several editions (1980, 1987, and 1990). It was through the preparation of these books that the major influences that impelled Willard’s thinking and practice at that stage of his career were revealed. In these volumes, the “big ideas” or broad generalizations in science were delineated: conservation of matter and energy, the second law of thermodynamics, and the law of mass production. Willard’s thesis here was that these big ideas were based upon the cumulative experience of the scientific community, and the understanding and appreciation of these ideas, along with practice in the processes of science, contribute to the scientific literacy of children and help to prepare them for participation in science and society. One of the innovative aspects of these books (at the time) was to draw the important connections among reading, language development, and science. Only years later would the educational community further attest to the importance of helping students to forge the links between science and literacy. A chapter in this book also dealt with science and the “exceptional child,” one of the first such texts to deal with this specific aspect of science instruction. In 1983, Willard and Abby co-authored the activity manual Science Activities for Children (Jacobson & Bergman, 1983). In this volume they outlined a host of science activities for elementary school students in the major themes of science. It is interesting to note that this book appealed to parents and students as much as to teachers. Because of Willard’s great concern for many of our world-wide problems, he became involved in population education. His book, Population Education: A Knowledge Base (Jacobson, 1979), grew out of a related seminar at Teachers College. It includes accurate and extensive background material and suggestions for discussions, as well as very interesting and appropriate activities for students. This was followed with suggestions for further

90   R. L. DORAN and A. B. BERGMAN

research. From this, he felt that students would more likely be able to make informed decisions about their future and the environment. Willard collaborated with other Teachers College faculty members in preparing the manuscript Promising Practices in Nutrition Education in the Elementary School (Jacobson, Boyd, & Hill, 1959). This monograph was a result of the Nutrition Education Research Project at Teachers College. Willard—A Leader of Leaders: Willard and Professional Associations Other remarkable contributions were Willard’s leadership roles in several professional science education organizations. The quality of his leadership might be measured by the number of organizations, the length of his involvement, or the level of his leadership (e.g., elected positions). What is particularly notable is not only Willard’s membership in these organizations, but his assumption of leadership roles in their operation and impact upon science, research, and teacher education in the United States. Willard must have “marched to a different drummer,” being a member, an active participant, and leader in almost all the major science education professional associations. He received many honors and awards from these groups for his vision and leadership. Willard would likely have viewed such service as thanks for what he received from these groups. Given his position at one of the premier educational institutions, Willard sought involvement and direction for many significant initiatives over the years. Some of his professional involvements, in no particular order, follow: National Association of Research in Science Teaching (NARST). Willard was a lifelong member and active participant in NARST, its conferences, and its journal. Willard was elected President of NARST, serving the academic year, 1969–1970. Joe Novak shared his observations of the early years of NARST and his interactions with Willard. When I first joined NARST in 1958, Willard was one of a dozen or so senior members who went out of his way to welcome me to the organization. In those days, annual meetings were held in February, either in Chicago or an Eastern city, with some 40 to 50 members in attendance. Admission to NARST was done by a vote of the membership in attendance at the meetings. NARST membership was about 70. From this first meeting onward, Willard was one colleague I could count on for support and encouragement. It was common for some senior members to sharply quiz younger participants during their presentation, so it was a relief to have Willard’s encouraging comments when I gave papers. (Joe Novak, personal communication, Oct 21, 2011)

Willard J. Jacobson    91

National Science Teachers Association (NSTA). Willard presented countless workshops, panels, and speeches at the national and regional conferences of NSTA. Bob Yager recalls one of Willard’s contributions to NSTA. In 1981, I was President-Elect of NSTA. One of my first efforts was to propose a Horizons Committee charged with study of the future needs and directions for science education. Willard was one of the ten asked to serve on that committee. The committee provided reports of research findings to recommend new directions for the field. Willard was a model thinker and provided many ideas for the Horizon Committee. (Bob Yager, personal communication, Oct 28, 2011)

Willard’s contributions to NSTA were recognized by its 1977 Citation for Distinguished Service to Science Education, and its 1987 Robert Carleton Award (the highest recognition given by NSTA). New York Academy of Science (NYAS). Instrumental in founding the Science Education Section of NYAS in the 1970s, Willard served as its first chair and was an active member until his retirement. Willard invited many of his graduate students to join the Academy, and some of them become leaders of the Science Education Section. June Miller was chair of the section and described how their first meeting of the year is now called the Willard Jacobson Forum in Science Education (June Miller, personal communication, Feb 3, 2012). American Association for the Advancement of Science (AAAS). Willard was elected a vice president of the American Association for the Advancement of Science (AAAS) and in 1968 was recognized as a fellow of the Association. The Joseph Priestly Society. This organization of scientists and science educators was initiated in the 1980s through Columbia University’s Physics Department. The mission of the society was to promote interactions among university faculty, high school teachers, and science museum administrators. The society organized discussions and seminars at Columbia about optimal ways to teach science and the important role of hands-on experiments for students. Willard took an active role in the society and urged Abby Bergman to join him leading the conversations about the future of science education in the United States and in the larger international arena. Council of Elementary Science International (CESI). Willard was a president of CESI. This was certainly a reflection of his dedication to the teaching of science in the elementary school and the early and long-term involvement of Teachers College science education faculty with CESI. At the 1932 national meeting, the journal Science Education was adopted as the official journal of the group (Burchett, 1985). The affiliation with NSTA was formally adopted in 1969.

92   R. L. DORAN and A. B. BERGMAN

Association for Science Teacher Education (ASTE). Willard was elected president of the group, then called the Association for the Education of Teachers in Science (AETS), for the 1962–1963 academic year. He was the guiding force behind the Eastern section (later, Northeastern) activities for decades. In 1977, Willard wrote “A Brief History of ASTE,” citing its origins in the late 1920s, with meetings organized by Teachers College science education faculty Gerald Craig and S. Ralph Powers. (Jacobson, 1977). In 1959 AETS began its affiliation with NSTA. Willard was recognized by AETS in 1994 with its Honorary Member Emeritus. Willard and the Second IEA Science Study According to John Keeves (International Coordinator of SISS), in 1980 it was proposed that “a second IEA Science Study should be undertaken with the Australian Council for Educational Research (ACER) as the International Coordinating Centre of the study and with Willard Jacobson as chairman of the International Study Committee. This study was planned as a replication and extension of the first study (undertaken in 1970–1971) to report on the changes over a period of 14 years in science education that were of great interest and importance to the countries involved” (John Keeves, personal communication, Feb 26, 2012). Also, Willard was the National Research Coordinator for the U.S. involvement in this study. He managed to keep this large, multi-year project moving along efficiently. In addition to all the in-house work, Willard hosted a meeting in August 1983 of all the SISS national research coordinators at Teachers College. The team of researchers from the 23 countries that participated in the Second International Science Study met once a year to do all the planning, drafting, pre-testing, and establishing procedures for use in these countries and many languages. As part of the Second IEA Science Study, each country was encouraged to form a National Advisory Committee comprised of leaders from the field. Willard suggested that this committee consist of the presidents of the several American science education organizations: NSTA, AETS, NARST, NSSA, CS3, and CESI. This group included Hans Anderson, Stan Helgeson, Don McCurdy, Ken Mechling, Doug Reynolds, and Essie Beck. Bill Aldridge also was part of this committee, as he was the executive director of NSTA at that time. One of the more innovative approaches included in the SISS was the “practical skills testing” (or performance assessment) with the fifth grade (age 10) and ninth grade (age 14) samples. Willard decided that the U.S. would participate in this optional component with five other countries (Hong Kong, Hungary, Israel, Japan, and Singapore). Pinchas Tamir was

Willard J. Jacobson    93

the experienced and enthusiastic leader of this research, building on his experience with lab skills testing, especially for Israeli high school biology courses. Tests were developed (trial tested and revised) to match the goals and approaches in all these countries, data were collected and analyzed, and results and conclusions compiled (Doran & Tamir, 1992). While this research was interesting and significant in its own right, it had a special impact in New York State. Doug Reynolds, a member of the SISS national advisory committee, was the chief of the Bureau of Science Education of the New York State Education Department. According to Doug Reynolds: I had been engaged in implementing such an approach statewide in high school Earth Science, but to try to do the same at the elementary and middle school levels across the whole State of New York tingled my imagination. Building upon “the trials and tribulations” of the SISS project, the new performance assessment was implemented statewide in May 1990 (all fourth grade students in public schools) with the full support of the Board of Regents. That form of assessment continues today, after over 20 years. Similar performance testing at grade eight was also implemented (beginning in 2001) and continues to this day. Willard’s leadership had a profound impact on the education of children across this state, with over 200,000 fourth grade students taking the performance assessment annually, and another 200,000 eight grade students annually doing the same. (Doug Reynolds, personal communication, Feb 5, 2012)

Malcolm Rosier, the International Coordinator of SISS, commented that: Individual countries like the U.S. worked on their national data sets to produce reports for their own audiences. In Willard’s case, he was also able to harness the energy of many doctoral students. (Malcolm Rosier, personal communication, Nov 12, 2011)

The strong American involvement of graduate students and staff in the direction (in addition to the details) of the study, and authorship of monographs, articles, and reports documenting the study’s procedures, results, and recommendations was unique and healthy. While clearly Willard was the recognized international scholar, he was always a “team player.” Willard treated each as equals on the team and preferred being addressed as “Willard,” not Professor Jacobson or Dr. Jacobson. Many of these doctoral students became co-authors of some of the SISS reports and monographs. One of these Teachers College doctoral students, Marilda Chandavarkar, received the International Association for the Evaluation of Educational Achievement (IEA)’s award of the year for best research done with their data.

94   R. L. DORAN and A. B. BERGMAN

Willard—The Mentor and Model of Scholarship Willard was the very model of scholarship and served as a mentor to countless graduate students. A longtime colleague of Willard, O. Roger Anderson, offered the following perspective: When I arrived at Columbia University Teachers College in 1964, Willard, though just over a decade after earning his doctorate, was already well on the road to international visibility as a science educator of broad vision, having served with AID in Afghanistan as one of the American education advisors and taking leadership in enhancing the breadth and intellectual foundations of elementary science education. He was a devoted advocate for science education reform, a forceful and cogent public speaker, and a loyal and honestly critical colleague who held leadership roles in major science education societies. He contributed to significant improvement of science education curricula, especially at the elementary level. He was uniformly respected and revered by his students and faculty colleagues. His legacy is written in the enduring contributions he has made to the profession and the lives of those who he touched through his dynamic role as teacher and mentor. (O. Roger Anderson, Dec 1, 2011)

Willard interacted with the hundreds of graduate students who took his courses and carried that experience to their own teaching and with the scores that completed a master’s thesis or doctoral dissertation “under his wings.” Many of these individuals co-wrote articles and books with Willard. This guided experience often became the precursor for subsequent individual efforts with many advisees. Willard had a keen eye for developing skills and talents in his students. In addition to his generous doses of encouragement, he had a well-developed sense for where each of his students and associates might make the greatest contributions. Many were drawn to Professor Jacobson by his obvious care for his students. Several of the individuals who wrote to us expressed that they always had the feeling that he cared about them as people and wanted them to achieve their fullest potentials. When it became time for students to do their year of residency at Teachers College, Willard worked intensely to find ways to support their study, whether it was supervising student teachers, teaching a methods course, or co-writing a book. Former students commented that Willard knew his strengths and also knew when something called for someone else’s area of expertise. Willard set high standards of scholarship, collegiality, and interpersonal relationships. Everyone, even beginning grad students, was treated with respect. Willard was always willing to listen and give a respectful response or correction, as needed.

Willard J. Jacobson    95

The Person Behind the Scholar It seems quite appropriate to include comments on Willard the person, beyond his professional and academic influence. His physical appearance did not detract from his intellectual presence. Malcolm Rosier (International Coordinator of the IEA Science Study) noted that “Willard was a stately and friendly person, with a crop of white hair that made me jealous as mine receded. Willard was older than most of the other national representatives, and added a gravitas to the SISS proceedings on that account” (Malcolm Rosier, Nov 13, 2011). To some, Willard always seemed to be a gentleman in the old-fashioned sense. Others commented that Willard was usually nattily dressed and appeared like a kind, soft spoken British Ambassador. Others were impressed by the competence, loyalty, and commitment of the people who worked with Willard and said Willard’s leadership made him seem like one of the nation’s premier field marshals of science education. A former advisee recalled Willard as a tall man, with blue eyes, magnificent white hair and pink complexion—a color pattern that we sometimes joked about while saluting him. Others recalled his big smile, deep and commanding voice, crisp white hair, always in place, and described him as tall and strong, quite ageless as a wizard of fairy tales. Willard brought about a respect and admiration such that some chose to address him as “Professor” not as “Willard.” Despite usually representing the United States, Teachers College, Columbia, or some prestigious organization, Willard always was an avid listener and learner. Some described how Willard would silently listen to each presentation and the debates among the participants that followed. When all had had their say, Willard would take the group’s diverse views and synthesize them, with grace, into a coherent single position that all would agree upon. Several colleagues remarked how Willard had an intuitive knack for choosing the right people to work on projects. He had this sixth sense about people, knowing who would fit and work well with the group. Willard chose the best students—the cream of the Teachers College crop. Then, Willard provided a rich environment in which they could thrive. Willard always welcomed American and international students into his “Teachers College family.” His apartment, just two blocks form the college, was often crowded with his doctoral students. Some recalled evenings there as the scene of lively discussion and debate, and a sharing of cultures and perspectives. The ever-gracious Carol entered into these conversations, all the while providing welcome refreshment and warm hospitality. Willard and Carol also hosted students at their summer home in Pine Plains, New York, where Willard enjoyed leading hikes around the area. Several former students remarked how they were made to feel like they were partners in his work. They felt like members of his family of students.

96   R. L. DORAN and A. B. BERGMAN

They learned much from this man, especially by his modeling as a person and a mentor. Willard’s wisdom affected them in ways that they still feel today. Final Remarks While science in the elementary schools may forever struggle for a bigger share in the curriculum, Willard’s work via textbooks, methods books, activities books, and his leadership involvement with the elementary science curriculum projects such as SCIS certainly kept it alive. Willard’s work to conduct and complete the U.S. part of the Second IEA Science Study (SISS) laid the groundwork for the Third IEA Math and Science Study (TIMSS) and other cross-national ventures. While international studies may generate as many (or more) questions than they answer, it is critical to focus on concrete, quantitative indicators of academic achievement in a variety of countries. We need to know as much as we can so that we can prepare our students for the future, permeated by science and technology and connected via the global economy. Several former students recalled how very busy Willard was all of the time! As one of the leaders in the world of science education, he hosted many world-wide, country-wide and city-wide meetings. In many cases, he often arranged evening activities for them. And still he got up at 5:30 every morning to do his writing for whatever publication he was working on at the time and somehow also managed to make time to play tennis. Some commented it is hard to catalog his legacy, for his was so complex a personality, with so many facets. Like good wine, or precious things that one has lost, our appreciation of Professor Jacobson, or Willard, as he so humbly allowed everyone to call him, grows deeper and fonder with time. Colleagues remarked that Willard was a good friend, a respected colleague, and a valued advisor. According to Bob Yager, himself a powerhouse in science education, “Willard Jacobson was one of the greatest of Science Educators ever—indeed a Giant in our field” (Bob Yager, personal communication, Oct 28, 2011). References Burchett, B. (1985). A brief history of CESI. Doran, R. L. & Tamir, P. (Eds.). (1992). An international assessment of science practical skills. Studies in Educational Evaluation, 18(3), 263–406. Jacobson, W. J. (1951). Science education and the development of abilities to cope with problematic life situations. Unpublished Ed.D. dissertation, Teachers College, Columbia University, New York, NY. Jacobson, W. J. (1970). The new elementary school science. New York, NY: Van Nostrand Reinhard Company.

Willard J. Jacobson    97 Jacobson, W. J. (1977). A brief history of ASTE. Jacobson, W. J. (1979). Population education: A knowledge base. New York, NY: Teachers College Press. Jacobson, W. J. & Bergman, A. B. (1980). Science for children—a book for teachers. Englewood Cliffs, NJ: Prentice-Hall Inc. Jacobson, W. J. & Bergman, A. B. (1983). Science activities for children. Englewood Cliffs, NJ: Prentice-Hall Inc. Jacobson, W. J., Boyd, F. L, & Hill, M. M. (1959) Promising practices in nutrition education in the elementary school. New York, NY: Teachers College Press. Jacobson, W. J., King, R. N., & Killie, L. E. (1959a). Adventures in science. New York, NY: American Book Company. Jacobson, W. J., King, R. N., & Killie, L. E. (1959b). Broadening worlds of science. New York, NY: American Book Company. Jacobson, W. J., King, R. N., & Killie, L. E. (1959c). Challenges in science. New York, NY: American Book Company. Jacobson, W. J., Kleinman, G. S., Hiack, P. S., Carr, A. B., & Sugarbaker, J. S. (1969). Field research in science (A textbook series). New York, NY: American Book Company. Jacobson, W. J. & Kondo, A. (1968). SCIS elementary science sourcebook. Berkeley, CA: The Regents of the University of California. Jacobson, W. J., Lauby, S. J, & Konicek, R. D. (1965). Thinking ahead in science (a textbook series). New York, NY: American Book Company. Jacobson, W. J., Victor, E., Bullett, M. A., Konicek, R. D., & Wong, H. H. (1975). Investigating in science (A textbook series). New York, NY: American Book Company.

98   R. L. DORAN and A. B. BERGMAN .

Chapter 6

Mary Budd Rowe What a Researcher Can Say to Science Teachers Julie A. Bianchini University of California, Santa Barbara Nicole I. Holthuis Because Science Matters Consulting

Acknowledgements This chapter is based on a previously published article on the life and work of Mary Budd Rowe written by Julie Bianchini in 2008 for the journal Cultural Studies of Science Education. We thank the former graduate students and colleagues of Mary Budd Rowe who took the time to share their favorite stories of her. We are particularly indebted to the helpful suggestions of Francis Lawlor and Thomas Keating.

Going Back for Our Future, pages 99–121 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

99

100    J. A. BIANCHINI and N. I. HOLTHULS

Mary Budd Rowe: What a Researcher Can Say to Science Teachers Mary Budd Rowe’s contributions to science education, just like the scientific enterprise itself, include both content and process. There is the content of Rowe’s work: her examination of the teaching and learning of science as inquiry; her identification of wait-time as central to productive classroom conversations; her drawing of connections across students’ lives, sense of fate, and science learning; and her exploration of ways new technologies could transform science curriculum and instruction. But there is also and, perhaps, more importantly, a tale of process. Rowe spent her career honing the simple yet powerful messages of science as story-making and of science teaching and learning as continuous inquiry. She then spread these messages to those she thought mattered most: teachers, in particular, elementary school teachers. It is indeed rare to find a scholar like Rowe whose research, teaching, and professional activities so closely aligned. We both worked with Mary Budd Rowe while young science education graduate students at Stanford University in the early 1990s. We entered Stanford with the knowledge that Rowe was a legend in the science education world—that her name was instantly recognized and her work revered. Over time, we came to see Rowe as an unassuming and approachable teacher and mentor. Nicole recalls meetings with Rowe that felt much like informal chats. Rowe expertly wove engaging and sometimes humorous stories culled from her vast experience into the fabric of meetings. Tucked within these narratives were important messages about inquiry, the nature of science, and teaching in general. We also came to recognize Rowe’s deep passion for science education—for how she breathed, ate, and slept science. Julie remembers Rowe as constantly in demand because of this passion. Meetings in her office were routinely punctuated by phone calls from other science education researchers, teachers, and policymakers (as well as by the eating of a granola bar as a late lunch). Rowe always wore two watches to help her manage her frequent travel to consult or collaborate: She set one of these watches ahead an hour each day for several days before a trip east to avoid jet lag once she arrived. Still, over 20 years after our first interactions with Rowe, we found the opportunity to reexamine her work and reposition her ideas in the larger history of science education to be very eye opening. As we examined her contributions en masse—her publications, her collections of videos and CD-ROMs, and her files stuffed with correspondence from a who’s who list of science education—we were struck by how clear, ubiquitous, and relevant her messages remain. Science is not a long list of dictionary definitions, rote facts, or right answers. Science is exploring, and exploring is fun.

Mary Budd Rowe    101

Rowe’s Storyline: Science As Story-Making Across four decades of scholarship, Rowe (as cited in Patrick, 1992, para. 20) underscored the need to understand science as “a special kind of story-making.” Science as story-making, Rowe elaborated, conveys the ideas of beginning with wonder, investigating questions, arguing from evidence, and proposing best-at-the-time explanations. “There are no right and wrong answers [in science], just better and better stories,” Rowe emphasized. “You live with the best story you have at the moment.” When students experience science as a story of excitement and adventure, Rowe concluded, “it changes the way [they] think of the world.” (See Bianchini, 2008, for further discussion.) Rowe rarely missed an opportunity to engage others in science as storymaking—to invite students, teachers, or colleagues to investigate simple natural phenomena with her. In 1979, for example, Rowe served as a member of a U.S. delegation to China sponsored by the National Academy of Sciences. J. Myron Atkin, now a professor emeritus at Stanford University, recounted how Rowe encouraged elementary students to explore themselves, their classroom, and, ultimately, him using hand lenses: By 1979, the Chinese were desperate to reconstruct science education. So they invited the National Academy of Sciences to put together a delegation who would come over for two weeks and visit schools in China and make recommendations to the Chinese about how their science education might be improved. Thirteen of us went. Paul Hurd chaired the delegation. Jim Rutherford was on it. And Mary Budd was on it. . . . [One day,] Mary and I were in an elementary school classroom. Mary had the children involved in looking at things through a hand lens. She looked at me; I was wearing short sleeves. “Take your hand lens and go over and look at Professor Atkin’s hairs on his arms.” The kids came all around me with their hand lenses to look at my arms. She got a good chuckle at that. (personal communication, September 6, 2007)

Equally important, as Nicole noted in the introduction, Rowe regularly infused her own stories about science and science education into informal conversations with doctoral students, scripted public presentations, and scholarly publications. Our favorite story told by Rowe was about the centrality of exploration in science. It was a strange sight: a man, standing before a fountain, watching the falling water and tilting his head from side to side. Drawing closer, I saw he was rapidly moving the fingers of his right hand up and down in front of his face. I was in seventh grade, visiting Princeton University with my science class, and the man at the fountain was Albert Einstein. For several minutes, he continued silently flicking his fingers. Then he turned and asked, “Can you do it? Can you see the individual drops?” Copying him, I spread my fingers and

102    J. A. BIANCHINI and N. I. HOLTHULS moved them up and down before my eyes. Suddenly the fountain’s stream seemed to freeze into individual droplets. For some time, the two of us stood there perfecting our strobe technique. Then, as the professor turned to leave, he looked me in the eye and said, “Never forget that science is just that kind of exploring and fun.” Nearly half a century later, I’ve spent an entire career trying to impart Einstein’s words to adults and children all over the world: Science is [emphasis in original] exploring, and exploring is fun. (Rowe, 1995, pp. 177–178)

Because of her encounter with Einstein, and, perhaps, also because of her own keen sense of humor, Rowe’s stories portrayed science as playful and fun. Rowe’s views of science learning as changing how students understand and experience their world and of science teaching as promoting wonder, adventure, and playfulness clashed with the realities of science education in U.S. schools. Rowe routinely criticized traditional science instruction. She argued that science teaching should not be equated with the mere conveying of science content: For the most part, teachers . . . concentrate on the question “What do I know?” Emphasis on this question tends to limit the teacher’s function to one of conveying information and correcting student recitations. . . . The students’ primary responsibility is to learn the official story well enough to be able to write it or recite it correctly, regardless of whether they understand it or believe it. Thus the coverage of large amounts of content becomes a primary objective. (Rowe, 1983b, p. 126)

Would it not be better for both teachers and students, Rowe countered, if science were taught and learned as inquiry? The structure and substance of science textbooks, Rowe continued, needed revision as well. “The plot, the storyline—the way in which ideas interact— have disappeared” from science texts (Rowe as cited in Patrick, 1992, para. 19). Student do not want to read texts that are “as dull as dictionaries” (Rowe, 1995, p. 178), she often underscored. Equally troublesome, the size and complexity of texts supported the teaching of science as a list of terms and facts to be mastered rather than as a series of questions to discuss and explore. Science books have grown by accretion, packed with more concepts per page, more pages per book, and more topics to be “covered” than ever before. High-school science texts average between seven and ten new concepts, terms, or symbols per page. Typically, the 300 to 350 pages assigned during a school year means that students are expected to learn between 2,400 and 3,000 terms and symbols per science course. Thus, in a school year . . . twenty concepts would have to be covered per period, an average of one every two minutes. That there is very little discussion or inquiry should come as no surprise—there is not time for it. (Rowe, 1983b, pp. 126–127)

Mary Budd Rowe    103

Rowe applied this same critical lens to educational programs and video series that could also convey the wrong storyline about science. Her files include scripts on which she wrote detailed comments and edits. One video producer, for example, made the mistake of including a teacher instructing students in the script: “By the end of this unit you can tell who had the correct [prediction] and who didn’t.” Rowe’s written response was direct and unequivocal. No, no, no. It is the wrong message. The children make predictions. Then the task is to investigate together the relationships. They are to find out how something works. The focus should be on the process and not on who is right and who is wrong. If you put the emphasis there, the kids stop thinking as soon as they get some kind of result. (I have watched little children who had to make predictions and mark them on their workbook . . . then, when things did not work the way they expected, quickly erase the predictions and change them.) Heavy emphasis on right or wrong and who is better than whom is too much a part of school. No matter how an experiment turns out, either predicted correctly or not, they must still confront the problem’s explanation [emphasis added]. (personal communication, 1994)

Rowe’s Early Years: Teaching Science as Continuous Inquiry Rowe’s passion for science and commitment to science teaching emerged early. According to Emily Girault, an old friend and now professor emeritus at the University of San Francisco, Rowe “‘was very aware of how she had been encouraged by some teachers early in her life.’ A junior high teacher, in particular, knowing that Rowe was blind in one eye, went beyond the curriculum to encourage her to study vision and the physics of light” (Science education innovator Mary Budd Rowe dies at 71, 1996, para. 7). Rowe’s chance encounter with Einstein at Princeton, we add, also made a lasting imprint (as well as a memorable story). Rowe earned a bachelor’s degree in biology and physics education at New Jersey State University in 1947. She later received a master’s in zoology from the University of California, Berkeley in 1954. Rowe taught science at a private, all-girls secondary school in La Jolla, California from 1948 to 1953, as well as at military schools in Germany from 1955 to 1958. During her tenure as a consultant in science and mathematics for the State of Colorado in the late 1950s and early 1960s, Rowe traveled in a 40-foot trailer—a fully equipped teaching laboratory on wheels—to rural areas of the state to demonstrate ways to investigate nature using hands-on activities. In 1964, Rowe earned her doctorate in science education from Stanford University. Her doctoral work was completed under the mentorship

104    J. A. BIANCHINI and N. I. HOLTHULS

of another giant in the field of science education, Paul DeHart Hurd. In her dissertation, The Influence of Context-learning on Solution of Task-oriented Science Problems Which Share Concepts: A Study in Elementary Science Education, Rowe (1964) investigated how 60 first grade boys approached and solved problems in magnetism across six different instructional contexts. Distilling her 102-page dissertation into a mere seven, she published her first singleauthored article in 1965 (Rowe, 1965). Rowe’s attention to keeping her message short persisted throughout her career. As both a graduate student and newly minted PhD, Rowe was involved in designing, disseminating, and researching the post-Sputnik elementary science curriculum projects funded by the National Science Foundation (NSF). Rowe worked with Robert Karplus, a physicist at the University of California, Berkeley, for example, on the Science Curriculum Improvement Study (or SCIS) elementary program. She also served as an advisor in the development of Science—A Process Approach, Elementary Science Study, and the Biological Sciences Curriculum Study: Biological Sciences. Rowe drew from her interests in group dynamics, change agent behavior, and engineering systems analysis to develop and implement intensive two-week professional development workshops on these curriculum materials (R. Hannapel, personal communication, February 4, 2008). Teams of teachers, administrators, science educators, and scientists from across the country came together to conduct hands-on investigations during these workshops and then organized similar workshops for colleagues back in their local contexts. Further, Rowe published several studies on teacher professional development and classroom implementation of these NSF-sponsored curricula, often with her advisor, Hurd (see Hurd & Rowe, 1964, 1966; Rowe, 1969a, 1971). While an assistant professor at Teachers College, Columbia University, Rowe initiated Science and the Inner City, an eight-year teacher professional development and curricular innovation project. Rowe worked at each of five elementary schools in Harlem to conduct professional development workshops for teams of teachers on NSF-funded curriculum materials and inquiry-oriented instructional approaches; to provide their classrooms with inexpensive, everyday materials for science lessons; and to support them when teaching science as inquiry to their students. Rowe’s first graduate student, Francis X. Lawlor, worked with Rowe on this Science and the Inner City project. A central lesson Rowe attempted to share with teacher participants, Lawlor noted, was the need to listen to and learn from their students—a recommendation still routinely offered today. Mary did not believe in the use of textbooks. . . . She thought that the kids were the teacher’s textbook. The teacher was doing inquiry with the children as they did inquiry with the materials. It was up to the teacher to learn from the students and from what difficulties they were finding. (personal communication, September 21, 2007)

Mary Budd Rowe    105

A second lesson Rowe underscored to these teachers was the need to improve the quality of discussions between their students and themselves. Rowe thought students more likely to look to their teacher for right answers than to use evidence grounded in materials to support their claims. Particularly in an inquiry classroom, given the demands placed on students when conducting investigations, students needed to learn to converse rather than to recite, and to give priority to evidence over authority. Children need to monitor their materials more carefully than they monitor the teacher’s face. Ideas can be modified or even discarded if the evidence requires. No particular point of view in the class is more sacred than another. What counts is what happens in the system of materials. Authority rests with the idea that “works.” (Rowe, 1969b, p. 12)

Indeed, Rowe’s examination of teacher-student interactions in these Harlem classrooms helped lead to and served as one site for her groundbreaking research on wait-time. In later years, Rowe turned her attention back to these post-Sputnik curriculum materials in an effort to preserve them for future generations. Rowe’s Major Contributions to Science Education: Science that Kids Can Live By The list of Mary Budd Rowe’s contributions to science education is long. We focus our discussion here on those three we think most central: waittime, technology in education, and science teacher education. Wait, Wait, Wait . . . First, Rowe remains widely recognized for her pioneering research on wait-time (see, for example, Rowe, 1974a, 1974c, 1974d, 1974e, 1986, 2003). Simply put, Rowe found that both the length and the quality of students’ responses increase when teachers give them more time to answer questions. Teachers typically wait less than one second before and after a student responds to a question; increasing the wait to at least three seconds enhances the language and logic of students’ answers. Indeed, Rowe identified two types of wait-time needed to improve teacher-student conversations: Waittime one is the time after a teacher asks a question; wait-time two is the time after a student responds. Rowe systematically investigated wait-time across small groups and whole classes of elementary students using hands-on, inquiry-oriented curriculum materials (see, again, Rowe, 1974a, 1974c, 1986). The idea for wait-time

106    J. A. BIANCHINI and N. I. HOLTHULS

had emerged from her examination of hundreds of tapes of classroom interactions culled from two research projects conducted in the late 1960s; across these tapes, a handful surfaced where teachers’ pacing of classroom discourse was slower and the quality of students’ responses, higher. In her studies of wait-time, Rowe employed a servo-chart plotter to track the speech, pauses, and silences of teachers and students; in most cases, teacher-student interactions proceeded so quickly that a stopwatch was found to be inaccurate. In classes where teachers did not attend to wait-time, whole class discussions consisted of a series of rapid-fire teacher questions, brief student responses, and teacher repetition of student ideas. Discussions led by teachers who had learned to extend the time they waited before and after questions, in contrast, looked markedly different: Teachers changed the number and kinds of questions they asked; more students participated in whole class discussions; and individual students were more likely to provide long responses, use evidence to support their answers, pose their own questions, and interact with each other. Although a simple and straight-forward idea, Rowe viewed extended waittime as critical to supporting and enhancing the teaching and learning of science through inquiry. She envisioned students as having adequate time to construct “sustained arguments based on multiple sources of evidence” in conversations with themselves and their teacher (Rowe, 1969a, p. 33). To “grow” a complex thought system requires a great deal of shared experience and conversation. It is in talking about what we have done and observed, and in arguing about what we make of our experiences, that ideas multiply, become refined, and finally produce new questions and further explorations. (Rowe, 1986, p. 43)

A number of other researchers took up Rowe’s wait-time concept and studied it in a wide range of contexts and grade levels. Rowe hoped her fellow researchers would not “mak[e] wait-time into a rigid formula that could produce major results in isolation from the whole student inquiry process” (Lawlor, personal communication, September 21, 2007). Further, despite its simplicity, studies of teacher training and implementation have made clear that wait-time is a difficult concept for teachers to incorporate into their everyday practice (Rowe, 1986). Riding a Beam of Light: Recognizing the Power of Technology Rowe was recognized as an innovator in the use of technology in science education. She purchased one of the first portable video recorders, a Sony camera, to capture classroom data for her wait-time research in the 1960s.

Mary Budd Rowe    107

Reflecting her view of science as adventurous, playful, and exciting, in later years, she served as a science advisor for the children’s television programs Reading Rainbow, 3-2-1 Contact, and Voyage of the Mimi. Rowe herself developed a video series to help elementary teachers better teach physical science concepts using inquiry; this is discussed in detail in the next section. Further, in the 1980s, she employed at that time cutting-edge CD-ROM technology to create a searchable database of science curriculum materials. More specifically, Rowe was a strong advocate of using technology to teach students. In the 1970s, Rowe (1978a) encouraged science teachers to examine existing research on student learning as a first step to integrating soon-to-arrive technologies into their teaching. She called for science teachers “to be in the vanguard” of this technological revolution. We stand on the brink of a major technological revolution that will have tremendous impact on education, particularly in science and mathematics. Technologies related to microprocessors have changed radically, even in the past two years, along the lines of miniaturization, speed of processing, and reduced costs . . . .Increasingly sophisticated laboratory simulations will be possible. Thus, time will be available in larger chunks, with less wasted on onerous calculations and recalcitrant equipment. (Rowe, 1978a, p. ii)

Rowe continued to recommend the infusion of technology into science teaching—by the late 1980s, such technologies included CD-ROMs, hypertext software, databases, spreadsheet programs, and electronic laboratory probes—so that students could practice science as creatively as real scientists do. She envisioned science classrooms where students used computers and other technological tools “to make meaningful explorations . . . draw valid, reliable conclusions . . . and perhaps most important of all, raise more questions for new explorations” (Rowe & McLeod, 1988, p. 73). Rowe (as cited in Patrick, 1992) also argued that teaching science using technology would help to level inequities that persisted in classrooms. Rowe saw technology not only as a tool for science teaching, but as a way to support teachers in their own work. In the 1980s, for example, Rowe (1987) developed the Science Helper K–8 CD to provide a searchable database of approximately one thousand lesson plans from out-of-print or hardto-find NSF-funded elementary science curriculum projects from the 1960s and 1970s. Rowe undertook this project because she worried that many of these potentially rich inquiry resources—curricula she had helped design, disseminate, and research—would be lost to future generations of teachers and students if not collected, organized, and preserved. Lawlor described the Science Helper project in detail: A major effort by Mary involved the preservation of the final products of all the years [of post-Sputnik, NSF-sponsored curriculum development]. She

108    J. A. BIANCHINI and N. I. HOLTHULS proposed archiving this material on some sort of new device [a CD-ROM disk]. When Mary called me from Washington at some time in the eighties, as I recall, very excited about a new technology developed by the Navy for access to super compact digitized equipment manuals on warships, I thought that she had gone “round the bend.” She had seen a demo of this radical storage technology at some military research center and immediately decided that it would be a fantastic educational tool for curriculum development and for classrooms. Thousands of pages of information on some sort of disk?...It was an idea very ahead of its time. . . . Mary introduced the first elementary school Science Helper CD at a time when no school had a means of playing one. She got a workshop funded to introduce school superintendents and science coordinators to the materials and had to supply each one with a CD player to be plugged into a computer. (personal communication, September 21, 2007)

Rowe, we emphasize, was one of the first outside of the military to use CD-ROM technology and one of the first to design an interface to allow users to sift through digital material. Although common today, in the 1980s, few had heard of CDs. In addition to the Science Helper CD, Rowe also developed a Culture and Technology CD (1996a) for similar out-of-print, NSFfunded social studies curricula for grades 5 through 12. The Science Helper CD is still published by the Learning Team (see http://learningteam.org/ index.html); the Culture and Technology CD is no longer available. Rowe’s interests in technology led her to work on electronic teacher communities in the 1990s. Rowe, for example, established and researched a computer network for teams of middle school teachers involved in the design and pilot testing of an integrated curriculum developed at Stanford University, Human Biology. Thomas M. Keating, a former graduate student of Rowe’s and collaborator on technology used in the Human Biology project, elaborated: On the technology side, Mary had an uncanny ability to anticipate where emerging technologies were headed and their implications for teaching and learning. In the early 90s, we worked together on a NSF-funded project connecting teachers engaged in the Stanford-based HumBio project via telephone and modem to form electronic communities. Early on in our project, a graduate student in the Computer Sciences introduced us to the NCSA [National Center for Supercomputing Applications] Mosaic web browser running on a Sun SPARC [from Scalable Processor Architecture] workstation. Mary’s mind was off and running on the potential of the nascent World Wide Web to connect students and teachers globally to participate in meaningful science. (personal communication, February 13, 2012)

While chat rooms, blogs, and on-line courses are plentiful today, at that point in time, encouraging teachers separated by grades, schools, and states to communicate with one another using computers was innovative.

Mary Budd Rowe    109

Research Can Help You In addition to being a skilled researcher and innovator, Rowe was a persuasive teacher educator. She was a staunch advocate both of sharing key research findings with teachers and of encouraging teachers to conduct research in their own classrooms. For Rowe, her message of science as storymaking and of science teaching and learning as inquiry needed to reach the classroom to make a difference. To introduce teachers to her message, Rowe (1973) wrote an elementary science methods text: Teaching Science as Continuous Inquiry. The text, first published in 1973, included three parts: the scientific framework, the pedagogical framework, and diagnosis and decision-making skills (evaluation). In the revised 1978 edition, Rowe outlined four sections: basic concepts and processes in inquiry science programs, experiments and investigations to do with children, strategies for teaching science as inquiry, and strategies for evaluating and maintaining inquiry science programs. Across editions and sections, Rowe (1978b) emphasized the importance of understanding and experiencing science as story: In this book, science is viewed as a kind of journey into the unknown, with all the uncertainties that new ventures entail. Doing science means using intuition; it means creating abstract ideas out of concrete instances, in order to find out:   1. How things work (description)   2. Why they probably work that way (explanation)   3. What must be done to make them happen in other situations (control) The reader accustomed to standard textbook formats needs to be alerted that much of this book reads more like nonfiction literature than a text. (p. x)

Also in her role as a teacher educator, Rowe worked with the National Science Teachers Association (NSTA) to translate relevant research findings into useful information for teachers. She helped initiate an edited series titled What Research Says to the Science Teacher and edited three of the six volumes (Rowe, 1978c, 1979, 1990) herself. She saw these volumes as serving a dual purpose: to disseminate important research findings to teachers and to encourage teachers themselves to conduct research in areas in need of additional investigation. [The volumes] of this series are part of an effort . . . to create a common frame of reference and language for teachers and research practitioners. They make tangible our belief that research has [emphasis in original] a valuable contribution to make to science education. (Rowe, 1979, pp. i–ii)

110    J. A. BIANCHINI and N. I. HOLTHULS

Rowe supplemented the publication of these volumes with the writing of articles for NSTA’s journals Science and Children, The Science Teacher, and the Journal of College Science Teaching. She tried to keep elementary through university teachers up-to-date on current reform initiatives and research findings in science education. Further, Rowe and colleagues created the Science Helper Video Series (Rowe, Fountain, Babich, & Beverly, 1992) to teach elementary teachers both central concepts in physical science and reform-based instructional strategies. Each of the eight videos in this series closes with a teaching excerpt of Rowe herself. Rowe is shown demonstrating a simple natural phenomenon while talking to students—eliciting their predictions, explanations, and misconceptions (although she does not label them as such). Teacher viewers, however, were probably not aware that they were being taught. Rowe never used academic jargon—no labels for labels’ sake, no complex flow charts, no regression analyses. She simply sat on the floor with a small group of children artfully modeling the process of inquiry and weaving in applicable findings from science education research. This video series is still available from the Learning Team (see again http://learningteam.org/index.html). Edward D. Britton, a former doctoral student of Rowe’s and a current associate director of Science, Technology, Engineering, and Mathematics at WestEd, summed up Rowe’s commitment to translating research findings into information teachers could and would use: She kept moving back and forth across the line between primary research around academic questions to build a body of knowledge and applied research on how to make a difference with teachers. . . . What motivated her to keep crossing that line? She just couldn’t help wanting to make a difference. She wanted people to use things and do things and change things. (personal communication, September 6, 2007)

The Evolution of Rowe’s Ideas The Relation of Wait-time and Rewards to the Development of Language, Logic, and Fate Control While most teachers and researchers are well-versed in the basics of Rowe’s wait-time construct described above, fewer are aware of the larger implications of her wait-time research for classroom instruction. Rowe (1974c), for example, recommended teachers not simply extend waittime when asking and answering questions, but reconceptualize the entire structure of classroom discourse as well. She described the classroom as a two-player system: The teacher was one player, and the students, the other.

Mary Budd Rowe    111

Theoretically, each player had access to four types of verbal moves: structuring, soliciting, responding, and reacting. In most classroom conversations, including those held in most inquiry classrooms, teachers routinely used all four types of verbal moves, while students used only one (responding). Rowe suggested teachers adopt a different model of classroom interaction, one of joint investigation and reasonable conversation, where both teachers and students employed all four moves available. Conversations where students suggested experiments (structuring) and responded to each other’s statements (responding and reacting), Rowe elaborated, better mirrored the purposes and practices of scientists conducting inquiry. Rowe’s investigations of wait-time also led her to examine differential opportunities for learning provided to high versus low achieving students (see Rowe, 1969b, 1974c, 1986). Rowe found that a few select students typically dominated classroom conversations; with increased wait-time, more students in a given class, particularly more students from underserved groups, participated in such discussions. Equally important, use of extended waittime appeared to change teachers’ expectations of what some students could do: Teachers expressed surprise at the important and salient insights shared by students who they had previously dismissed as low achieving. In one study, for example, Rowe (1974c) asked 26 elementary teachers to identify their five highest and five lowest achieving students. Before instruction in wait-time, these teachers routinely waited two seconds for their top five students to respond to a question, while they waited less than one second for their bottom five. When wait-time was extended to three seconds for all students, participation of marginalized students increased. Differences in wait-time given to those students perceived as high or low achieving, in turn, led Rowe (1969b, 1974c) to examine the reward structure of classroom conversations. She found most teachers frequently and regularly praised and sanctioned their students. She also found that those students labeled low achieving received less relevant praise and more criticism from their teachers than those students considered high achieving. As with much of her work, Rowe used a constellation of theories to support her argument that frequent verbal rewards constrained students’ innovative problem solving and shortened their persistence on inquiry tasks. Rowe encouraged teachers to revise their reward structure—their pattern of praise and sanctions—and adopt a neutral stance toward student responses. In short, Rowe asked teachers to construct safe and supportive learning communities where students could test ideas, propose explanations, and argue evidence, again, toward the goal of implementing authentic inquiry instruction in science classrooms. Further, Rowe translated lessons learned from research on wait-time into suggestions for those lecturing in high school and college classrooms. In her 1983 article “Getting Chemistry Off the Killer Course List,” she argued

112    J. A. BIANCHINI and N. I. HOLTHULS

that students rarely capture more than 30% of information provided in a lecture because of lapses in memory. To narrow this gap between information provided by teachers and noted by students, Rowe recommended that teachers pause for two minutes during every eight to 12 minutes of lecture so that triads of students could share notes and clarify questions. Present in this short article was Rowe’s common critique of traditional science instruction: too many concepts and symbols covered in too many pages of text. Present too was Rowe’s longtime interest with technology in education: She suggested computers could help improve student understanding of chemistry concepts by providing specialized applications and problems to supplement lectures and texts. Finally, Rowe’s work on wait-time led to research into student agency, or fate control (Rowe, 1974e). Fate control is the sense that what one does matters. Rowe’s interest in this construct emerged from investigating the interaction between teachers’ verbal rewards and students’ ability to persist in completing inquiry tasks. Rowe understood students’ confidence to connect to how they view life—as a “crap shoot” with no influence on outcomes or as a “bowling match” where outcomes could be influenced. If students understand the world as a craps game (if they have an external locus-of-control), she explained, their problem-solving processes are different than if they view the world as a bowling game (if they have an internal locus-of-control). The craps view stresses luck, chance, and powerful others; the bowler, effort, practice, and persistence. Views of the world as a game of chance, Rowe continued, are incompatible with the norms and methods of science: Science requires its practitioners to be attuned to cause-and-effect, to understand that phenomena emerge from processes they can discover, and to take thoughtful and reasoned action. As with wait-time, Rowe’s purpose in investigating fate control was to better understand how to help all students learn science through inquiry (see also Rowe, 1973, 1978b, 1983b). While at Teachers College, Rowe (1974e) began her research into fate control with a pilot study to determine if short wait-times coupled with intense sanctioning would maintain or foster a craps model in students. In her early years at the University of Florida at Gainesville, Rowe received a grant from the National Institute of Mental Health (NIMH) to study fate control on a much larger scale (J. N. and P. R. Swift, personal communication, September 24, 2007). Reminiscent of her time in Colorado, Rowe constructed a research trailer with a mini-classroom at each end and then traveled to 105 classrooms at 29 schools located in 12 counties in Florida (Main & Rowe, 1993). The sample was representative of rural, urban, and suburban areas and of diverse socioeconomic levels. Three hundred student pairs were investigated—matched by fate control orientation, sex, and ethnicity. These pairs of students completed three inquiry tasks: the assembly of a crystal radio, the rolling of cylinders down a ramp, and the spinning

Mary Budd Rowe    113

of a table. The latter two tasks were adapted from NSF-funded elementary science programs. We think it important to note that even pioneers in science education can struggle in their research. One of her collaborators on this fate control research project, J. Nathan Swift, a now retired professor from State University of New York, Oswego, noted Rowe’s findings “never materialized completely although she had very intriguing data. . . . The data didn’t have any clear sensibility to it” (personal communication, September 24, 2007), he elaborated. Students were pre-tested using two different fate orientation instruments, but the two instruments did not correlate in any sensible way. What was clear, he continued, was that students found the inquiry tasks engaging and intriguing: “When the kids got into the research trailer, doing those simple experiments. What they were doing in the research trailer was so much more interesting than what they were doing in their classrooms.” Rowe eventually published a subset of findings from this extensive research project with a former doctoral student, June Dewey Main (Main & Rowe, 1993). The Centrality of Questions in Science and Science Teaching: So What? How Come? At the intersection of Rowe’s passion for science as inquiry, her research on wait-time and fate control, and her commitment to working with teachers was her focus on powerful questions. One can see Rowe’s translation of fate control ideas into her insistence that science teachers and students routinely ask the question, “So what?” Indeed, Rowe developed a “So What?” chart for science teachers, curriculum developers, and policymakers to use in designing and implementing instruction (1978b, 1983b). Teachers and students were to move through four components: • Ways of Knowing (What do I know? Why do I believe it? What is the evidence?); • Actions/Applications (What do I infer? What must I do with what I know? What are the options? Do I know how to take action? Do I know when to take action?); • Consequences (Do I know what would happen?); and • Values/Who Cares? (Do I care? Do I value the outcome? Who cares?). Rowe argued movement through these four components would ensure that the teaching and learning of science was made interesting and relevant. Rowe also consistently underscored the need to ask “How come?” rather than “Why?” questions when teaching science. A Stanford colleague, Ra-

114    J. A. BIANCHINI and N. I. HOLTHULS

chel A. Lotan (personal communication, January 8, 2012), recalled one of her first meetings with Rowe in which Rowe told Lotan to re-think the “Why?” question in her small group curriculum materials, to ask instead, “How come?” Rowe explained to Lotan that asking why puts students, and people in general, on the defensive; they feel as though they have to defend their thinking, beliefs, values, and so on, rather than explain them. As we see with all her messages, Rowe practiced what she preached. In one episode from the Science Helper Video Series (Rowe et al., 1992), for example, Rowe held a marble and a long clear hose that circled around itself like a snake, the end of which lay on a large piece of paper. Rowe asked two students to predict the path a marble would take upon exit if she dropped it through the hose. The students discussed their ideas and made predictions by drawing the expected path on a piece of paper. Before dropping the marble, Rowe asked them, “How come . . . .?” She then waited and listened to the students’ explanations. Keeping Rowe’s Storyline Alive: Getting Science Off the Killer Course List Rowe led a busy—one might even say adventurous—professional life. She taught science in California and Colorado, as well as in Germany. She served as an assistant professor at Teachers College before making her way to the University of Florida at Gainesville in the early 1970s. She was a division director for Research in Science Education at NSF from 1978 to 1980. In 1983, she became an advisory board member to NSF’s Research in Science Education program. She received NSTA’s most prestigious honor, the Robert H. Carleton Teacher/Scholar Award, in 1981 and was later elected president of NSTA in 1988. She was a member of the National Academy of Education and a fellow for the American Association for the Advancement of Science. The New York Times recognized her as one of a handful of “local leaders of educational innovation”—an innovator who has received little public recognition but has helped make schools better nonetheless (Celis & Nieves, 1991). Rowe was a visiting professor at Stanford’s School of Education in the years before her death in 1996. Rowe was said to be more interested in generating new ideas and collecting data than in publishing findings (P. R. Swift, personal communication, September 24, 2007). Still, the articles Rowe did write made an indelible mark on the field. Rowe’s research paper on wait-time published in the Journal of Research in Science Teaching (JRST) in 1974, for example, was awarded Best Article by the National Association for Research in Science Teaching the following year. This article was later reprinted in the 2003 supplemental edition of JRST as one of the 13 most influential articles in its 40-year

Mary Budd Rowe    115

history (see Rowe, 1974c, 2003). A second article on wait-time published in NSTA’s Science and Children in 1969 was reprinted after Rowe’s death in 1996 to honor her many and significant contributions to science education (see Rowe, 1969b, 1996b). Finally, Rowe’s article on the importance of teacher caring first published in The Science Teacher in 1977 was reprinted in 2000 as part of that journal’s millennium issue; the editor had examined articles written over the journal’s 66 years to select those most significant to the 21st century (see Rowe 1977b, 2000). The fact that her published work fundamentally and unalterably changed the field of science education underscores the point that quality of thought can be more crucial than number of publications. Rowe left lasting imprints on members of the science education community not only through her publications, but through her personal interactions as well—whether chance encounters, attendance as a guest lecturer, or sustained involvement in projects. In brief, Rowe listened to her own advice to teachers about the importance of showing one cared (see, again, Rowe, 1977b). Lawlor underscored Mary’s ability to inspire others: “Rowe was fiercely task-oriented but even more fiercely person-centered. She had a marvelous way of making personal contact and of imparting her vision of what could be done.” Mary exerted influence, Lawlor continued, without ever losing her perspective on self or her sense of humor. I saw just what impact Mary had [on science teachers] at a NSTA conference in Kansas City. Mary and I were walking down a hallway in the conference center when a group of teachers approached. . . . One of the teachers, loaded down with sacks of promotional materials, came up to Mary, dropped his precious cargo, and shook Mary’s hand exclaiming, “I can’t believe that I am meeting a living legend!” Naturally, we all addressed her as “living legend” for the rest of the weekend. (personal communication, September 21, 2007)

While academia can be rife with over-theorized and under-utilized ideas, across publications and personal interactions Rowe held tightly to simple and memorable messages. “Just wait,” she often told us. “Ask a question, and then wait for the answer, and then wait some more.” Even the titles of Rowe’s publications convey the importance of simple messages. These titles include “Wait Time: Slowing Down May be a Way of Speeding Up!” (Rowe, 1986); “Getting Chemistry Off the Killer Course List” (Rowe, 1983a); and “The Uncommon Common Sense of Science” (Rowe & Holland, 1990).1 Rowe thought that these important messages should be delivered modestly, directly, and often. And so that is what she did—she repeated these ideas across teachers, graduate students, academics, scientists, and policymakers from decade to decade using diverse means of communication. What are Rowe’s memorable messages we think important to keep alive today? Clearly, Rowe’s recommendations for ways to structure and facilitate

116    J. A. BIANCHINI and N. I. HOLTHULS

students’ investigations of and conversations about science phenomena remain relevant. In the recent publication Ready, Set, Science, for example, Michaels, Shouse, and Schweingruber (2007) wrote: The kind of discourse that encourages scientific talk and argument is different—in subtle and not subtle ways—from the I-R-E pattern of discourse. To begin with, teachers ask questions that do not have “right” or “wrong” answers or to which they themselves don’t know the answers. For example, a teacher might ask, “What outcome do you predict?” and follow up the initial question with comments such as, “Say more about that.” (p. 90)

The authors continued: They also may follow questions with “thinking” or “wait” time, so that students have a chance to develop more complex ideas and so that a greater number of students have a chance to contribute, not just those who raise their hands first. (p. 90)

These notions of science teaching as engaging students in conversations rather than in rapid-fire questioning, of the importance of evidence and argumentation, of deemphasizing right and wrong answers, of asking students questions and waiting for their responses, of attending to all students, and of the need to grow students’ complex ideas are present in Rowe’s science methods text for elementary teachers written in 1973. We wonder why it has been so difficult to make significant changes in the way science is taught and learned. If read closely, Rowe’s work can be understood to align with current science education research into the intersection of equity, language, and context. Science, Rowe began, should be understood as its own culture with a language and set of practices useful in constructing powerful stories of how the world works. Science teachers serve the “dual role of travel agents for science (i.e., they attract customers and encourage investment in a journey) and guides (i.e., they speak the language of the science culture and are practiced in the ways of knowing in both cultures)” (Rowe & Holland, 1990, p. 88). Although learning a new language is part of learning science, Rowe cautioned, science teaching and science textbooks should not distort and constrain the image of science so that students simply master a foreign language (Rowe, 1983a, 1983b). Rather, Rowe (1969a) emphasized, teachers should begin with students’ everyday language and introduce scientific terms in appropriate contexts: “If teachers let young children start by using their available language, the children learn alternatives by using new [science] words in appropriate contexts rather than by definition” (p. 31). Working together, Rowe concluded, teachers and students can construct scientifically-appropriate stories about how the world works: “Conversing

Mary Budd Rowe    117

freely within the scientific culture and absorption of the common-sense knowledge that marks it becomes a reasonable goal just as it is in any venture to foreign places” (Rowe & Holland, 1990, p. 96). Further, Rowe called for teachers and researchers to transform science education to meet the needs and interests of all students. Rowe (1969a, 1973, 1974b) argued that engaging students in the language and practices of science was a concrete way to help underserved students learn to exert what we today call agency in their lives—to see that, rather than having no control, they could actually transform their lived reality. Rowe (1969a) described this agency in terms of fate control. Science and prediction . . . go together. The more I know about a system, the more I am likely to be able to act on it in definite ways and expect certain results. Prediction rests on the belief that events are not totally capricious, that what I do to the system makes a difference in how the parts act. I can, in some way, act to control the fate of the system. Probably the building of this belief represents the single greatest contribution science can make to the education of the inner-city “disadvantaged” child. (p. 32)

Rowe (1977a) also expressed concern over the low numbers of African American students and students from other ethnic minority groups who enrolled in science courses and elected to pursue science careers. Returning to her research on wait-time and reward schedules, Rowe encouraged science teachers to view all their students as academically competent and to set high expectations for all. A Final Story We close our discussion of Mary Budd Rowe’s contributions to science education with another story Rowe herself often told. Several themes discussed above—the idea of science as story-making, the need for students to engage with natural phenomena, and the importance of substantive and sustained conversations between teachers and students—are revisited here. On a flight to Europe several years ago, I sat next to a sixth-grade boy who watched me use my calculator [in other versions, it is a slide rule] to analyze some data. “When you’re finished, where are you going to look up the answers?” he asked. “No book has an answer to the problems I’m working on,” I said. “It’s up to me to find the answers.” “Then will your teacher tell you you’re right?” “No,” I replied. “I’ll show my results to other people, and I’ll explain my answers, and we’ll talk it over.” “And then will your teacher tell you if you are right?” he persisted. “No, I’m afraid not.” He sighed sympathetically, “Some teachers are like that, you know.”

118    J. A. BIANCHINI and N. I. HOLTHULS For this boy, the world was full of right and wrong answers. He didn’t realize that science is not just facts, but the meaning that people give to them—by weaving information into a story about how nature probably operates. The best way to respond to a child’s question is to begin that process of storymaking together. (Rowe, 1995, pp. 178, 181)

Note 1. To give readers an additional sense of Rowe’s views of science as exploration and playfulness, we integrated some of Rowe’s memorable article titles and phrases into the title and section headings of our chapter. Several titles used come from articles not included in our reference list.

References Bianchini, J. A. (2008). Mary Budd Rowe: A storyteller of science. Cultural Studies of Science Education, 3(4), 799–810. doi:10.1007/s11422-008-9132-y Celis, W., & Nieves, W. (1991, November 3). The flowering of innovation. The New York Times. Retrieved from http://www.nytimes.com/1991/11/03/education/ special-report-the-flowering-of-innovation.html?pagewanted=all&src=pm Hurd, P. D., & Rowe, M. B. (1964). Science in the secondary school. Review of Educational Research, 34(4), 286–297. Retrieved from http://www.jstor.org/stable/1169405 Hurd, P. D., & Rowe, M. B. (1966). A study of small group dynamics and productivity in the BSCS laboratory block program. Journal of Research in Science Teaching, 4, 67–73. doi:10.1002/tea.3660040203 Main, J. D., & Rowe, M. B. (1993). The relation of locus-of-control orientation and task structure to problem-solving performance of sixth-grade student pairs. Journal of Research in Science Teaching, 30(4), 401–426. doi:10.1002/ tea.3660300407 Michaels, S., Shouse, A. W., & Schweingruber, H. A. (2007). Ready, set, science: Putting research to work in K–8 science classrooms. Washington, DC: The National Academies Press. Patrick, C. (1992, February 18). Science “a special kind of story-making” to educator Rowe. Stanford University Campus Report. Retrieved from http://news. stanford.edu/pr/92/920218Arc2408.html Rowe, M. B. (1964). The influence of context-learning on solution of task-oriented science problems which share concepts: A study in elementary science education. Unpublished doctoral dissertation, Stanford University, Stanford, CA. Rowe, M. B. (1965). Influence of context-learning on solution of task-oriented science problems. Journal of Research in Science Teaching, 3(1), 12–18. doi:10.1002/ tea.3660030104 Rowe, M. B. (1969a). Science and soul. The Urban Review, 4(2), 31–33.

Mary Budd Rowe    119 Rowe, M. B. (1969b). Science, silence, and sanctions. Science and Children, 6(6), 11–13. Rowe, M. B. (1971). The fate of ten scientist-science educator teams three years after participation in a leadership training program. New York, NY: Teachers College, Columbia. Rowe, M. B. (1973). Teaching science as continuous inquiry. New York, NY: McGrawHill. Rowe, M. B. (1974a). Pausing phenomena: Influence on the quality of instruction. Journal of Psycholinguistic Research, 3(3), 203–224. doi:10.1007/BF01069238 Rowe, M. B. (1974b). Science that kids can live by. Learning: The Magazine for Creating Teaching, 2(9), 16–21. Rowe, M. B. (1974c). Wait-time and rewards as instructional variables, their influence on language, logic, and fate control: Part one—wait-time. Journal of Research in Science Teaching, 11(2), 81–94. doi:10.1002/tea.3660110202 Rowe, M. B. (1974d). Reflections on wait-time: Some methodological questions. Journal of Research in Science Teaching, 11(3), 263–279. doi:10.1002/ tea.3660110309 Rowe, M. B. (1974e). Relation of wait-time and rewards to the development of language, logic, and fate control: Part two—rewards. Journal of Research in Science Teaching, 11(4), 291–308. doi:10.1002/tea.3660110403 Rowe, M. B. (1977a). The forum: Why don’t blacks pick science? The Science Teacher, 44(2), 34–35. Rowe, M. B. (1977b). Teachers who care. The Science Teacher, 44(5), 37. Rowe, M. B. (1978a). Introduction: Research can help you. In M. B. Rowe (Ed.), What research says to the science teacher (Vol. 1, pp. i-ii). Washington, DC: National Science Teachers Association. Rowe, M. B. (1978b). Teaching science as continuous inquiry: A basic (2nd ed.). New York, NY: McGraw-Hill. Rowe, M. B. (Ed.). (1978c). What research says to the science teacher (Vol. 1). Washington, DC: National Science Teachers Association. Rowe, M. B. (Ed.). (1979). What research says to the science teacher (Vol. 2). Washington, DC: National Science Teachers Association. Rowe, M. B. (1983a). Getting chemistry off the killer course list. Journal of Chemical Education, 60(11), 954–956. Rowe, M. B. (1983b). Science education: A framework for decision-makers. Daedalus, 112(2), 123–142. Retrieved from http://www.jstor.org/stable/20024856 Rowe, M. B. (1986). Wait time: Slowing down may be a way of speeding up! Journal of Teacher Education, 37(1), 43–50. doi:10.1177/002248718603700110 Rowe, M. B. (Producer). (1987). Science helper K–8 [CD]. Sunnyvale, CA: PC Software Interest Group. Rowe, M. B. (Ed.). (1990). What research says to the science teacher: The process of knowing (Vol. 6). Washington, D.C.: National Science Teachers Association. Rowe, M. B. (1995, May). Teach your child to wonder. Reader’s Digest, 177–187, 181, 184. Rowe, M. B. (Producer). (1996a). Culture and Technology [CD]. Armonk, NY: The Learning Team.

120    J. A. BIANCHINI and N. I. HOLTHULS Rowe, M. B. (1996b). Science, silence, and sanctions [Reprint]. Science and Children, 34(1), 34–37. Rowe, M. B. (2000). Teachers who care [Reprint]. The Science Teacher, 67(1), 30–31. Rowe, M. B. (2003). Wait-time and rewards as instructional variables, their influence on language, logic, and fate control: Part one—wait-time [Reprint]. Journal of Research in Science Teaching, 40(Supplement), S19–S32. doi:10.1002/tea.10090 Rowe, M. B. (Principal Investigator), Fountain, K., Babich, M., & Beverly, S. (1992). Science helper series [Video]. Gainesville, FL: University of Florida. Rowe, M. B., & Holland, C. (1990). The uncommon common sense of science. In M. B. Rowe (Ed.), What research says to the science teacher: The process of knowing (Vol. 6, pp. 87–96). Washington, DC: National Science Teachers Association. Rowe, M. B., & McLeod, R. (1988). Science and math instruction: A new partnership. Media and Methods, 25(2), 13–16, 72–73. Science education innovator Mary Budd Rowe dies at 71. (1996). Stanford University News Release. Retrieved from http://news-service.stanford.edu/ pr/96/960625roweobit.html

Mary Budd Rowe    121

This page intentionally left blank.

Chapter 7

A Career of Opportunities Rodger W. Bybee BSCS Emeritus

In life, one has to deal with the realities and consequences of both fortunate circumstances and adverse situations. Both prosperity and adversity come in various forms, and the consequences vary depending on their form, frequency, and duration. The consequences of fortunate or unfortunate situations in early years may be positive or negative respectively. Although this observation may generally be true, I have a different story. My story begins with the revelation that both of my parents were alcoholics and their alcoholism significantly influenced my formative years. In later years, I exhibited all of the classic characteristics of an adult-child of alcoholics. Paradoxically, some characteristics of an adult-child of alcoholics positively influenced my motivations, thinking, and decisions while pursuing a career in science education. In my career, I have been fortunate to have some very significant opportunities: participating in athletics that offset a difficult family situation, attending a community college that opened doors to a world of ideas and leadership, studying ecology that set a foundation of science concepts, teaching science in a laboratory school that provided a diversity of insights about teaching and learning, writing and defending my dissertation that introduced new dimensions of science education, working on national stanGoing Back for Our Future, pages 123–154 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

123

124   R. W. BYBEE

dards that expanded my knowledge of policies, and guiding development of an international assessment that broadened and deepened my understanding of science education. These are a few of the possibilities that became realities. Each opportunity had some anxiety that I had to overcome, and each led to my own professional development. The story in this chapter is not one about the scars and problems of an early life with alcoholic parents. While this was a reality, the story here is positive and centers on my perceptions of and responses to the opportunities I had and continue to have. I begin with a brief review of the perceptions I held—many of which I still hold—as a result of growing up with alcoholic parents. The story continues with discussion of my thinking, influential individuals, and development as a science educator. Family Matters: Insights from an Adult-child of Alcoholics My parents’ drinking became a problem when I was about 10 years old. I can recall commitments of both parents to state hospitals, my mother’s almost successful suicide, time in county jails, automobile accidents, lost jobs, marital problems, and bankruptcies, all associated with their drinking. These were things relatives, law enforcement, employers, clergy, and others knew about. There also were the experiences that went on inside the houses and apartments where we lived that others did not know about. As my parents’ problems continued during my pre-college years, I spent time with grandparents and relatives and finally moved out of my parents’ apartment late in my senior year of high school. Their various problems associated with drinking continued until my sophomore year in college, when they joined Alcoholics Anonymous. They remained sober for many years. Both, however, fell off the wagon late in their lives. In my late 40s, I experienced emotional problems and sought professional help only to discover that I was an “adult-child of alcoholics.” Until then, I had lived and worked with that history but never had to acknowledge the influence, if any, experiences in my early years had on my life and work in later years. Adult children of alcoholics demonstrate a variety of characteristics, and because both of my parents were alcoholic, I have a full constellation of the characteristics. As already mentioned, this is not a story about adult children. That said, let me give several examples of the characteristics, as they set the stage for later discussions of the opportunities and development in my professional work and life. If one’s early experiences are with chaotic, uncontrollable people and environments, then it is natural and helpful to be hyper vigilant, well

A Career of Opportunities    125

planned, and in control. The motto might be: don’t get caught by surprise, plan ahead, and be ready. If your experiences indicate that you cannot trust others to fulfill promises or meet basic obligations, and others continually indicate “nothing is wrong” in clearly troubling situations, then it is reasonable to assume personal responsibility for anything one wants completed. The motto could be: don’t depend on others. If you want it done, do it yourself. If your feelings toward those who are supposed to be responsible for you were not acknowledged, then avoiding or reducing emotions toward others is a reasonable response in later life. What is the motto here? Don’t get invested in another person or project; just assume responsibility for yourself. If one’s personal needs were only marginally met by those responsible for you, then taking care of yourself and getting things done is an understandable response. Don’t expect help from others; do it yourself may be the motto. If one’s experiences at home are dysfunctional, then not sharing those experiences and compartmentalizing aspects of life are natural consequences. What happened at home stays at home; keeping secrets about different parts of one’s life is okay. This is a motto to live by. Finally, low self esteem and not accepting credit for accomplishments is a common response. The motto is: I don’t really believe in compliments from others, so I’ll just deflect the acknowledgement, seem humble, and continue. Back to the observation that these problems can be paradoxical: On the one hand, they are indeed problems for many adult-children. On the other hand, they may be realized as positive attributes in a professional life if one is well planned, completes commitments, remains committed, ignores one’s own needs, and is humble about accomplishments. As I proceed through discussions of my career opportunities and actualities, the characteristics just described should remain in the reader’s understanding of why and how I developed the courage to overcome the anxieties that accompany professional possibilities. My Early Science Education: Not Much to Report I was born in February 1942, a few months after the attack on Pearl Harbor. My parents lived in San Francisco, where I began kindergarten at Longfellow Elementary School in 1947. In 1950, my parents enrolled me in Mission Delores, a Catholic School in San Francisco, where my education continued through sixth grade. I can recall nothing of a science education in elementary school. Mostly, my formal education consisted of preparing for atom bomb attacks, memorizing answers to catechetical questions about

126   R. W. BYBEE

God and Catholicism, hearing about the terrible Communists, memorizing multiplication tables, and diagramming sentences. Most of these experiences had little or no meaning for me. Informally, my science education consisted of the usual experiences of childhood: observing nature, collecting bugs, becoming aware of changing seasons, and occasionally going to the zoo and aquarium. But there were no especially meaningful experiences with science. My parents moved to Denver in 1955, where I went to Morey Junior High School. My clearest memory of science education was a science project on the growth of pea seeds. I had to write a letter to a professor at Denver University who kindly responded to my questions. (I do not remember the letter, questions, or the professor’s answers.) Science education at Morey Junior High School did not capture my interest or leave any memories that I could truly count as influential. In 1956, my parents returned to the town where they grew up—Kimball, Nebraska. I attended Kimball County High School (KCHS) in the small town with a population of 5,000. Kimball is located in the western panhandle of Nebraska. My class had 60 students, and academically I was an average student, graduating in the middle of the class. While in high school, I participated in all athletics, which in the 1950s at KCHS only included football, basketball, and track. I participated all four years, lettering three years. As mentioned, my academic achievements were average, my study skills poor. My primary identity was through athletics, and my academic motivation was to pass courses so I remained eligible to play. The coaches also set curfews for how late we could stay out and rules against smoking and drinking, thus setting rules that influenced my behavior. My motivation to participate in sports, the need to remain eligible, and the coaches’ rules acted as countervailing forces to a home environment that lacked goals and rules. To this day, I acknowledge the important role of athletics in my career. Any interest in and understanding of college did not really emerge until my senior year. What about my science education in high school? I remember general science class, but not any particular content. My high school biology course was adequate and interesting. I can recall only one laboratory activity, the dissection of an earthworm. I also was intrigued with the various systems of a frog. The systems were revealed by overlain transparencies in our book, Modern Biology. Perhaps we had other laboratories; if so, the experiences have long since left my memory. I took physics, but it, too, left little impression beyond interesting demonstrations and memorization of terms such as matter and energy. My science education in high school may have fomented some interest, but it certainly did not provide a conceptual foundation for further study. To be truthful, what did sustain me was an interest

A Career of Opportunities    127

in athletics and my friends. That is, science took a back seat to sports and social concerns. However, high school in Kimball, Nebraska did include what I now recognize as a significant experience in my career. In my junior and senior years of high school, I took courses in industrial arts—otherwise known as “shop.” The concrete “activity-based, project-centered” approach appealed to me. Learning to weld and work with wood, combined with the visible results of the course, held my interest. In fact, during my senior year, the teacher had to leave for two professional meetings. On those occasions, I was designated “student teacher” with responsibilities for taking roll and seeing that the class began and ended on time and without mishap. The principal occasionally looked in to see that all was going well. The experience of “being the teacher” made a deep and lasting impression on me. I simultaneously was very excited and quite anxious about the opportunity. Things turned out well—no fingers lost in saws and no discipline problems. The other students accepted me, and I did not overplay the role of teacher. I liked the role and responsibility. A Turning Point: An Opportunity to Attend Junior College During my senior year in high school, a local family, Paul and Luella Mockett, asked to meet with me. After a discussion about my goals, they offered me the chance to attend college if I would agree to go to Otero Junior College in La Junta, Colorado. One son and a daughter of theirs attended Otero, and they thought it would be a great opportunity for me. The Mocketts made three stipulations on their support for my two years of junior college. First, I had to make passing grades. Second, I had to prepare a proposed budget for each term and then report on actual expenses at the end of the term. Finally, I had to abide the college’s policies. After applying to Otero, I visited the college and, while there, tried out for the basketball team. As a result, I received an athletic grant that waived my tuition, books, and fees. The Mocketts agreed to pay the equivalent of the grant if I would place the money in a savings account for my college education after Otero. I agreed. This opportunity was both unexpected and a significant turning point in my life. I may have thought people in Kimball, Nebraska did not know about my parents, but they did. Later, I discovered that Paul Mockett asked my uncle what he thought about giving me the opportunity to attend college. My uncle told him I either would do exceptionally well or fail miserably. I can only surmise that the Mocketts understood my plight and realized my potential.

128   R. W. BYBEE

For two years, I met all the stipulations, and the Mocketts supported my education. I have been thankful for their generosity, confidence, and stipulations all of my adult life. This also provided me with an introduction to preparing and reporting on proposed and actual budgets, experiences that grew in importance as I developed as a professional. When reasonable and possible, I have tried to provide others with support and the same stipulations. In 1960, I began my college career at Otero Junior College. Although I did not realize it at the time, my career and possible opportunities in science and technology education also began at this community college. During my two years at Otero Junior College, I had experiences that provided a substantial academic foundation, and I formed professional qualities that helped me as I later worked on projects such as the National Science Education Standards (National Research Council, 1996), labored at the National Research Council, helped frame the Program for International Student Assessment (PISA), and directed the Biological Sciences Curriculum Study (BSCS), a nationally recognized science education organization. The transition to college life presented order, security, and community. This life directly contrasted to the environment created by my parents. The transition also presented the realities of my poor preparation for the academic rigors of college. Again, my identity with athletics, in this case basketball, provided a structure, motivation, and goals that complemented those set by the Mocketts. At the conclusion of my first year at Otero, I entered a race for student body vice president and won. I had emerging study skills, established some academic foundations, realized more of my non-athletic potential, and developed some political skills. I remember very clearly confronting the reality that I was only five foot eight inches tall, and though I played very well in high school and junior college, a career in athletics was not an option. I would have to rely on something other than basketball for a profession. During this identity crisis, I realized that I did very well in the sciences. In fact, the emerging field of ecology captured my interest and became my primary field of study within biology. I did, after all, become a person of the 1960s. My first courses at Otero included zoology, English, and history of western civilization. The teachers and coaches were wonderful. I especially remember Harry Robinson, my zoology professor, and later Dexter Hess, my botany and ecology professor. William Boast taught English. And Jim Weiand was my basketball coach. In my science courses, I simultaneously realized the excitement of the natural world and my lack of understanding of basic concepts. In English class, I found the intrigue of ideas and interest in writing clashed with my deficit in grammar. I already mentioned my love of sports and limits they had for a career. Without exception, these teachers challenged me to my

A Career of Opportunities    129

limits. They saw through my faults and weaknesses and recognized my unrealized potential. I understand this now. At the time, it was only hard work and meeting challenges, be they athletic or academic. I began college about the same time environmental concerns and ecology came to the public’s attention because of critical issues such as smogfilled skies and polluted rivers. My initial year of study included an elective that established my interest and understanding that has continued to this day. I enrolled in a plant ecology course that included field study and an independent investigation. At the time, this was just an interesting elective course. Now, I realize it engaged a theme of my career in science education. The textbook for the course was Eugene Odum’s Fundamentals of Ecology (1959). What I remember most is struggling to understand concepts such as autotrophic and heterotrophic components of ecosystems, biogeochemical cycles, energy in biological systems, and limiting factors. I also remember being very interested and highly motivated by the field trips and inquiry into simple, and not very significant, ecological questions. Scientific inquiry also became a theme in my career, but I did not realize it at the time. Becoming a Science Teacher: A Life Decision After two years at Otero Junior College, I left and continued at the University of Colorado, Boulder, where I took other science courses including animal ecology from Gordon Alexander, a renowned scientist. I still struggled with the concepts, as this was an upper division and graduate course. My enthusiasm carried me on, because by this time I had the initial glimmers of pursuing a career in the emerging scientific field of ecology. In the spring of 1963, I applied for an assistantship in a field study of alpine ecology and was not accepted. So, I did not have research experience as a field biologist. In the fall of 1963, I transferred to the University of Northern Colorado (then Colorado State College) in Greeley, Colorado and began a program of study to become a biology teacher. Until that time, now fully three years of undergraduate school, I had not really decided what to do with a biology major. Should I go into research? Some health-related career? Become a biology teacher? Like most, I explored these ideas primarily with friends. One day while thinking about such questions and walking across campus, the answer came in a moment of insight. I would become a biology teacher and perhaps an assistant coach. Coaching would provide another unique context to teach students. I soon dropped the idea of minoring in physical education and pursued a minor, and eventually a major, in fine arts. So, you ask—Why fine arts? At Otero Junior College, I took drawing classes and did quite well. As this interest continued, I realized the connections between science and art—for example, medical illustration.

130   R. W. BYBEE

To this day, I remember the decision-making experience and my exact location on the campus when it occurred. On visiting the university, I often walk by the spot and recall the moment and new-found direction for my life. Although my career has varied from the original goal of being a biology teacher, it has not varied from the general aim of science education. That first year at the University of Northern Colorado, I had to decide on an emphasis for my biology major, and I elected ecology. In spring 1964, I took an advanced course on biotic communities that furthered my understanding and enthusiasm for ecology and the possibilities of teaching. My early introductions to ecology were generally without connections to humans or social issues. For example, the book by Odum that I mentioned earlier had nine pages on “Applications: Human Society.” The book was 496 pages of text, and this was the final section. As I recall the courses and review the textbooks, it is clear that at this time humans really were not considered a part of ecology. In retrospect, I developed an understanding of ecology that later expanded with the discipline’s introduction of humans as significant influences in ecosystems. The initial ideas of human ecology had a history that began much earlier than 1964 (see Bybee, 1984, 2003), but I estimate it was around this time that I first heard about Silent Spring by Rachel Carson (1962). To say I read the book from cover to cover would be an overstatement. I did read sections and had an emotional response that initiated my concerns about society and science education. One of the next significant steps in this story of possibilities and becoming a science teacher was a course on “Teaching Science in Secondary Schools” taught by Robert B. Sund. Students in the class had to purchase the high school science book they likely would use. BSCS Biology: An Ecological Approach was my selection. I was quite excited with this purchase, because it aligned with my interest in ecology, as well as made a major connection to science teaching, a career possibility about which I was very clear and most enthusiastic. When I think about that BSCS book now, I realize that it represented a concrete connection between my studies as an undergraduate and my career in science education. Another important factor in my decision to pursue becoming a science teacher was Bob Sund, my first mentor. Bob took a personal interest in my development as a science teacher and science educator. I was greatly influenced by his varied interests, enthusiasm, and intellectual rigor. When I took his methods course, he was just completing a manuscript for the first edition of Teaching Science by Inquiry in the Secondary School (Sund, 1967). He asked me to join as a co-author for the third edition of the textbook. Unfortunately, Bob died in 1979 before we completed the third edition. The book is now in its tenth edition, published as Teaching Secondary School Science: Strategies for Developing Scientific Literacy.

A Career of Opportunities    131

Being the executive director of BSCS and seeing “The Green Version” biology textbook go to its tenth edition 45 years after my introduction to the text certainly gave me enormous satisfaction. In this same period, roughly 1964 to 1965, I had to take a geology class. This stirred within me new interests in Earth sciences, ones that would complement my ecological understanding and social concerns. The Earth sciences stimulated enough interest that I diverted studies in biology (I had already completed a major) and launched into Earth science, including a year of teaching 9th grade Earth science. The combination of life and Earth sciences provided a sound basis for what later became for me an enduring concern about population, resources, and environmental issues and the need to address them in school science programs. Being a Science Teacher: A Diversity of Experiences After a year-long internship in 1965–1966 at Heath Junior High in Greeley, Colorado public schools, I continued teaching junior high school science at the University of Northern Colorado Laboratory School. Teaching in a laboratory school presented diverse and unique opportunities that established a foundation that influenced my career. During this period, I routinely taught science to ninth grade students and elementary students grades K–6. In addition, I worked with profoundly deaf preschool students, severely developmentally disabled students, and in the summer, disadvantaged high school students in an Upward Bound Program. Also, undergraduate science teacher education students were daily observers in my classes. I will return to my discussion of these opportunities later in this chapter. While teaching at the laboratory school, I introduced and taught several innovative science curricula. The programs were produced in the Sputnik era. I taught the Earth Science Curriculum Project (ESCP), Science Curriculum Improvement Study (SCIS), Elementary School Science (ESS), and Science—A Process Approach (S—APA). In the summer program, I taught Biological Sciences Curriculum Study (BSCS) Green Version. The diversity of science teaching experiences provided a broad and deep foundation for my later work. The experience of teaching science to junior high (now middle school) students is one all future teachers should have. Students who are simultaneously mature enough to be really engaged by science and distracted by issues of personal and social development present unique challenges, to say the least. While at the laboratory school, I also taught elementary students for two years. This provided insights about cognitive development and learning science that could not be gained from a textbook. This was a period when Jean Piaget’s explanations of cogni-

132   R. W. BYBEE

tive development were very popular in science education. After Bob Sund’s death, I completed a revision of a book on Piaget that Bob initially wrote (see Bybee & Sund, 1990). All of these experiences helped me better present ideas in my later professional writings. A rare opportunity I enjoyed was teaching science to profoundly deaf preschool students. I taught these young boys and girls several times a week for a year. I contend that this was one of the most insightful and knowledgebuilding experiences of my career. The fact that these children were deaf was key. Whatever science I wanted to teach simply could not be accomplished by talking. I had to provide conceptually meaningful experiences that established and built fundamental ideas about, for instance, size, texture, color, and shapes. In addition, I had to pay attention to the students’ interests. When they lost interest, they would cease looking at me and/or the materials. Early attempts to regain their attention proved futile. So, I learned to not worry about the “typical lesson,” which is defined as a class period. For the time we had, I prepared several lessons of increasing interest. The lessons might begin with sorting objects of different sizes, shapes, and colors and move to the students’ observations of a live turtle that I kept in the classroom. At the next class session, I would briefly return to sorting objects and applying simple terms, introducing students to other experiences such as “challenges” to describe the turtle as large or small, smooth or rough, fast or slow (Bybee & Hendricks, 1972). During this period, I also taught students with special needs by employing some of the same lessons adapted from the Science Curriculum Improvement Study (SCIS). The experiences provided me with very different insights about students’ learning. While the deaf children were excited about science class, they often would communicate questions about when would they get to go to the science room, while the mentally retarded children exhibited a very low or flat affect. They seemed to be continually puzzled but not necessarily enthusiastic about the science lessons. One of my clearest memories of teaching the retarded children was watching them initially grasp an idea—for example, learn the difference between small and large when describing an organism—and then literally “lose” the concept in the next few minutes. I would ask them to repeat the idea, and they would respond with a puzzled gesture that communicated—what idea? what concept? what organism? what object? These experiences directly contributed to the later formation of the instructional approach known as the BSCS 5E instructional model (Bybee, Taylor, Gardner, & Van Scotter, 2006). In the summers of 1967 and 1968, I taught Upward Bound students science. These students came from diverse cultures and disadvantaged backgrounds. They were high school students of Hispanic, African American, Native American, and Caucasian backgrounds. All had potential to go to college but were not achieving at levels commensurate with their abilities.

A Career of Opportunities    133

My goal in teaching them was to establish some basic concepts of the sciences and to ignite their interest in the natural world. Each day, I tried to have a lesson that engaged their thinking and sustained their interest while developing concepts related to energy, evolution, or Earth systems. Teaching science to these students presented challenges of motivation and developing concepts that many students learn in elementary or middle school. I also did not have time to think in terms of extended lessons, units of study, or full courses. Each day, or at best two or three days, I had to provide ample experiences to sustain their interest and facilitate learning of basic scientific ideas. I mentioned the fact that undergraduate students observed my teaching daily. Before long, I realized that it was easier for them to sit in the back of the room and critique my science teaching than it was to teach science. So I did two things. First, as early as possible in the quarter of observation, I had these undergraduates teach a single concept, usually about a five minute lesson that was, if possible, a demonstration. Later, they would teach a laboratory and full lesson. These experiences helped them understand the complexities of a classroom and the decisions science teachers have to make on a moment to moment basis. Second, I began preparing short activities that would have the undergraduate students “investigate” various aspects of science teaching. They would keep track of the type of questions I asked or the characteristics of perceived discipline problems, or they would compare student attention in different lessons. These lessons later became important activities in the methods textbook, Teaching Secondary School Science: Strategies for Developing Scientific Literacy (Bybee, Powell, & Trowbridge, 2007). Forming New Ideas about Science Education: Graduate School Here I must digress. I graduated with a bachelor’s degree from the University of Northern Colorado in 1966 after three intense years that included full programs in summer school and a year’s student teaching as an intern. The internship was half a day, so I continued taking courses during that same time. In the fall of 1966, I began teaching at the university laboratory school and simultaneously enrolled in a master’s program, also at the University of Northern Colorado. Beginning my graduate education immediately after finishing my bachelor’s degree confronted me with a critical choice and new possibilities— should I pursue a master’s degree in science or in science education? I remember thinking about the higher prestige of science and the lower reputation of education. While in this frame of mind and unable to overcome anxieties attending the choice, I pursued both goals, which included in-

134   R. W. BYBEE

troductory graduate courses in life and the Earth sciences and in scientific research. I also enrolled in several science education courses. At the year’s conclusion and after several long discussions with my internship supervisor, Donald Adams, I decided on science education. One experience stands out as particularly important for me at that time. At the University of Northern Colorado, I took a course on introduction to scientific research from Albert M. Winchester, a geneticist of worldwide reputation. Within the term, we had to identify a question, design an investigation, conduct the investigation, write the equivalent of a scientific paper, and report the results in a mock scientific meeting. The latter was the final exam. This experience left two impressions with me. First, it enhanced my understanding and interest in scientific inquiry as an essential component of science education. Second, the question and investigation I worked on was the chronic low-intensity effects of radiation on the egg, pupa, and larvae stages of the fruit fly, Drosophila melanogaster. My second point centers on the concept of “chronic low-intensity.” When you think about it, most educators maintain a view that learning occurs in “acute high-intensity” situations. Examples of this perception include the reliance on a lesson, activity, or demonstration as the central experience of learning. Some things, however, are better taught and learned in little doses over extended periods of time. For instance, I think that many years of different types of inquiryoriented activities could slowly develop students’ understanding and abilities much better than a single experience, such as a science fair project. Students need time and opportunities to learn. Focused repetition on key concepts and processes is not necessarily bad. Marketing uses this approach in the repetition of 10 to 30 second TV commercials. My graduate work in science education included the standard courses on curriculum, assessment, research, statistics, and other topics of interest to my professors, in particular Robert B. Sund and Leslie W. Trowbridge. Graduate study for my master’s program culminated with a thesis typical of the period. I compared the learning outcomes for a college-level introductory Earth science course taught as lecture-demonstration with a laboratoryoriented course (Bybee, 1969). Students’ attitudes favored the laboratory course, and there was no significant difference in the Earth science content they learned (Bybee, 1970). Near the end of my master’s program, Bob Sund asked me to work with him on a reader for elementary school science. His publisher had asked for such a book to complement an elementary methods textbook he had published with Arthur Carin. Actually, Bob turned most of the work over to me. As it turned out, the general outline, selection of articles, obtaining permissions, writing introductions and summaries, and taking care of the publishing process all became a wonderful introduction to preparing a book—without actually writing the book! The collection of articles was

A Career of Opportunities    135

published in 1973 as Becoming A Better Elementary Science Teacher (Sund & Bybee, 1973). During my graduate study at the University of Northern Colorado, I realized the importance of continuing graduate study at a different university. I briefly entertained the idea of graduate study in psychology based on the experiences of teaching and an interest in the humanistic psychology of Rollo May, Carl Rogers, and Abraham Maslow. I decided to continue work in science education and applied for a Teacher of Teacher Trainers program at New York University (NYU). I began study at NYU in the fall of 1970 under the direction of J. Darrell Barnard. Upon arriving in New York, I found out that F. James Rutherford had accepted a position as chair of science and mathematics education at NYU. In the early 1970s, the science education faculty at NYU included J. Darrell Barnard and Morris Shamos. The center of their work was Conceptually Oriented Program in Elementary Science (COPES). Graduate students were expected to do their dissertation on COPES. Although I appreciated and supported the COPES project, I wanted to study other areas that I thought important for science education—philosophy of science, philosophy of education, and psychology as it applied to learning and development. While continuing to study under Darrell Barnard, I slowly shifted to Jim Rutherford as my thesis advisor. My professional relations with Jim Rutherford have grown and matured as we continued our careers after leaving NYU. Coursework contributed to my continuing growth and interest in fields associated with science education—for example, psychology, philosophy, history of education, interdisciplinary analysis of educational issues, and the usual statistics, computer programming, and seminars. The PhD dissertation I completed clearly brought many of my interests together in a fairly major undertaking. Work on my dissertation set an academic foundation that complemented my experiences teaching science. I combined my interest in humanistic psychology with my experiences in science education and used a book by Jim Robinson, The Nature of Science and Science Teaching (1968), as a model. For my dissertation, I studied the implications of Abraham H. Maslow’s philosophy and psychology for science education in the United States (Bybee, 1975). After a thorough study of Maslow’s work, I established a framework for exploring the implications of his work. I decided to use the goals of science education, as they were a deeper structure that influenced all facets of the discipline. I first identified the enduring goals, which are (1) scientific knowledge, (2) scientific inquiry, (3) personal development, (4) career awareness, and (5) aspirations of achieving a better society. In a study such as mine, one has to first establish how each goal changes over time; second, where the relationship or priority of goals was found (e.g., in

136   R. W. BYBEE

a policy document or a curriculum); and finally, if a goal (e.g., social aspiration) has ascended or descended in importance when compared with the other goals. The point here is that I identified the important goals for science education and realized how those goals change in the course of history and how they influenced the form and function of policies, programs, and practices. The exact implications of Maslow’s ideas have had little or no influence on science education. As an aside, I did prepare an article and submitted it to Science Education, and after an initial review and response, I have neither heard from the editor nor had the article rejected. That was 40 years ago! Although Maslow’s ideas may not have had an impact on science education, the intellectual undertaking has affected my career’s work. I owe a great debt to the advice, counsel, and constructive criticism of my major advisor and now colleague—F. James Rutherford. What lessons did I learn from graduate school, especially the completion and defense of a dissertation? With time and reflection, I suggest the following answers to this question. I learned to propose and defend a project. This lesson was especially helpful in writing proposals, for example, to the National Science Foundation (NSF). I learned the importance of understanding the history of science education as a discipline and its knowledge and application—for example, through curriculum and instruction. I also learned the value of constructive criticism and its contributions to clear thinking and writing. Synthesizing My Ideas: Life as an Academic Near the completion of my doctoral program at NYU, I was offered a position in the Education Department at Carleton College in Northfield, Minnesota. I accepted immediately. When I applied, I really did not realize the prestige of Carleton College. Rather, it was an opportunity to combine my experiences in science education with my interests in psychology, the environment, and other areas both within and outside of education. At this point, I must return to the theme of ecology and my earlier concerns about the environment. My graduate study did not include environmental studies. I did, however, have time to read. In the late 1960s and early 1970s, there were books such as Paul Ehrlich’s Population Bomb (1968) and Garrett Hardin’s classic 1968 article, “The Tragedy of the Commons.” The reality is, however, I did not have the time to formally pursue my interest or concerns during my graduate studies or during the early years as an assistant professor of education at Carleton College. My first year at Carleton, 1971, was spent getting to know the Northfield school system where we placed students as tutors, observers, and student

A Career of Opportunities    137

teachers. I also had to devote considerable time to preparing and teaching education courses. With time and adjustment to this new academic environment, I extended my interests and re-engaged in ecological issues. In this period, I attended a symposium held at Carleton and organized by Ian Barbour. The symposium centered on environmental issues and included a major session by Donella and Dennis Meadows, both Carleton alumni. The Meadows had been lead authors on a text sponsored by a group referred to as The Club of Rome. They published what for me was a powerful book entitled The Limits to Growth (Meadows, Meadows, Randers, & Behrens, 1972). Ecological issues took on new meaning and depth as I continued reading and understanding the complex interactions of population growth, resource use, and environmental quality. My study quickly extended beyond the “pure” ecological concepts to the dynamics of systems, the social issues of economics and politics, and the human dimensions of basic needs, values, and ethics. A small liberal arts college turned out to be the ideal place for me to re-engage my interest in human ecology. First came the insight that ecology and economics had a connection, and that finite resources such as oil could become a central concern of nations, even a reason to go to war. Second came the implications of these ideas for science education. Finally, I began using ecological concepts as metaphors for understanding school systems and the dynamics of educational change and reform. With this newly developed understanding came my deeper realization that education, especially science education, had a responsibility to address these issues. Two ideas seemed fundamental. First, it was essential to incorporate means for students to realize the science associated with many social issues. Second, introducing scientific ideas in their “pure” form did not necessarily lead students to making the connections, much less applications and actions, to the fundamental issues of human ecology— populations, resources, and environments. Although I developed the view that social issues should be addressed in science education, I am more an introvert than extrovert, so inserting myself into the social aspects of advocating for human ecology was not easy for me. I am an INTJ on the Meyers Briggs Type Indicator. So, my approach to science education takes the form of teaching courses, writing articles, speaking at professional meetings, and working on policies and programs. By the late 1970s, I was among the senior faculty at Carleton College and thus had the opportunity to teach one or two courses a year outside the Education Department. I also had shared my interests in environmental issues with Ian Barbour, a colleague in the Religion Department who was noted for his writing on science and religion. Talking to and eventually teaching with Ian challenged me more than I ever imagined. He had long studied, thought, and taught about ethics, technology, the environment, and resources. In the early 1980s, Ian and I taught a course on “The Sus-

138   R. W. BYBEE

tainable Society.” Ian had just published a book on Technology, Environment, and Human Values (Barbour, 1980), and Lester Brown had just published Building a Sustainable Society (Brown, 1981). During this period, I taught one other course of much interest to me: “Science, Technology, and Public Education.” By the late 1970s, my academic work in ecology and educational interests merged in several articles. Historical documents in science education had always proclaimed this goal of social applications of science, albeit in different forms and contexts in different historical periods. I published a detailed discussion of the process of reform in an article titled “The New Transformation of Science Education” (Bybee, 1977a). In this article, I described a series of stages in science education reform and the contemporary social factors that were facilitating what I proposed was a contemporary transformation. Among the factors were national energy issues and global problems. I began work on several articles that would eventually clarify my ideas and set a foundation for my work during the next decades. I did extensive reading and reflection in areas outside of science education for these articles. I read some on Judeo-Christian traditions and ecology, personal and cultural values, scientific and social paradigms, the tragedy of the commons, and limits to growth. My challenge was simply stated but immensely complex: How did society get to this contemporary position relative to population growth, resource use, and environmental quality? What should we do now? And what role should science education play? Work in this period produced several publications (Bybee, 1977b, 1979a). Looking back on this period, I can truly say that my experiences at Carleton College presented new opportunities and avenues of exploration that formed my thinking and values relative to science education, because I brought together rich sets of ideas from diverse sources. In the end, I developed several ideas that continue to be critical to my understanding of social issues and science education. One idea became very important and is attributable to Jonas Salk, with whom I also had an opportunity to discuss the concept. Salk (1973) presented a self-regulative growth model based on growth of many organisms in limited physical systems. The growth, if plotted as population numbers and time, approximates a sigmoid or S curve. In early phases, growth is quite rapid. Then, at the point of inflection, the limits of the system are sensed by the organisms, and population growth slows to sustainable levels. Organisms such as fruit flies respond instinctively. The response for fruit flies is genetically programmed. But, what about humans? I had to enter the realm of ethics and values where works such as Theodosius Dobzhansky’s The Biological Basis of Human Freedom (1956) and Jacob Bronowski’s The Ascent of Man (1973) helped formulate a model of growth and values for an ecological society. I published two articles that presented my synthesis of these ideas. These articles are little recognized and

A Career of Opportunities    139

seldom cited, but for me they are among my most significant intellectual accomplishments. The first article was “Science Education and the Emerging Ecological Society” (Bybee, 1979b). At least for me, if not for the educational community, I made the case for reform. Science education had a new sense of social purpose, one that supported slower rates and different directions of growth. That purpose would influence policy by taking into account four basic factors that I derived from the fundamentals of ecosystems and ideas of Abraham Maslow. 1. Development of the individual places emphasis on the fulfillment of basic human needs and discovering means of nurturing continued healthy personal growth. 2. Development of environmental quality includes the protection, conservation, or improvement of all the factors—for example, air, water, noise, and stress—that affect individual and community development. 3. Development of resources means deciding what natural resources are to be used and the degree to which they are used, recycled, and conserved. 4. Development of community entails greater recognition that there are groups of humans at local, regional, national, and international levels that are dependent on one another for the basic requirements of individual development and that we must cooperate in the elimination of racism, sexism, and war. These fundamentals would establish and maintain a social order, one characterized by the title, “ecological society.” I wrote the second article as a complement to the argument for an ecological society titled “Science Education Policies for an Ecological Society: Aims and Goals” (Bybee, 1979c). By the late 1970s, the nature of many social problems led me to conclude that the burden for responding to the problems would be placed on science education because (1) they were related to science and technology and (2) many social changes could be accomplished through education as a least intrusive form of social changes in comparison to, for instance, taxes or laws. Today, in an era of global climate change, I would make the same case for science education and an even stronger need to reform. During this period, I attended an Aspen Institute Seminar on Aristotle’s Ethics taught by Mortimer Adler. This was a significant intellectual and educational experience in many ways. One result of the seminar was my development of what has become known as the 4Ps. This model uses the terms Purposes, Policies, Programs, and Practices to characterize different domains of an educational system. One can think of the 4Ps as goals, plans, instructional materials, and teaching strategies. As it turns out, many discussions, presentations, reports, and books can be characterized using one of these domains. In my experience, few discussions take what would

140   R. W. BYBEE

be a holistic perspective. Typically, an educational discussion has a single perspective and stays within that domain (e.g., practices). Understanding the totality of these perspectives and individuals’ tendency to stay within one perspective has helped me clarify many discussions and reduced my frustrations when people criticized one perspective from that of another. For example, if a person gives a presentation on policy, another may accuse the presenter of not understanding the “real” world of classroom practice. By the way, the reverse also can be heard. Presentations from a perspective of classroom practice often miss the place, intention, and reality of district, state, or national policies. The 1979 essay on aims and goals (referenced above) elaborates the defining characteristics of Social Purposes, Educational Policies, School Programs, and Classroom Practices. At this time, I wanted to establish the idea of social purposes and educational policies to connect the ideas from the prior essay to this one. I assumed the purposes implied by the idea of an “ecological society” had been clarified, and there was a need to describe the major policies that would inform programs and practices. Here is what I stated in that essay: In the next decade, policies for science education programs and practices should include the appropriate cognitive, affective, psychomotor and social objectives to • Fulfill basic human needs and facilitate personal development; • Maintain and improve the physical and human environment; • Conserve and efficiently use our natural resources; and • Develop greater community at the local, regional, national, and global levels. (Bybee, 1979c, p. 249)

These statements of major policies represent a synthesis of my prior interests, study, teaching, and experience. You can see fundamental ideas from ecology, the psychology of basic needs and self actualization from Abraham Maslow, the philosophical teaching of Mortimer Adler, the educational writings of Joseph Schwab and John Dewey, and the ethical and theological influences of Paul Tillich and Martin Buber. In sections of the essay following this initial statement of policies or general guides for programs and practices, I expanded each of the statements. For each of the policies, I suggested knowledge and values that also should guide the design of curriculum materials and instructional practices. Such an effort was not easy, and I soon had to clarify the criteria for my selection of values, as well as the values themselves. Indicating what knowledge may be of most worth for policies, such as “maintain and improve the physical and human environment” and “conserve and efficiently use our natural resources,” made this more than the usual educational exercise and brought the statements clearly into the realm of social policy.

A Career of Opportunities    141

So you ask, “What values did I propose?” For “fulfill basic human needs and facilitate personal development,” I suggested equity and beneficence (and non-maleficence). For “maintain and improve the physical and natural environment” and “conserve and efficiently use our natural resources,” I proposed stewardship and prudence. And for “development of greater community,” I recommended cooperation and agape (the far-reaching love and compassion for humanity). I completed all of the work described in the preceding paragraphs during the late 1970s and early 1980s. My work in the mid to late 1980s broadened to include scientific literacy and the science-technology-society theme. In the early 1980s, I began using the term “scientific literacy” to express the aim of science education. Scientific literacy conveys the general education aim for citizens as opposed to a specific aim, such as preparation for careers in science and engineering. To my way of thinking, our policies, programs, and practices had to help students understand the many and varied issues associated with science and my continuing interest in human ecology. My interest and understanding had developed due to the experiences at Carleton College. And as I looked at the larger landscape of science education, I found very few examples of policy-related articles that included significant social issues, even though many authors espoused the aim of scientific literacy (Bybee, 1985). For a number of reasons, most of which I would have difficulty identifying, my interests and writing took on a larger perspective in the 1980s. This occurred without eliminating or ignoring environmental concerns, but I thought science education had entered a new era, one that was symbolized by the report, A Nation at Risk (National Commission on Excellence in Education, 1983). The prominence of science education in A Nation at Risk did not escape my attention. This turned out to be a new era of reform, even if it was not as clear and as well defined as the Sputnik era. In the early 1980s, the theme of scientific literacy for all students replaced the dominant theme of the 1960s—science manpower. Science-technology-society emerged as an emphasis for programs and practices in the 1980s. During a sabbatical leave from Carleton College, I edited the 1985 NSTA Yearbook. The yearbook theme and title was Science-Technology-Society. About this time, I read Lawrence Cremin’s book Public Education (1976) and used a paraphrase of his question about educational goals as the title and theme for an essay in the NSTA yearbook: “What should the scientifically and technologically literate person know, value, and do—as a citizen?” More precisely, I stated this as the Sisyphean question in science education, using the Greek myth of Sisyphus to express the need to continuously ask the Sisyphean question and to consider reviewing our purposes in light of ever-changing social circumstances. Several articles that I wrote during this period expressed my interest in environmental issues through the S-T-

142   R. W. BYBEE

S theme. An example is “Science Education and the Science-TechnologySociety (S-T-S) Theme” (Bybee, 1987). The Sisyphean question has become a question that I have used since the mid 1980s. For instance, I used it in developing the assessment framework for PISA Science 2006 and the basis for a proposed environmental literacy framework for PISA 2015. My early advocacy of the S-T-S theme changed when I realized that most attempts to implement the theme were quite superficial. They did not really explore the science and technology associated with various societal issues (not to mention political, economic, and ethical), many of which, as it turned out, were about environmental or resource problems. As societal issues worsened my articles became more direct and forceful (see e.g., Bybee, 1991). My articles also developed the theme of scientific literacy. The influence of John Goodlad’s What Schools Are For (1979), Mortimer Adler’s The Paideia Proposal (1982), Lawrence Cremin’s Public Education (1976), Ernest Boyer’s High School (1983), and science educators and colleagues such as Paul DeHart Hurd and Doug Roberts is clearly evident in my discussion of scientific literacy. My work in scientific literacy eventually resulted in a book published in 1997, Achieving Scientific Literacy: From Purposes to Practices. My time at Carleton College significantly expanded my knowledge and understanding in general and of science education in particular. Teaching, reading, and writing resulted in the synthesis of ideas and themes that continue to this day. In the early 1980s, I really began developing my ideas about scientific literacy and broader, more encompassing views of science education. This expansion was due in part to the change in social climate vis á vis education. A significant aspect of my work at Carleton College centered on teacher preparation in the sciences. I worked on the methods textbook mentioned earlier and continued my involvement in science education by attending and presenting at meetings such as the National Science Teachers Association (NSTA) and the National Association of Biology Teachers (NABT). Only in the environment of a small liberal arts college could I have had opportunities to teach in education, psychology, religion, and science technology and public policy programs. The students and faculty were both challenging and supportive. Moving Beyond the Academic World: Joining the Biological Sciences Curriculum Study I left Carleton College and joined the Biological Sciences Curriculum Study (BSCS) in 1985 and served as associate director from 1985 to 1995. At BSCS, I was principal investigator for four new National Science Foundation (NSF) programs: an elementary school program entitled Science for Life and Living: Integrating Science, Technology, and Health; a middle school

A Career of Opportunities    143

program entitled Middle School Science & Technology; a high school biology program titled Biological Sciences: A Human Approach; and a program primarily used in community colleges, Biological Perspectives. My work at BSCS also included serving as principal investigator for projects to develop curriculum frameworks for teaching about the history and nature of science and technology for biology education at high schools, community colleges, and four-year colleges, and curriculum reform based on national standards. As part of my work on the Science for Life and Living project, along with colleagues, I developed the BSCS 5E instructional model, a sequence of teaching strategies described as Engage, Explore, Explain, Elaborate, and Evaluate. This instructional model is now widely used by state departments of education, curriculum developers, and science teachers. It also has been a central feature of many BSCS programs since the late 1980s. In 2006, with support from the National Institutes of Health, Office of Science Education, we completed a review of the instructional model (Bybee et al., 2006). Designing, developing, and implementing curricula became a major part of my job at BSCS. The complexity of the task challenges one’s knowledge, abilities, time, and budget. Multiple other criteria must be considered, including accurate science content; appropriate developmental levels; contemporary research on learning, usability and manageability by science teachers; and marketability by publishers. At BSCS we had to consider these and other criteria and propose programs that were innovative but implementable in school systems. Serving as chair of the Education Department at Carleton did not really prepare me for the administrative duties at BSCS. As associate director, I had a wonderful mentor in Joseph D. McInerney, who was executive director. Joe introduced me to administrative functions such as strategic planning, conducting board of directors meetings, and yes, even hiring and firing personnel. Joe encouraged me to develop proposals and interact with national organizations (e.g., NABT), federal funders (e.g., NSF), and other funders (e.g., IBM). My work also included interactions such as contract negotiations with the private sector, namely commercial publishers. Ten years as associate director served as a wonderful apprenticeship to the internal workings of a non-profit organization and the importance of understanding external priorities and politics as they influence decisions about school science programs and classroom practices. Exploring a New World: The National Experience In 1992, I was invited to work on a project to develop national standards for science education. At first, I worked as a member of the content group. Later, I served as chair of that group. This experience provided new perspectives

144   R. W. BYBEE

on educational reform, especially the influence of national policies such as standards and on the various aspects of science education. The challenges associated with the development of national standards were numerous and often intense. But that is another story, too long and complex for this discussion. Working on content standards provided an opportunity to argue for the inclusion of standards that addressed issues aligned with human ecology. Truthfully, the task of including environmental issues in the national standards turned out to be relatively easy. As soon as we identified life science standards, content related to ecology was essential. In grades K–4, we included standards on “organisms and environments,” for grades 5–8 “populations and ecosystems,” and for grades 9–12 “interdependence of organisms.” The standards, however, presented another opportunity. Early discussions made it clear that we would have to acknowledge the connections between science and the decisions individuals make about various issues such as health and, yes, the environment. We grouped these standards under a title reminiscent of long standing social application goals for science education, “Science in Personal and Social Perspectives.” It was here that I proposed variations on the themes of population growth, resource use, and environmental quality. We were very clear about not including contemporary topics such as acid precipitation or global warming, even though they presented a considerable scientific debate and a wonderful opportunity to introduce the nature of science, as well as a host of issues ranging from atmospheric chemistry to emerging and re-emerging infectious diseases. When it came time to justify the content to the national board, I made the case that including a topic such as global warming would be too close to suggesting curricular topics, thus subjecting the standards to the criticism that they were political statements, and more importantly, setting the stage for a national curriculum for science. I argued that issues of populations, resources, and environments were fundamental to many contemporary problems and potential future problems, ones that applied to developed as well as developing countries. For example, understanding the dynamics of population growth, the difference in renewable and non-renewable resources, and the limited capacity of the environment to receive and degrade waste was essential scientific knowledge and not a political statement. Rather, the understanding included scientific knowledge essential for all citizens. The argument carried the day. This particular example brings out the lesson of having one’s own values and priorities and balancing them with the national politics of education. Just as working at BSCS required putting the organization before personal preferences, so too did work on national standards require an understanding of policies and their potential influence on science education programs and practices. Work on the national standards brought a whole new realm of insights about American science education and the limits and possibilities of re-

A Career of Opportunities    145

form. We had to be very sensitive to any statement of standards that would suggest a national curriculum. We had to design standards for grade-level sets—K–4, 5–8, and 9–12. This structure then resulted in criticism, because we were not clear about what exactly should be the science content for grade 3 or 6 or 10. We also had numerous individuals lobby for their program or organization’s priority. These various issues required the content group and smaller group consisting of Angelo Collins, Audrey Champagne, Karen Worth, Harold Pratt, and me to develop and maintain a clear understanding of national policies in science education. Working on the National Science Education Standards (NRC, 1996) was an opportunity to influence science education at a national level that also included review, criticism, and sensitivity to the entire science education system. Developing national standards included three clearly distinct groups— scientists, science educators, and science teachers. To say the least, these three groups did not have similar priorities for the standards. Scientists had the highest aspirations for accurate content often at advanced levels—for example, graduate work. We had to temper their views concerning what elementary children should and could learn. Science educators had priorities for research-based content and constructivist approaches. Our task here was to incorporate some of these ideas. One of the strongest and most difficult meetings centered on the scientists’ objections to any use of the term constructivism because of several post-modern publications. Finally, teachers often asked for concrete examples and activities that would help students “meet the standards,” so we included short vignettes illustrating sample standards. Our task was to mediate, moderate, and accommodate these very different perspectives. I have been amazed at the positive and constructive use of the national standards. Their influence is evident in state standards, textbooks, and the national assessment. I have been somewhat disappointed in the influence the national standards have had on state and local standards related to the aforementioned themes of science in personal and social perspectives and other standards we developed for teaching, assessment, programs, and systems. For reasons that are difficult to explain, state committees formulating standards for science generally have referred to the National Science Education Standards but have not seen the richness, depth, and clarity of all the standards. Developing the standards required careful and clear thought about what the standards were, how they would influence the science education system, and ultimately what students should learn. We had to do this without crossing the line of advocating for or against a particular point of view, except the obvious one of science and related content. As we were completing work on the national standards, Bruce Alberts, then president of the National Academy of Science, asked me to come by his office. I thought it was a meeting to thank me and others for our work.

146   R. W. BYBEE

After a few minutes of conversation, Bruce told me of his plans to consolidate a number of different initiatives and projects in the National Research Council (NRC), the working arm of the National Academy of Science, National Academy of Engineering, and the Institute of Medicine. Bruce then asked if I would be interested in heading this group. After a brief consideration, I agreed and was offered the position as executive director of the new center. In September 1995, I left the Biological Sciences Curriculum Study (BSCS) where I had been associate director since 1985 and joined the National Research Council as executive director of the newly formed Center for Science, Mathematics, and Engineering Education. From 1995 to mid1999, I worked in Washington, DC. My work in science education at the NRC centered on disseminating the National Science Education Standards. The standards were released late fall 1995 with a 1996 copyright. I also had responsibility for mathematics, engineering, and undergraduate science education, all of which had boards and active projects. I quickly learned to work with the boards and help them focus on priorities, projects, and finally reports that would be completed appropriately and adequately within the NRC system of objectivity, balance, and report review. During my five years at the NRC, we completed reports on topics that included: international assessments, the role of teacher education, addenda to the National Science Education Standards (NRC, 1996), the role of science and mathematics standards in state policies, and investigation of the influence of science, mathematics, and technology standards. Working with the NRC required continual meetings with federal agencies such as the National Science Foundation (NSF), National Institutes of Health (NIH), National Aeronautics and Space Administration (NASA), and the U.S. Department of Education (US DoED). These meetings had different goals. In some meetings, I was presenting proposals for initiatives suggested by one or more of the various boards in the Center. Other meetings were called by the agencies in an effort to pursue a project the agency had developed. There also were meetings to review programs and budgets, advise on results of projects, and provide key staff with information they could use in hearings “on the Hill.” My national experience at the NRC also included several testimonies before congressional committees and briefings for the Secretary of Education on U.S. results on international assessments. These experiences made clear the political aspects of science education and a gave a new appreciation for the careful statements and measured responses to questions from committee members or the Secretary and his deputies. My five years in Washington, DC provided new opportunities and growth beyond anything I had ever considered. Interacting with top scientists, government officials, members of Congress, the Secretary of Education, and

A Career of Opportunities    147

The White House brought broader and deeper perspectives on science education in the United States. Science education is more than classrooms, school districts, and state agencies. Little did I know how these experiences would prepare me for the next opportunities and challenges. Returning to BSCS: Organizational Possibilities At a BSCS Board of Directors meeting in the spring of 1999, Joe McInerney announced his resignation as executive director. Tim Goldsmith, chair of the BSCS board, approached me with an invitation to return to the organization as executive director. When I went to Washington in 1995, I intended to stay long enough to have a “Washington experience” but never planned to finish my career in Washington. The opportunity to return to BSCS came at a good time. I had been in Washington five years, and the chance to return to Colorado and BSCS was ideal. My vision for BSCS included maintaining the curriculum work, expanding initiatives in professional development, and initiating research and evaluation. While at the NRC, I had hired and worked with Susan LoucksHorsley and rapidly came to appreciate her leadership, especially in the area of professional development. I convinced Susan to join BSCS. She led the development of a proposal to establish a center that provided professional development for school districts implementing NSF supported science curricula for secondary schools. We worked on the proposal in the fall of 1999, and Susan joined BSCS in 2000. Susan died in an unfortunate accident later that year. NSF did fund the proposal, and a very courageous staff that included Nancy Landes and Jim Short initiated the program and made real the possibilities of professional development at BSCS. In January 2001, Janet Carlson returned to BSCS as associate director, and I asked her to establish a research and evaluation program as a part of our work at BSCS. In time, we had major projects in curriculum, professional development, and research and evaluation. I established BSCS centers for these complementary aspects of the organization. From 1999 to 2006, the Board of Directors supported work on a strategic plan and development of a business orientation for BSCS. We worked on our national brand, a consistent image conveyed in presentations, assuring a presence at national meetings and establishing an image of BSCS as an organization with a mission of curriculum development, professional development, and research and evaluation. I believe it is safe to say we accomplished all of the above.

148   R. W. BYBEE

Expanding Horizons: Opportunities for International Experience Early in my career, I had several opportunities to travel overseas for international conferences. I always found the perspectives gained from these meetings a counterpoint to those that only focus on issues in the U.S. The international experiences likely affected my work, but I was never in a position to really appreciate the benefits of the meetings. My time in Washington at the National Research Council and at BSCS changed my views by broadening them through several international experiences. One of the first significant international policy experiences came after I returned to BSCS. Ina Mullis, whom I had known for years, contacted me in 2000 with a request to join the International Study Panel in Mathematics and Science for TIMSS. While at the National Research Council, the Center for Science, Engineering, and Mathematics Education completed several reports on TIMSS and, as a result, I developed a deep understanding of this international assessment. My decision to join the TIMSS Study Panel was, as they say, a no-brainer. In 2000, Teresa Smith (now Teresa Neidorf), then of the International Study Center at Boston College, contacted me to help with preparation of TIMSS Assessment Frameworks and Specifications: 2003. I told Terry that my contribution would be directly related to the U.S. National Science Education Standards, because I believed they represented important science content for students to learn. And the national standards had been thoroughly critiqued as part of the report review process at the National Research Council. In my view, there was no better model for a science assessment framework; certainly none of the state frameworks would compare. Terry said that was precisely why I was invited to contribute. She described the importance of including basic science content and the need to include a vision that would inform future cycles of TIMSS as we entered the new millennium. Terry opened the door for inclusion of more than basic ideas in the science content domain. My task included describing life science, Earth science, and environmental science content for grades 4 and 8. Environmental science was included in the framework as a separate content domain to show the relative importance placed internationally on educating students about factors affecting the environment and resources. Although Terry Smith made it very clear that the material I submitted would be reviewed, edited, and in some cases omitted, work on the TIMSS framework represented something that I believed did represent a vision for the future. The basic contemporary problems, some of which varied among the countries, was important. I also was pleased to see consistencies between the National Science Education Standards and future international

A Career of Opportunities    149

assessments—TIMSS. I did not realize I would have another major opportunity to work on an international assessment of science. But I did. In June 2002, I received an email from Eugene Owens, a colleague I had worked with while in Washington. Gene directed international programs at the U.S. Department of Education, National Center of Educational Statistics (NCES). Gene asked permission to forward my name as a representative of the U.S. to participate in a Science Forum for the Program for International Student Assessment (PISA). I agreed because I knew and respected Gene. At the time, I knew little about PISA and the challenges that lay ahead. As I understood Gene’s invitation, I might not be accepted to join the assessment team. So, I gave little thought to the request. Several weeks later, I received an email from Ray Adams, a senior administrator at the Australian Council for Educational Research (ACER) in Melbourne. Ray asked if I could attend a meeting of the Science Expert Group for PISA 2003 to be held in Melbourne that October. He also asked if I could attend a Science Forum in Paris that December. And if I could, would I be willing to help him chair the meeting. So I agreed and that December, I traveled to Paris for the meeting that was held at the headquarters for the Organization for Economic Cooperation and Development (OECD). The Science Forum consisted of representatives from 30 countries that periodically meet to advise and monitor progress on the PISA science program. I was to chair the Science Forum and a subcommittee of the Forum called the Science Expert Group, which had responsibility for developing the framework for PISA 2006 Science. Developing the framework for 2006 presented another opportunity for me to advocate for the assessment of science in the context of issues related to the environment and resources. In addition to scientific knowledge, OECD leadership asked that we take on the additional, innovative, and difficult task of including an attitudinal dimension to the assessment. Along with Barry McCrae and other colleagues, the work on PISA 2006 Science survey resulted in several publications that detail the results and implications of PISA (Bybee, Fensham, & Laurie, 2009; Bybee & McCrae, 2009, 2011). Reflecting on a Career of Opportunities: Conclusion Reflecting on one’s work is at once a measure of success and of possibilities unrealized. One gains new perspectives and appreciations for those individuals and opportunities that have guided more than 40 years of a career. From an early age, most of us learned to be modest and humble about accomplishments. At least I did. It is hard to imagine, much less say, that I have been a pioneer and played a significant leadership role in science

150   R. W. BYBEE

education. Yet, my career has included elementary, middle, high school and college teaching; directing national organizations; participating in the development of national and international policies and programs; and influencing instructional practice. When I stop and think about it, I am overwhelmed by the opportunities I have had to influence science education. Working on the National Science Education Standards and current work on the Next Generation Science Standards certainly could be seen as a major achievement for any science educator in the United States. To say the least, it also was a major challenge and significant opportunity to influence national policies. Several years later, I was presented the possibility of affecting the development and implementation of an assessment that will impact the international community. Then came the opportunity to influence the U.S. National Assessment of Education Progress from 2009 to 2019 and the potential of an international assessment of environmental literacy in 2015. These opportunities are both remarkable and humbling. My work will have an impact on science education programs and policies, not only in the United States, but in 60 countries. I have been fortunate to be presented opportunities to advocate for my ideas. At the same time, there always have been challenges of time to do the work and, inevitably, the sacrifices. As I reflect on this work, one insight becomes clear—seldom are one’s views taken without review, reservation, and revision. Sometimes the revision improves the statement and other times . . . well, need I say more? A second insight, not unrelated to the first, centers on politics. As the night follows the day, there will be politics associated with the opportunities. In my career, these have sometimes been politics with a small p and other times with a capital P. How do I assess my contributions? Long ago I came to the realization that one person cannot bring about the major changes implied by the issues associated with reforming science education. That said, I deeply believe that each person has some responsibility to society and future generations. Blessed with numerous opportunities, I am comfortable saying that I have given back to society, contributed to some changes, and witnessed positive educational improvement. Was this easy? Not at all. Compared to the 1960s, educational change in the contemporary era presents confounding sets of views and approaches often offsetting the possibilities that one is presented. I return to personal revelations about my family history and being an adult-child of alcoholics. I truly believe that the experiences early in my life contributed to a set of characteristics that both motivated me to engage in the opportunities I was presented and to accomplish what I have. I must also make it clear that the positive outcomes in my professional life were countered with sacrifices, trade-offs, and issues in my personal life.

A Career of Opportunities    151

In the end, one has to overcome past experiences and the anxiety that attends each opportunity, and continue moving forward. As I look to the science education community, I believe all individuals have opportunities to make contributions. I conclude with the following: Blinded by our search for larger-than life leaders, we have not recognized the individuals or the forms and processes of leadership that will be most responsible for the contemporary reform of science education. We must realize that a majority, not a minority, of individuals in science education have leadership responsibilities. These include policy makers, curriculum developers, teacher educators, science supervisors, and classroom teachers. (Bybee, 1993, p. 149)

References Adler, M. (1982). The paideia proposal: An educational manifesto. New York, NY: Macmillan. Barbour, I. (1980). Technology, environment, and human values. New York, NY: Praeger. Boyer, E. (1983). High school: A report on secondary education in America. New York, NY: Harper & Row. Bronowski, J. (1973). The ascent of man. Boston, MA: Little, Brown. Brown, L. (1981). Building a sustainable society. New York, NY: W.W. Norton & Company. Bybee, R. (1969). A comparison between an individualized laboratory method and a lecturedemonstration method of teaching two general education earth science classes. Unpublished master’s thesis, Colorado State College, Greeley, Colorado. Bybee, R. (1970). The effectiveness of an individualized approach to a general education earth science laboratory. Science Education, 54(2), 157–161. Bybee, R. (1975). The implications of Abraham H. Maslow’s philosophy and psychology for science education in the United States. Unpublished doctoral dissertation, New York University, New York. Bybee, R. (1977a). The new transformation of science education. Science Education, 61(1), 85–97. Also reprinted in Bybee, R. (1993), Reforming science education: Social perspectives and personal reflections (pp. 5–19). New York, NY: Teachers College Press. Bybee, R. (1977b). Toward a third century of science education. The American Biology Teacher, 39(6), 338–341, 357. Bybee, R. (1979a). Science education for an ecological society. The American Biology Teacher, 41(3), 154–163. Bybee, R. (1979b). Science education and the emerging ecological society. Science Education, 63(1), 95–109. Also reprinted in Bybee, R. (1993), Reforming science education: Social perspectives and personal reflections (pp. 20–37). New York, NY: Teachers College Press. Bybee, R. (1979c). Science education policies for an ecological society: Aims and goals. Science Education, 63(2), 245–255. Also reprinted in Bybee, R. (1993),

152   R. W. BYBEE Reforming science education: Social perspectives and personal reflections (pp. 38– 51). New York, NY: Teachers College Press. Bybee, R. (1984). Human ecology: A perspective for biology education. Monograph Series II. Reston, VA: National Association of Biology Teachers. Bybee, R. (1985). The restoration of confidence in science and technology education. School Science and Mathematics, 85(2), 95–108. Also reprinted in Bybee, R. (1993), Reforming science education: Social perspectives and personal reflections as The Restoration of Confidence in Science Education (pp. 59–70). New York, NY: Teachers College Press. Bybee, R. (1987). Science education and the science-technology-society (S-T-S) theme. Science Education, 71(5), 667–683. Bybee, R. (1991). Planet earth in crisis: How should science educators respond. The American Biology Teacher, 53(3), 146–153. Also reprinted in Bybee, R. (1993), Reforming science education: Social perspectives and personal reflections (pp. 118– 132). New York, NY: Teachers College Press. Bybee, R. (1993). Reforming science education: Social perspectives and personal reflections. New York, NY: Teachers College Press. Bybee, R. (1997). Achieving scientific literacy: From purposes to practices. Portsmouth, NH: Heinemann. Bybee, R. (2003). Integrating urban ecosystem education into educational reform. In A. Berkowitz, C. Nilon, & K. Hollweg (Eds.), Understanding urban ecosystems (pp. 430–449). New York, NY: Springer. Bybee, R., Fensham, P., & Laurie, R. (Eds.). (2009). Special issue: Scientific literacy and contexts in PISA science. Journal of Research in Science Teaching, 46(8), 861–960. Bybee, R., & Hendricks, P. (1972). Teaching science concepts to pre-school deaf children to aid language development. Science Education, 56(3), 303–310. Bybee, R., & McCrae, B. (2009). PISA science 2006: Implications for science teachers and teaching. Arlington, VA: National Science Teachers Press. Bybee R., & McCrae, B. (2011). Scientific literacy and student attitudes: Perspectives from PISA 2006 science. International Journal of Science Education, 33(1), 7–26. Bybee, R., Powell, J., & Trowbridge, L. (2007). Teaching secondary school science: Strategies for developing scientific literacy. Upper Saddle River, NJ: Prentice Hall. Bybee, R., & Sund, R. (1990). Piaget for educators. Long Grove, IL: Waveland Press. Bybee, R., Taylor, J., Gardner, A., & Van Scotter, P. (2006). The BSCS 5e instructional model: Origins and effectiveness. Colorado Springs, CO: BSCS. Carson, R. (1962). Silent spring. Boston, MA: Houghton Mifflin. Cremin, L. (1976). Public education. New York, NY: Basic Books. Dobzhansky, T. (1956). The biological basis of human freedom. New York, NY: Columbia University Press. Ehrlich, P. (1968). Population bomb. New York, NY: Ballantine. Goodlad, J. (1979). What schools are for. Bloomington, IN: Phi Delta Kappa Educational Foundation. Hardin, G. (1968). The tragedy of the commons. Science, 162(3859), 1243–1248. Meadows, D., Meadows, D., Randers, J., & Behrens, W. (1972). The limits to growth. New York, NY: The American Library.

A Career of Opportunities    153 National Commission on Excellence in Education (NCEE). (1983). A nation at risk: The imperative for educational reform. Washington, DC: U.S. Department of Education. National Research Council (NRC). (1996). National science education standards. Washington, DC: National Academies Press. Odum. E. (1959). Fundamentals of ecology. Philadelphia, PA: W.B. Saunders Company. Robinson, J. (1968). The nature of science and science teaching. Belmont, CA: Wadsworth Publishing Company. Salk, J. (1973). The survival of the wisest. New York, NY: Harper & Row. Sund, R. (1967). Teaching science by inquiry in the secondary school. Columbus, OH: C.E. Merrill Books. Sund, R., & Bybee, R. (1973). Becoming a better elementary science teacher—A reader. Columbus, OH: Charles E. Merrill.

154   R. W. BYBEE

Chapter 8

Susan Loucks-Horsley, PhD Transformational Leader and Spark for Educational Change1 Susan Mundry STEM & Learning Innovations Programs, WestEd

Introduction Susan Loucks-Horsley had a smile that brightened every room she entered and a superior intellect and problem solving orientation. She was a national expert in many areas: educational leadership, school change, curriculum implementation, science and mathematics education, and more. But it wasn’t only what Susan knew—and she knew so much—that made her stand out as a pioneer in science education. What set her apart were her special gifts as a teacher, connector, and bridge-builder. She bridged the worlds of research and practice and connected networks of people around the nation to work together to transform education. Her keen understanding of how to collaborate with people and apply systems thinking to effect change, as well as her tireless work ethic and commitment to applying quality research and common sense to the difficult challenges facing education, were the

Going Back for Our Future, pages 155–176 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

155

156    S. MUNDRY

magic ingredients in her successful career. As Senta Raizen, who was a close colleague of Susan for many years, told me recently: Susan was remarkable because she was so good at collaboration. She could bring people together around solving a problem or producing a product, even people with differing perspectives and ideas. (Raizen, personal correspondence, 2011)

Susan was a great collaborator, and she cared deeply about people. Her hopes for her colleagues were that they would develop their talents as excellent education change agents and researchers, and she eagerly modeled, coached, and mentored to support their growth. Susan was first and foremost a teacher, and she continues to teach us today. When faced with a tough dilemma or work problem, my colleague, Kathy Stiles and I will still often say, “What would Susan do?” And we can usually guess, because Susan not only helped us learn what to do and how to do it, but also spent a tremendous amount of time focusing on why we should approach the work the way we do. She stressed the need for us to be crystal clear about what she called the “underpinnings” of our work. For example, she would clarify what research and models would drive our work and carefully articulate project goals and strategies aligned to the research. Kathy DiRanna of WestEd pointed out that Susan was a “talent magnet” not a micro-manager and a “debate maker.” She would “open the debate and encourage people to come up with the best ideas.” As a national expert, Susan established relationships with some of the most important thinkers and doers in the field. She wanted her staff and colleagues to be directly connected to these major thought-leaders of the time and made it happen by leading discussions of papers on cutting-edge research and developments in the field, always providing food and drink to liven the discussion. She invited experts to provide professional development for staff and serve as advisors on our projects and insisted that (and supported) the staff invest in their own learning by attending national and regional conferences and building relationships. As Dennis Sparks (2000), former executive director of the National Staff Development Council, wrote for an AERA Symposium on Susan’s contributions to education: Susan seldom, if ever, was the sole author of a book or article, choosing instead to use the writing process as a powerful means of professional development for herself and her collaborators. She continually pushed intellectual boundaries to advance practice and taught educators important new ways of thinking about professional learning.

In fact, because of the way Susan lived her life and conducted her work, she personally developed a multitude of leaders, both on our staff and in

Susan Loucks-Horsley, PhD    157

state departments of education, professional associations, school districts, and other organizations who have continued to be champions for Susan’s work and ideas. I am honored to be among the people who worked closely with Susan for almost 20 years and have benefited greatly from her friendship and mentoring. Committed to Teachers and Students Of the many clients we served in education, Susan was most devoted to teachers. As a former science teacher herself, Susan understood the work and the challenges teachers faced. She empathized with teachers who worked under impossible conditions without the training, leadership, and support they needed to be effective. Susan committed herself to improving the ways teachers learn on the job and to creating working conditions that support them to be effective. In 1987 she and colleagues wrote, “. . . supporting the continual development of teachers is critical to attracting and keeping the best and the brightest people in the profession. A person who has opportunities to learn and to grow can best provide such opportunities for our young people. Schools that are true learning communities for adults and children alike are the only kind that hold promise for the future” (Loucks-Horsley, Harding, Arbuckle, Murray, Dubea, & Williams, 1987, p. 1). In 2012, many years after Susan’s first call for the development of learning communities, professional learning communities in schools are beginning to blossom and show promise for providing the culture and support for teachers that Susan envisioned so many years ago. Rodger Bybee, former Executive Director of BSCS wrote of Susan’s commitment to teachers: Personally, she conveyed a belief that, given the opportunity, each science teacher had the potential to improve, and that each teacher wanted her or his students to learn science. Susan let all she touched know that she understood their concerns and recognized their daily struggles to change. She supported their dignity, integrity, and worth as individuals. Susan Loucks-Horsley clearly recognized that the central issue of reform is not educational material; the essential factor is how leaders think and respond to the personal concerns of teachers, how they learn, and what has meaning for them. (Bybee, 2001)

While Susan was dedicated to teachers and supporting them, she never took her eye off the real prize—ensuring an excellent education for all young people, especially those challenged by poverty, learning disabilities, and discrimination. In her bridge-builder role, she believed that we needed to make it possible for all children to understand and connect with the natural and the designed world around them. She spoke of the need for a scientifically literate citizenry as a necessity and a right for all in our coun-

158    S. MUNDRY

try. As described later in this chapter, her work establishing the National Center for Improving Science Education with Senta Raizen, Rodger Bybee and others provided a forum for Susan to advocate for making science education a basic right for all students, and to provide young people with curriculum that nurtured deep conceptual understanding and developed scientific dispositions and skills. Early Work in Education Susan began her career in education in the late 1960s as a 9th grade science and mathematics teacher at the Midland Independent School District in Midland, Texas. Her BA was in science and mathematics education, specializing in geology, from the State University of New York, in Binghamton. Susan started teaching at a time of great reform in science education, and she encountered some of the challenges we still face today to improve instruction to meet the needs of all students. She began a graduate program at the University of Texas Austin in Curriculum and Instruction, and she was drawn to the innovative work underway to understand how to support teachers to make changes in instructional practice. She began work through her graduate studies with Shirley Hord and Gene Hall on the Concerns-Based Adoption Model (CBAM). She became the assistant project director for the Research on the Concerns-Based Adoption Model, where she conducted research on educational change in schools and colleges and met many of the leading science educators she would continue to collaborate with throughout her career. She quickly put her talents to work, developing deep expertise in the areas of curriculum design and organization, evaluation, school change, and staff development, and developing tools for using CBAM in the field (Heck, Stiegelbauer, Hall, & Loucks 1981; Loucks, Newlove, & Hall, 1975). The basic assumptions of CBAM are that: (1) Change is a process, not an event; (2) change happens first with individuals and then institutions; (3) change is a highly personal experience; and (4) change involves developmental growth in both feelings (affective) and skills/knowledge (cognitive) (Hall & Hord, 2010). Susan would go on to apply this seminal model of educational change to all of her work. Its principles shaped her thinking about how to design professional development, implement the National Science Education Standards, and orchestrate school reform. One of the major studies Susan worked on during her early career involved research on the implementation of a revised science curriculum in 80 elementary schools. This work was done in collaboration with national science education leader Harold Pratt, who was science coordinator in the Jefferson County, CO schools and later an NSTA president. Susan often

Susan Loucks-Horsley, PhD    159

spoke of her deep admiration of Harold for his thoughtful and pragmatic approach to guiding science education and for his work developing local leaders to improve science learning. This study contributed significantly to Susan’s subsequent work and was at the intersection of two of her major interests: understanding (1) how teachers come to make changes in their practices and (2) how to design and provide professional development based on research on adult learning and educational change. Susan left the University of Texas in 1981 to join the staff of The NETWORK, Inc., a research and development organization in Andover, MA that was a dynamic environment for promoting educational innovation and, later, housed the federally funded Regional Laboratory for Educational Improvement in the Northeast and Islands. Susan served as a senior researcher on one of the largest studies of that time on the dissemination of educational innovations, called Dissemination Efforts Supporting School Improvement (DESSI). This position afforded her the opportunity to work closely with giants in the field of educational innovation, including Matt Miles from Teachers College, Columbia University, whose early research on organizational development and school reform produced seminal work informing educational change, and Michael Huberman, an internationallyrecognized theorist, researcher and consultant on education in the United States, Canada and Europe. She co-authored a series of groundbreaking research reports with these researchers, with David Crandall, the former director of the Regional Education Laboratory, and others. Her research revealed substantial new ideas about how to introduce and sustain innovations in classrooms and schools, and Susan was eager to help local practitioners apply these important findings to their work (Loucks, Cox, Miles, Huberman, 1982; Crandall & Loucks, 1983). I first met Susan when I joined The NETWORK in 1982, and what stood out to me was her pure dedication to the work and the joy she derived from it. She worked hard, but Susan also loved to have fun, plan dinners, listen to live music, and just be with people she enjoyed. She could attend a meeting in the morning and engage her colleagues in deep, thoughtful work (as if she had nothing else pressing on her time) and then go to her office and write a stunning paper or proposal, often due the next day! Then after all that, she would round up a few people to go out for dinner. Although I was a few years younger than Susan, it was hard to keep up with her. I often marveled at how she was able to come into a project and quickly understand what needed to be done and how to address problems. She always had a way of adding value. In baseball terms, Susan was both a “utility infielder,” able to play any position with flair, and a reliable “closer”—often brought in at the end to add value to a promising project or to save a losing one. In the “closer” role, she almost always provided help to

160    S. MUNDRY

our staff on how to translate research and develop user-friendly tools and products useful to practitioners. After completing the DESSI study and personally editing the ten volumes of reports, Susan turned her attention to disseminating the important findings so they could be used. To reach the practitioner audience, Susan convinced the Association for Supervision and Curriculum Development to dedicate an entire issue of its journal, Educational Leadership, to the findings of the study. Susan authored the lead article, entitled “At Last: Some Good News from a Study of School Improvement” (Loucks, 1983). In it she writes that the study had found “classrooms where teachers and students pursue learning with excitement; in schools where goals are set simply so they can be achieved and higher ones aspired to; and in districts where careful planning and high-quality development work contribute to challenging experiences for students” (p. 4). The article identifies five factors that contributed to success in these schools: 1. “. . . the commitment of teachers, developed either through actual use of the new practice or before they began to use it,” 2. “curricula or instructional practices . . . that were carefully developed, well-defined, and determined to be effective,” 3. “training by credible people—often former teachers—that included follow up activities,” 4. “assistance and support by an array of players, including other teachers, principals, district staff, and external trainers and linkers,” and 5. “attention to . . . institutionalization (ensuring that the new practice remained), including line items on budgets, orienting new or reassigned staff and writing the new program into curriculum guidelines” (Loucks, 1983, p. 5). Susan took these research-based lessons learned into the field by reaching out to educators across the nation to encourage them to make wiser, research-based choices as they enacted educational reforms. She underscored the need for common sense when choosing educational innovations—that teachers should only be asked to implement practices that are clearly-defined, classroom-friendly, and proven as effective. She also cautioned that teachers should implement new practices with fidelity, without tinkering with them or dropping key elements. Without fidelity of implementation, she warned, teachers may do all the work but not achieve improvements in learning. In another article in the series, she challenged administrators who were planning school improvement efforts to ensure that they provide “ample, appropriate help” (Loucks & Zacchei, 1983, p. 27) and that they learn to be change leaders and provide clear direction. She writes, “ . . . with clear and

Susan Loucks-Horsley, PhD    161

continuous direction and assistance from administrators, improvement efforts resulted in committed, skilled teachers who described many benefits of using the [new] practice” (p. 28). Susan coached educational leaders to incorporate new practices into the life and culture of the school or district. For example, rather than allowing just the few volunteer teachers to enact new practices, Susan counseled administrators to build a critical mass of teachers in each school to work together so teachers could support and help each other so that the new practices would stick. Building on this work, Susan and colleague Leslie Hergert (now at the Education Development Center) wrote a book, An Action Guide to School Improvement (Loucks-Horsley & Hergert, 1985). In it they provided a sevenstep planning and implementation process to guide school reform and offered research-based guidance and examples for each step. This work was significant because it stressed the need for leaders to carefully guide school improvement and to pay attention to selecting solutions or innovations for their schools that fit with their contexts. It was at this time that Susan began to challenge the “one size fits all” thinking that was often applied to school improvement. The DESSI study resulted in a ten-volume set of scholarly research reports, another dozen journal articles written for practitioners, a practical “how-to” guide for school improvement, and workshops and professional development for educational change leaders to help them use this research. Susan would go on to apply this same comprehensive dissemination formula to her work in the National Center for Improving Science Education, National Institute for Science Education and at the National Research Council to bridge the worlds of research, policy, and practice. The 1980s and 1990s were the heyday of educational innovation, and Susan stood in the middle of it all. Studying and implementing innovations were central to her work. At the time, the U.S. Department of Education had a policy of supporting not only development of innovation, but also the wide-scale dissemination of effective practices. Susan applied the learning from her research and wrote about the specific steps needed to support teachers to adopt, implement, and sustain instructional improvements. She directed a number of technical assistance projects aimed at supporting educators to adopt these research-based principles. Innovation was a centerpiece of Susan’s career—she was very good at putting together hunches and ideas about what might work in education. In the early 1980s she got excited about the potential for using technology to support learning. While it is hard to believe in 2013, in the early 1980s few educators were embracing the idea that technology could change how students learn. Susan collaborated with the technology firm Bolt, Beranak and Newman in Cambridge, MA to develop and conduct research on the impact of a computer-based writing program called QUILL. The QUILL

162    S. MUNDRY

product was developed, and research results showed that the program significantly improved students’ abilities to plan and execute writing tasks. Based on these promising results, Susan pursued a grant from the U.S. Department of Education to disseminate the QUILL program throughout the country. She used the knowledge from CBAM and her research on educational change to design a training program and follow-up assistance for teachers to ensure that teachers would gain the knowledge and skills needed to use this new technology in their classrooms. The program was widely adopted throughout the country, and it supported countless teachers and students to become comfortable using computers in the classroom. Her early career resulted in a deep understanding of education change and all of the necessary components to support change—professional development, leadership, proven practices and quality curriculum, and effective policy. Susan would go on to apply this rich background to science education. Taking on Science Education Reform Throughout the 1980s, Susan continued to be connected with science educators involved in her earlier work at the University of Texas, especially in Colorado where she provided technical assistance to Jefferson County and worked with BSCS around processes for curriculum implementation and teacher professional development. She was influenced by a group of close friends who were all working in science education—Rodger Bybee, Karen Worth, Mark St. John, Harold Pratt, Paul Kuerbis and others. Susan became more and more convinced that to improve science education and to provide effective science instruction to many more students, educators needed to focus on improving teacher quality and implement new curriculum and other innovations more effectively to produce and sustain changes in practice. Susan believed that change in teacher practice and ongoing improvements were essential to improve student outcomes, especially to reach all students with quality science experience. Her background and experiences had taught her that changes in teacher practice would be elusive until we applied earlier learning about organizational change to the problem of improving science education. In the late 1980s, the U.S. Department of Education proposed funding for a national research center on science education. Susan and several of us working at The NETWORK, Inc. saw this center as an opportunity to provide relevant guidance to the field about the need to improve science education. We invited Senta Raizen, a former program officer at the National Science Foundation and National Institute for Education, and Rodger Bybee, a noted science education expert, to work with us to establish

Susan Loucks-Horsley, PhD    163

the National Center for Improving Science Education (NCISE). We competed for and were awarded the center contract in 1987. Senta Raizen was the primary founder and served as the NCISE director for over 20 years. Susan was NCISE associate director and led the research on teacher development. Rodger Bybee and colleague, Paul Kuerbis of Colorado College, led the work on curriculum and instruction. Senta, Susan, and Rodger developed a unique vision for NCISE as a place that would synthesize existing research to produce guidance for policymakers and practitioners to inform the field in three critical areas: science curriculum and instruction; student assessment; and teacher professional development. They shared the belief that there was substantial knowledge about what was needed in science education, but that the knowledge was not widely known or available. NCISE took on bridging the gaps among research, practice, and policy and aimed to promote cooperation and collaboration among the many organizations working to improve science education. State and local education agencies across the U.S. drew upon the center’s resources to inform systemic improvements in science education. The Congressional Office of Technology Assessment and the National Education Goals Panel drew upon NCISE’s reports, and these reports also informed the development of the National Science Education Standards. From her position as associate director of NCISE, Susan immersed herself fully in the world of science education reform and became a champion for transforming teacher development. The research agenda for the first year was to synthesize the research on elementary science. Susan led the panel focused on the development of elementary teachers in science and wrote a report that challenged the status quo. She and her colleagues wrote, . . . the rhetoric of cooperation between science faculties and education school faculties will have to become a reality. The long-lamented obstacles to better coordination between the colleges and the schools will need to be removed . . . In our view, major shifts in institutional norms and boundaries are required if policymakers and educators at all levels are seriously interested in creating a scientifically literate population in the next generation. (NCISE, 1989, p. 39)

Susan and a panel of experts laid out a vision for reforming teacher preparation, induction, and professional development to fundamentally change teaching. In keeping with the mission of NCISE to inform the field, Susan brought this research directly to practitioners by presenting at conferences held by the National Science Teachers Association (NSTA), the Association for Supervision and Curriculum Development (ASCD), the National Staff Development Council (NSDC), the Association of Teacher Educators (ATE), and the National Association for Research in Science Teaching (NARST).

164    S. MUNDRY

Susan encouraged her colleagues to apply the findings from the center’s work to the development of a model of professional development for elementary teachers. With her support, Maura O’Brien Carlson, Gregg Humphrey, and others at The NETWORK, Inc. developed the Vermont Elementary Science Project to put the center’s recommendations into practice. This model changed teacher professional development in at least two fundamental ways. First, teachers were completely engaged in inquiry as the focal point of their learning. Second, the model included embedded formative assessment through which teachers learned to observe and listen carefully to their students’ ideas. In addition, Susan, Maura, and others wrote a practitioner-oriented guide, published by ASCD and called Elementary School Science for the 90s, to encourage others nationally to transform their thinking about how to prepare teachers and how to instruct students (Loucks-Horsley, Kapitan, & Carlson, 1990). Maura, Gregg, and Karen Reinhardt would go on to write another book on their model for formative assessment in science classrooms (Carlson, Humphrey, & Reinhardt, 2003). In subsequent years, the center produced research syntheses for middle school and then high school echoing some of the same themes and calling for more rigorous curriculum, formative assessment of student learning, and ongoing teacher development. Over these years Susan deepened her commitment to reforming professional development. She was compelled by findings from a 1993 national survey of teachers conducted periodically by Iris Weiss (1994) of Horizon Research, Inc. that reported that “76 percent of elementary teachers assigned to teach all four subjects indicated they felt very well qualified to teach reading/language arts, compared to roughly 60 percent for both mathematics and social studies, but only 28 percent for life science” (p. 8). Susan dedicated herself to working with in-service and pre-service leaders to change the ways teachers are prepared and supported to teach science. The center research was conducted “pre-national standards,” but soon work on the national standards at the National Research Council (NRC) and on benchmarks at the American Association for the Advancement of Science (AAAS) was underway, and Susan was committed to emphasizing the knowledge gained about teacher professional development in the national standards. Starting in 1992, she served on the Working Group on Science Teaching Standards for the National Science Education Standards (NRC, 1996). A few years later, Susan accepted a position as the Professional Development Director at the National Research Council and became a constant advisor to the team at the National Research Council rolling out national standards, led by Rodger Bybee. She also continued to work at a reduced percent in her job as the then co-director of the Math and Science program at WestEd. She had two “full time” jobs but was extremely produc-

Susan Loucks-Horsley, PhD    165

tive and happy and somehow made it work, spending half of her time in Washington and the other half in her WestEd office in Arizona. Just as Susan had conceptualized a full service approach of research to practice for the National Center for Improving Science Education, she and Rodger Bybee applied the same thinking to supporting the adoption and use of the National Science Education Standards (NRC, 1996). They, along with Harold Pratt and others at the NRC, identified the need to develop practical tools on inquiry, curriculum selection, and assessment that would help educators understand and apply the standards. They also crafted a plan to develop leaders across the nation who would support the implementation of the standards. They co-authored articles and chapters that recommended that schools provide professional development to implement the standards and discussed what successful implementation of the standards might look like (Loucks-Horsley & Bybee, 1998; Bybee & LoucksHorsley, 2001). They write: Before implementing the Standards, a short pause is warranted. At this stage of the journey, it is imperative to determine how we will know when we have completed this task. The ultimate answer is that all students will have achieved the outcomes specified in the con­tent standards. But at intermediate points along the journey we need to ask questions such as, “If we are successful in 2, 5, or 10 years, what will look different? If we look in classrooms, schools, and communities, what will we see that is different from today?” Such visualization is necessary to keep sight of our objectives. Science teachers simply do not have the time or money to waste making changes without a clear image of their destination and the intermediate points along the journey. (LoucksHorsley & Bybee, 1998, n.p.)

As science educators now plan to implement the Next Generation Science Standards, it is timely to revisit the recommendations that Susan and her colleagues offered for the NSES implementation and learn from this early work. Susan and her NRC colleagues clearly understood what it would take to bridge policy and practice and support people to make changes in science education. National Institute for Science Education In 1995 Andrew Porter (then at the University of Wisconsin-Madison) invited the National Center for Improving Science Education to form a partnership with the Wisconsin Center for Education Research (WCER) to pursue a National Science Foundation grant for a research institute. NCISE joined with WCER and competed for and won a five-year grant to create the National Institute for Science Education. This center provided the environ-

166    S. MUNDRY

ment and the national venue for Susan to transform beliefs and practice in the area of teacher development. As Dr. Porter wrote in 2000 at an AERA event honoring Susan’s career, From the beginning of NISE, Susan Loucks-Horsley had a profound effect on our entire community. Committing to produce a book by the end of the first year of NSF funding seemed an ambitious—if not unattainable—goal. Susan and her coworkers did produce the book. It was a huge success and continues to have a profound effect on the thinking of people across the country as they design, deliver, and do research on professional development for teachers of science and mathematics. The book was NISE’s first “home run.” Susan’s having set and met such a goal inspired all of us in NISE to demand that level of accomplishment from ourselves. (n.p.)

The book Andy was referring to is Designing Professional Development for Teachers of Science and Mathematics (Loucks-Horsley, Stiles, Mundry, Love, & Hewson, 2010), now in its 3rd edition and widely used by education leaders and staff developers. Susan often shared the story that writing this book transformed her own thinking about professional development. In 1989 she and then executive director of the National Staff Development Council, Dennis Sparks, wrote an article for the Journal of Staff Development outlining five overall models for professional development and the key assumptions and research supporting their use. The models include: (1) individuallyguided staff development, (2) observation/assessment, (3) involvement in a development/improvement process, (4) training, and (5) inquiry (Sparks & Loucks-Horsley, 1989). Susan’s idea for the first NISE book on professional development was to expand on the descriptions of these five models specifically for science and mathematics teachers. Susan identified science and mathematics professional development experts who she invited to collaborate on the book. The panel included a group of respected leaders in science and mathematics reform: Hubert and Rebecca Dyasi from the Workshop Center at City College, New York; Susan Friel, a professor of mathematics education and curriculum developer at the University of North Carolina; Judith Mumme, who directed the state-wide Math Renaissance project for the state of California; Cary Sneider, director and curriculum developer of Global System Science at the Lawrence Hall of Science, University of California, Berkeley and later, vice president of the Museum of Science, Boston; and Karen Worth of the Education Development Center and faculty at Wheelock College. Susan convened the collaborators and staff and began a dialogue about her ideas for developing “thick and rich descriptions” of the models for science and mathematics professional development. As the discussions evolved, the collaborators challenged the idea that we should write about models, raising the concern that models might imply a rigidity that would

Susan Loucks-Horsley, PhD    167

not reflect the many different contexts and circumstances in schools and districts. The group began to coalesce around the idea that what the field of science and mathematics education needed is a better understanding of the professional development design process and how one considers important inputs and makes decisions to create and provide effective professional development designs. We were excited about this new direction for the work. The group began to use the metaphor of a bridge to think about professional development. As Susan wrote: “A bridge, like professional development, is a critical link between where one is and where one wants to be . . . Each bridge requires careful design that considers its purpose, who will use it, the conditions that exist at its anchor points . . . and the resources required to instruct it” (1999, p. 2). The collaborators and staff shifted their focus to understanding the major inputs and processes that are essential to effective professional development and developed a design framework to inform professional development planning, implementation, and evaluation (See Figure 8.1). Once again, Susan turned her attention to how to get the ideas in the book applied to practice. She suggested that the field needed visual examples of what professional development looks like in practice and to hear the thinking of professional developers themselves about why they design programs they way they do. We proposed a major project with WGBH, Boston to videotape 18 examples of professional development and to develop facilitation materials that would help professional developers learn from these examples. We produced the video series and facilitation guide Teachers as Learners: Professional Development in Science and Mathematics (Mundry

Figure 8.1  Professional development design framework (Loucks-Horsley, Stiles, Mundry, Love, & Hewson, 2010).

168    S. MUNDRY

& Dunne, 2003). The series provided vivid examples of what different professional development strategies such as case discussion and immersion in inquiry looked like in action. We continue to build on this work today. The book is now in its third edition, the video series has been widely used to support professional developers, and we have recently developed a simulation game and facilitation materials to help leaders apply the principles of the framework to their work. Susan’s work on professional development continues to have a significant impact in the field. Building Leaders for Science Education Susan also worked with NRC staff and colleagues from WestEd to develop the State Leadership Institute for science and math leaders and supervisors to provide them with the know-how to be advocates for the national standards. In collaboration with the Council of State Science Supervisors, Susan and WestEd staff Kathy Stiles and Kathy Dunne created a model for developing leaders for the science standards. The curriculum they developed was guided by Susan’s careful thinking about what leaders needed to know and understand to lead significant reform, stemming from her early work on CBAM and her research on school improvement. It included time to learn the underpinnings of the standards—the what as well as the why— and to develop the necessary skills for advocating for change and facilitating groups to integrate the standards into their work. This was one of the happiest and most optimistic times I ever saw Susan—she was thrilled with the early success of the state science academy and delighted at how receptive and even “hungry” the science and math leaders were for the professional development and support she and her team provided. Participants claimed that their thinking about science and mathematics education was transformed by the experience, and they eagerly networked with one another about how to bring about science education reform in their states. At this same time, the federal government began to make a large investment in state, local, and urban systemic reform for science and mathematics education. This was the perfect context for Susan’s work—she consulted with many states from Massachusetts to California on their National Science Foundation State Systemic Initiatives and stressed that systemic reform models address instruction that focused on depth over breadth, learning over time, collaboration, and community building. She labeled one time professional development as the “sizzle without the steak” and “flash in the pan” change. She advocated for carefully applying the research on educational change and common sense to the investments being made in science

Susan Loucks-Horsley, PhD    169

reform. She would ask—what is it going to take to make it (the reform) stick? For example, she was hired to be the external evaluator for the California state systemic initiative, CAMS. She began working with the team and helped them to gather and use data. She modeled data use and guided the team to see that they needed to change what they were doing because it did not seem to be working. She guided the leadership by asking, how can you rethink what this looks like? According to Kathy DiRanna, who was state science director for CAMS at the time, Susan provided them with the insight that “we can’t do this alone, we need to build a community and develop many change agents.” Susan also believed that there was a great need for leadership development to prepare the younger generation to sustain improvements in science education. She feared that all the “gray hairs” and elders of the field would begin to retire and there would be few who understood the research on change and professional development. She envisioned creating a leadership academy for new leaders for science education who would have the opportunity to meet and be mentored by the current leadership. We worked together to seek funding from the NSF and began the National Academy for Science Education Leadership at WestEd in 1997. The Leadership Academy has since helped about 200 educators develop and enhance their leadership knowledge, skills, and strategies. Academy participants learned to use some of the same research findings and strategies in professional development and organizational change that guided Susan Loucks-Horsley for years and, more importantly, they became connected to a network of collaborative professionals in science education. Kathy Stiles and I still conduct the leadership academies for organizations and school districts interesting in developing their next generation of science leaders. We co-authored a book based on our Leadership Academy curriculum (with Susan and Joyce Kaser), Leading Every Day: 124 Actions for Effective Leadership (Kaser, Mundry, Stiles, & Loucks-Horsley, 2002), which received a Book of the Year Award from Learning Forward, and we are now preparing the third edition. Therefore, Susan’s work to invest in the leaders of tomorrow continues. Primary Aims and Career Focus As this review of Susan’s career and countless contributions to the field of education reveals, through all of her scholarship, Susan focused on applying research, building capacity and leadership, and finding solutions. She empathized with teachers who were continuously asked to try new and better ideas and practices every year, but who were not given the support, or the necessary conditions, to make new practices work effectively. This con-

170    S. MUNDRY

dition of education frustrated Susan, and she believed that our work had to focus much more on developing the capacity of leaders and teachers to be more successful. She succeeded in doing this by teaching, mentoring, and leading others. Susan’s aim was to bring great ideas to reality in schools. Many outstanding researchers in the field informed her thinking and helped shape her work. She was greatly influenced by the science education leaders I have already mentioned as well as the work of Matt Miles, Michael Huberman, Milbury McLaughlin, Michael Fullan, Dennis Sparks, Gene Hall, and Shirley Hord in the area of school and organizational change. In the area of teacher development, she was highly influenced by Bob Garmston’s work on coaching, Judith Warren Little’s work on building community, Bruce Joyce and Marsha Weil’s work on models of teaching, Lee Shulman’s work on pedagogical knowledge, and Deborah Ball and others who contributed to our understanding of mathematics pedagogical content knowledge. At the time of her death in 2000, Susan had just begun her job at BSCS as the associate executive director, and she eagerly anticipated guiding BSCS’s work and had also become very interested in applying research on cognition to our work. I have no doubt that she would have continued to deepen her interest in this new research and worked actively to put its principles into practice. Like any of us attempting the challenging work of educational reform, Susan faced obstacles. As an outside consultant and researcher, she did not have control to mandate changes in schools, nor to work deeply in one school district or setting over time. She struggled as we all do, with the difficulty of leadership turn-over that sometimes caused good ideas and practices to be abandoned, and for improvement efforts to be stalled. Interestingly, however, I never heard Susan talk about these as obstacles. Rather, she saw them as contextual challenges that needed to be addressed as part of the role we played as external technical assisters. She focused on how to ensure that there was a critical mass of teachers using new practices, as quickly as possible, and on building broad leadership in a school or district so that when one person left, others were there to carry on. Her optimism and problem-solving approach led her to focus on solutions rather than lament the obstacles. There were several major facilitating factors that contributed to Susan’s work. First was her belief that we must “stand on the shoulders of giants.” Susan always involved the “giants” in her work, either in person or by drawing upon seminal work in the field. When starting a new project or faced with a challenge, one of the first things Susan asked was, “Who knows something about this, and how can we tap that knowledge to inform our own thinking and work?” Many of the people I have mentioned in this chapter were those “giants,” and there were many, many more. She believed that isolation was

Susan Loucks-Horsley, PhD    171

the enemy of improvement and that we needed to create collaborative communities where professionals could build upon one another’s ideas. Second, Susan’s dispositions and work ethic facilitated her work. She had tremendous focus and vision and managed to keep the big picture in mind, even as she tackled the details of a major project. She was a natural data-user and constantly analyzed what was working and what was not and made ongoing changes to improve outcomes. Third, Susan attracted and developed talent. She surrounded herself with people who could deliver and who she genuinely liked to work with. She invested in people, pushed their thinking, and supported their growth. This built teams of people with commitment to each other and to the work. Fourth, Susan was credible. She did what she said she would do and established herself as a reliable, knowledgeable professional who would add value to any endeavor. As a result, she was able to influence many people to consider and adopt new ways of working and thinking. Since so much of our work involves persuading others to do things they may not be inclined to do, credibility is essential. Summary of Major Contributions to the Field As discussed in this chapter, Susan contributed substantial and lasting ideas and models to the field. While at the Research and Development Center for Teacher Education at the University of Texas, Austin, she worked on the development team of the Concerns-Based Adoption Model (CBAM). It is a model that has been used and adapted for multiple purposes over the past 25 years. Her other major works include: 1. Designing Professional Development for Teachers of Science and Mathematics (Loucks-Horsley, Stiles, Mundry, Love, & Hewson, 2010) 2. Facilitating Systemic Change in Science and Mathematics Education: A Toolkit for Professional Developers (Regional Educational Laboratories, 1995) 3. Continuing to Learn: A Guidebook to Teacher Development (Loucks-Horsley, Harding, Arbuckle, Murray, Dubea, & Williams, 1987) 4. An Action Guide to School Improvement (Loucks-Horsley & Hergert, 1985) 5. Elementary School Science for the ’90s (Loucks-Horsley, Kapitan, & Carlson, 1990) 6. Leading Every Day: 124 Action for Effective Leadership (Kaser, Mundry, Stiles, & Loucks-Horsley, 2002) 7. Teachers as Learners: Professional Development in Science and Mathematics (Mundry & Dunne, 2003)

172    S. MUNDRY

8. Key reports from the National Center for Improving Science Education on teacher development and support 9. Numerous chapters and articles that provide guidance on the implementation of educational innovations and the National Science Education Standards Most notably, Susan knew how to get the ideas and models represented in her body of work out into the hands of policymakers and practitioners alike so that they were put into action. Legacy of her Work Susan left a lasting legacy in her writing about effective professional development, educational change, and how to improve science education, but more importantly she left behind the talented people she mentored. In turn, many of these people are now mentoring new leaders for the future. As Kathy DiRanna, director of the California statewide K12 Alliance recently told me, From the very beginning, Susan was extremely vocal and an advocate for inquiry learning, student-centered instruction, and science for all kids. She pointed out the contradiction that so many in the field were pushing for student-centered learning while most professional development for teachers was still primarily lecture. She validated our work and pushed our thinking to lead science education to the next level.

The other way that DiRanna pointed out that Susan is still informing current initiatives is that Susan made sure that all of her colleagues were connected. Before social networks were popular, Susan built networks among everyone she knew. She used the occasion of national conferences to arrange for people she worked with in California or Florida or Colorado to meet the people she was working with in Texas and Massachusetts and Illinois. I remember so many times Susan would ask me, “Do you know so and so?” If I said no, she would say, “You have to meet him/her, I will arrange it”—and she did. As Dennis Sparks wrote, When I last saw Susan . . . she told me that she saw her current work as introducing the math and science communities to the professional development community so that we could learn from and support one another’s efforts to improve teaching and learning. She continually sought ways to connect people to one another and to ideas, often doing so behind the scenes and always without desire for acknowledgement or recognition. (personal communication)

Susan Loucks-Horsley, PhD    173

Long after Susan made these connections, we are all still connecting with one another and bringing more people into our “network.” Our very first cohort of WestEd’s National Leadership Academy in 1997 included up and coming leaders for science. In the subsequent years, many of them have made outstanding contributions to science education themselves, such as writing their own books, implementing exciting projects, and serving as national leaders for organizations such as NSTA. They are still out there making a difference for students of science. Susan’s legacy is also evident in how she approached her work. As Dr. Glen Harvey (2000), CEO of WestEd, observed: Susan “address[ed] real problems, working with real people, to create real solutions. Susan was, in many ways, of the earth. She could never be accused of dwelling in the proverbial ivory tower. The issues she took on were real; they were gritty. She sought the difficult issues that she thought would make a difference in education.” She went on to say that we must keep the themes of Susan’s work alive in the work we do, and my colleagues and I have made this commitment. Harvey also wrote: We need to address real problems, partnering with the practitioners and policy makers who struggle to support kids daily; our solutions need to meet the test of utility, quality, and making a demonstrable difference. We need to be accountable for our work and its results. And we need to hold tight to our passion. (personal communication)

Susan has left us with a tremendous body of work that we can build on and continue to apply. Rodger Bybee summed up how so many of us feel about Susan: Her life made a difference in the lives of others; now we have lost one of our best and brightest. We are left with her inspiration and dedication. Fulfilling her vision passes to all of us. As we look to the future without Susan LoucksHorsley, we can be thankful for her professional achievements and contributions, and we must be grateful for her personal inspiration and grace.

Note 1. I wish to acknowledge Katherine Stiles and Rodger Bybee for reviewing and providing input to this chapter.

References Bybee, R. W., & Loucks-Horsley, S. (2001). The National Science Education Standards as a catalyst for change: The essential role of professional development.

174    S. MUNDRY In J. Rhoton & P. Bowers (Eds.), Professional development planning and design (pp. 1–12). Arlington, VA: NSTA Press. Bybee, R. W. (2001). Dedication. In J. Rhoton & P. Bowers (Eds.), Professional development planning and design (dedication page). Arlington, VA: NSTA Press. Carlson, M. O., Humphrey, G. E., & Reinhardt, K. S. (2003). Weaving science inquiry and continuous assessment: Using formative assessment to improve learning. Thousand Oaks, CA: Corwin Press. Crandall, D., & Loucks, S. (1983). Volume X: A roadmap for school improvement. Andover, MA: The NETWORK, Inc. Hall, G., & Hord, S. (2010). Implementing change: Patterns, principles, and potholes (3rd ed.). Boston, MA: Allyn & Bacon. Harvey, G. (2000, September). Unpublished paper presented at AERA Symposium. Heck, S., Stiegelbauer, S., Hall, G., & Loucks, S. (1981). Measuring innovation configurations: Procedures and applications. Austin, TX: University of TX at Austin. Kaser, J., Mundry, S., Stiles, K., & Loucks-Horsley, S. (2002). Leading every day: 124 actions for effective leadership. Thousand Oaks, CA: Corwin Press. Loucks, S. (1983). At last some good news from a study of school improvement. Educational Leadership, 41(3), 4–5. Loucks, S., Cox, P., Miles, M., & Huberman, M. (1982). Portraits of the changes, the players and the contexts. A study of the dissemination efforts supporting school improvement. People, policies and practices: Examining the chain of school improvement, Vol. II. Network of Innovative Schools, Andover, MA. ERIC no. ED240714. Loucks, S., Newlove, B., & Hall, G. (1975). Measuring levels of use of the innovation: A manual for trainers, interviewers, and raters. Austin, TX: University of TX at Austin, Research and Development Center for Teacher Education. Loucks, S., & Zacchei, D. A. (1983). Applying our findings to today’s innovations. Educational Leadership, 41(3), 26–32. Loucks-Horsley, S. (1999). Effective professional development for teachers of mathematics. In Eisenhower National Clearinghouse, Ideas that work: Mathematics professional development. Columbus, OH: Eisenhower National Clearinghouse. Loucks-Horsley, S., & Bybee, R. W (1998). Implementing the National Science Education Standards. Science Teacher, 65(6), 22–26. Retrieved from http://lsc-net. terc.edu/do/paper/8105/show/use_set-sci_ref.html Loucks-Horsley, S. Harding, C. Arbuckle, M. Murray, L., Dubea, C., & Williams, M. (1987). Continuing to learn: A guidebook for teacher development. Andover, MA: Regional Laboratory for Educational Improvements of the Northeast and Islands. Loucks-Horsley, S., & Hergert, L. F. (1985). An action guide to school improvement. Alexandria, VA: Association for Supervision and Curriculum Development. Loucks-Horsley, S. Kapitan, R., & Carlson, M. (1990). Elementary school science for the 90s. Alexandria,VA: Association for Supervision and Curriculum Development. Loucks-Horsley, S., Stiles, K. E., Mundry, S., Love, N., & Hewson, P. W. (2010). Designing professional development for teachers of science and mathematics (3rd ed.) Thousand Oaks, CA: Corwin. Mundry, S., & Dunne, K. (2003). Teachers as learners: Professional development in science and mathematics facilitator guide. Thousand Oaks, CA: Corwin Press.

Susan Loucks-Horsley, PhD    175 Mundry, S., & Loucks-Horsley, S. (1999). Designing professional development for science and mathematics teachers: Decision points and dilemmas. NISE Brief. Madison, WI: University of Wisconsin-Madison. National Center for Improving Science Education. (1989). Getting started in science: A blueprint for elementary science education. Washington, DC: Author. National Research Council (NRC). (1996). National Science Education Standards. Washington, DC: National Academy Press. Porter, A. (2000, September). Unpublished paper presented at an AERA Symposium. Regional Educational Laboratories. (1995). Facilitating systemic change in science and mathematics education: A toolkit for professional developers. San Francisco, CA: Author. Sparks, D. (2000, September). Unpublished paper presented at an AERA Symposium. Sparks, D., & Loucks-Horsley, S. (1989). Five models of staff development. Journal of Staff Development, 10(4), 40–57. Weiss, I. R. (1994). A profile of science and mathematics education in the United States, 1993. Chapel Hill, NC: Horizon Research, Inc.

176    S. MUNDRY

Chapter 9

Robert Karplus (1927–1990) Science Education Pioneer Robert G. Fuller University of Nebraska Lincoln, Emeritus Beverly Karplus Hartline Montana Tech

Robert Karplus placed a toy truck in front of a child. He rolled the truck slowly across the desk. “Did the truck move?” he asked. “No,” replied the child. (It is difficult to learn about motion when an object that goes from one point to another does not move. Perhaps the child had misunderstood.) Karplus moved the truck back to its starting position. Again, he slowly rolled the toy truck across the desk to a new location. “Did the truck move?” he asked again. “No,” the child replied once again. “Can you explain to me why you say the truck did not move?” Karplus asked.1 Going Back for Our Future, pages 177–198 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

177

178   R. G. FULLER and B. K. HARTLINE

“It did not move,” responded the child triumphantly. “You moved it!” This surprise answer, early in his explorations of science education, triggered Karplus’ curiosity and hooked his interest. How do children learn? And how can teachers and scientists help them learn science and inquiry better? Before Science Education—Theoretical Physics Robert Karplus was born in 1927 in Vienna, Austria into a successful and highly educated family of Karpluses and Goldsterns. Bob and his younger brother, Martin, exhibited significant intellectual ability at an early age, and both went on to notable scientific careers. At the age of nine, Bob was playing with friends and fell about 40 feet out of a tree he had climbed. This accident fractured his skull, left him unconscious for several days, made him deaf in his right ear, and resulted in subtle balance problems. His recovery kept him out of school for about six months to rest and relearn things, like talking. He could not stand the enforced naps, so he took a flashlight under the covers and read while he was to be sleeping. One can only imagine the stress on secular Jewish families, like the Karpluses, in Austria in the 1930s. Excellent teachers who happened to be Jewish were dismissed from their positions. Former playmates suddenly spurned Bob and Martin. The boys had been studying English, tutored by their aunt, who had lived for a short time in England, and were already more fluent than their parents. A few days after the Nazis annexed Austria in March 1938, the boys left Vienna early with their mother, Isabella Lucie Karplus, on a planned ski vacation to Switzerland, not to return until well after the war. Their father, Hans Karplus, stayed longer in Vienna to take care of some matters but was jailed by the Nazis. Meanwhile, Bob and Martin started school in Zurich, finishing the term. Then Lucie and the boys went to Le Havre, France, for the summer, while she continued trying to obtain a visa for the family to emigrate to England or America. Their father was released from jail in exchange for signing ownership of their property to the Nazis, just in time to join the family to sail for New York, as refugees on the steamship Ile de France. The immigration visa had been arranged by Bob’s uncle, Eduard Karplus, an electrical engineer who had been in the Boston area working for the General Radio Company since 1930. The company president, Melville Eastham, had signed the required affidavit to be responsible for the family. Upon landing in New York in the fall of 1938, they continued to the Boston area, settling initially in Allston, where Bob enrolled in the Boston Latin School, an excellent college-preparatory public school. Soon the family moved to

Robert Karplus (1927–1990)     179

Newton, Massachusetts so the boys could attend Newton’s highly regarded public schools. After graduating as valedictorian from Newton High School in 1943, Karplus entered Harvard to major in chemistry. He was able to take advantage of the accelerated curriculum Harvard offered during World War II, along with a very low military-draft priority (due to the deaf ear and balance problems), to earn his bachelor’s and master’s degrees in three years and complete his PhD in 1948 when he was just 21 years old! His dissertation research in chemical physics was directed by E. Bright Wilson, Jr., who later remarked that Karplus had been his best PhD student. The thesis combined experiments on microwave spectroscopy with theoretical work directed by Julian Schwinger.2 After graduate school, Karplus received a Jewett Fellowship ($3,000 per year) and started his career in what we would call today a two-year post-doctoral position at the Institute for Advanced Study in Princeton, NJ, where J. Robert Oppenheimer was director, having returned there from his stint at Los Alamos with the Manhattan Project. When Karplus arrived in Princeton, quantum electrodynamics (QED) was one of the hottest topics in theoretical physics. Freeman Dyson, also at the institute, had proposed a systematic scheme to calculate the predictions of QED theory. Karplus and another theorist, Norman Kroll, decided to use Dyson’s method to calculate an actual physical observable, where QED would give a result that would be measurably different from the results of classical physics. One can only imagine their excitement. If they succeeded, they would have a result that could be compared with experiment and would provide a test of the validity of the QED theory and its usefulness in predicting the behavior of a physical system. Karplus and Kroll worked by hand for about a year, carrying out very difficult mathematical calculations. Their collaboration produced a famous paper, published in the 1950 Physical Review (Karplus & Kroll, 1950). Their results were in agreement with the experimental results. It was a triumph, not only for QED, but also for Karplus and Kroll. In the early 1950s, Karplus published several other important papers, such as one on the scattering of light by light (Karplus & Neuman, 1951) and papers on positronium (Karplus & Klein, 1952a, 1952b, 1952c). In the summer of 1950, Karplus returned to Harvard as an assistant professor. He continued to do research in theoretical physics as well as teach upper-level physics courses. Since Harvard was unlikely to give tenure to an alumnus, Karplus began to look for a position elsewhere. In March, 1953, during a visit to Harvard, Edward Teller, professor of physics at the University of California, Berkeley, had a long interview with Karplus and was quite impressed. Karplus was offered and accepted a position for fall 1953 as a visiting lecturer at Berkeley with rank equivalent to associate professor. He took a one-term sabbatical from Harvard for this period.

180   R. G. FULLER and B. K. HARTLINE

At the end of the semester, the senior physicists at Berkeley began to urge a permanent faculty appointment for Karplus. They wrote such things as “I am most enthusiastic about Karplus as a theoretical physicist, as a teacher and as a man . . . He has done something to the atmosphere of the place that makes it very reminiscent of the pre-war days when Robert Oppenheimer was at his best” (Alvarez, 1953). Internationally famous physicists Hans Bethe and J. Robert Oppenheimer also weighed in. “In my opinion, Dr. Karplus is one of the best, possibly the best, theoretical physicist of his generation” (Bethe, 1953). “I now write this to you in confidence that this appointment will be a source of pride and pleasure to you and to your colleagues” (Oppenheimer, 1954). Spurred by such glowing letters of support, Berkeley offered Karplus a tenured appointment as associate professor of physics in the summer of 1954. He was only 27 years old. He did not disappoint the Berkeley Physics Department. He began to attract students, and he published with several collaborators on a wide range of topics, including the anomalous Hall effect in ferromagnetic materials (Karplus & Luttinger, 1954), the first application of dispersion relations to particles other than protons (Karplus & Ruderman, 1955), and the drift of particles in the earth’s magnetic field (Hamlin, Karplus, Vik, & Watson, 1961). Karplus became highly respected as a teacher: He is unusually interested in teaching and I have heard numerous favorable accounts of his classroom work. I have sat in several of his classes . . . and I thought his approach was clear, challenging and quite original. He also has a strong experimental and chemical background for a theoretical physicist, thus giving a satisfying breadth to his teaching. (Kittel, 1957)

Teller wrote to “particularly emphasize the contributions Dr. Karplus has made to building up our Physics Department. . . . Due greatly to his initiative we can now boast of a theoretical physics staff which is second to no other one in the United States” (Teller, 1957, n.p.). With these recommendations, in summer 1958 Karplus was promoted to full professor of physics at Berkeley, not quite 10 years after completing his PhD. His contributions to theoretical physics continued to be numerous and exceptional. By 1967, he had published 49 refereed papers in physics—over two per year. More than 90% of his 32 co-authors became fellows of the American Physical Society, and two were Nobel prizewinners in physics. With a buoyant, outgoing personality, Karplus had not buried himself completely in his studies and research, but had spent many evenings in graduate school folk dancing. In fall 1947, while still at Harvard, at one of these dances he had met Elizabeth Frazier, a Wellesley graduate student in physics, with a bachelor’s degree from Oberlin. Bob and Betty were married in December 1948 and honeymooned in New York City. They stayed in the Boston area during January, while Betty finished her master’s degree. When

Robert Karplus (1927–1990)     181

they arrived in Princeton, Betty got a job as the head of the radiochemistry lab in Princeton University’s physics department. During their time in Princeton, she also worked some evenings, doing extensive calculations using a mechanical Marchant calculator for John von Neumann. He needed these numerical results to validate whether the computer he was building was doing weather-prediction problems accurately. The couple started their family in 1950 with the birth in Princeton of Beverly, co-author of this chapter. Their second daughter, Margaret, was born in Cambridge, MA in 1952. Their third child, Richard, was born in Berkeley in 1953, while Bob was a visiting lecturer there. They bought property in Orinda3 in 1954, when Bob returned to Berkeley as a tenured associate professor, and built a house. Four more children were added: Barbara in 1955, Andrew in 1957, David in 1960, and Peter in 1962—all born in nearby Oakland, CA. Interactions with his growing and inquisitive children and their friends triggered Karplus’ career transition into science education starting around 1958. This new career was to be every bit as exceptional as, and even more pioneering than, his physics career. Entering Science Education By the late 1950s Karplus had been doing theoretical physics for a decade. He found it lonely and intellectually exhausting, and the excitement may have started to wane. Only a few people in the world were able to understand in detail what he was trying to do. At the same time, in his university teaching, he noticed that many students had serious problems understanding some elementary concepts in physics because of common misunderstandings they brought from earlier experiences or previous science classes. At home, Karplus was fully engaged in being an intellectual mentor of his children, guiding them as they explored and learned about the world. One day, Beverly volunteered her dad to come to her second-grade class to excite her friends about science. Although his simplified physics lecture was a disaster, the kids were very intrigued by the family’s hand-cranked Wimshurst (static electricity) machine that created sparks and shocks! He began to wonder whether students might have fewer misunderstandings and more interest in science in college, if they were taught differently in elementary and secondary school—through experiences and discovery, rather than textbook facts and vocabulary. Karplus began his journey of discovery by reading the works of Lev Vygotsky, Jean Piaget, and other pioneers in learning and intellectual development, and by watching and interacting with his own children and their friends to validate the readings through experience.

182   R. G. FULLER and B. K. HARTLINE

As Karplus himself summarized in 1977: Little did I realize in 1958, when I visited my daughter Beverly’s second-grade class to teach a few science lessons, that my professional career would be redirected from theoretical physics to science education. My fate was sealed when the National Science Foundation’s very first grant for course content improvement in elementary school science was awarded to three colleagues and me at the University of California, Berkeley, in 1959. At that time there began an intensive learning experience that introduced me to a new field and one that is still continuing . . . (in Thier, 2002, p. 322)

After a transition period that extended from 1959 to 1963, Karplus worked almost exclusively in science education, making contributions from elementary school to the university level. The early 1960s saw the birth of the Science Curriculum Improvement Study (SCIS), co-led with Herbert Thier, a science teacher and school administrator Karplus recruited in 1963 to be the project’s deputy director. His university physics teaching also received attention. He authored an introductory textbook, Introductory Physics: A Model Approach (Karplus, 1969a) for non-science majors, and he used self-paced instruction in his courses for physics undergraduates. At the graduate level, he collaborated with Berkeley faculty in other departments to develop the Graduate Group in Science and Mathematics Education, in which Berkeley PhD students could conduct their dissertation research on an education-related project in their discipline. In addition, he led a pioneering effort at Berkeley in the late 1970s and early 1980s to recruit more women and minority faculty to all departments. This effort was premised on a belief that “Berkeley’s continued greatness as a university depends on the faculty becoming more diversified by including more women and more members of ethnic minority groups than is the case at present” (University of California Berkeley, Berkeley Committee on the Status of Women and Ethnic Minorities, 1981). Karplus was one of the first scientists to apply the inquiry approach that had been so successful in his own research to the challenges of teaching and learning science. Already in the 1960s he had shifted the paradigm to focus on student learning—not just of content facts, but of inquiry and reasoning skills as well. As a scientist he thought of the problem as a system with many interacting components and forces that needed to be explored, discovered and understood. Educating children, Karplus found, offers intellectual challenges that are at least as great as those facing physicists.

Robert Karplus (1927–1990)     183

Karplus’ Enduring Contributions to Science Education We feature in this profile five of the major contributions to science education pioneered by Karplus. For more detail about each of these and his many other contributions, we refer the reader to A Love of Discovery: Science Education—The Second Career of Robert Karplus (Fuller, 2002), edited by coauthor of this chapter, Robert Fuller. Karplus’ many papers in science education are compiled in that volume for easy access and use. 1. Intellectual development. He endeavored to understand the fundamentals of learning and intellectual development, integrating psychological and learning theory with observation and practice to advance both. 2. The learning cycle. As he and his colleagues, including teachers, struggled, they realized it was critical to create a systematic framework—the learning cycle—that would enable classroom teachers to help their students learn and enjoy science. 3. Feedback loops to guide curriculum development. Karplus employed multiple levels of interacting feedback loops to guide curriculum development and learning, a process demonstrated distinctively in the SCIS project. 4. Teacher training. He realized early the importance of augmenting the SCIS elementary science materials with a systematic teaching process. His model for teacher training fostered teachers’ enjoyment of science while helping them become skilled and confident with the curriculum and the effective techniques they needed to help all students—not just the elite—learn science. 5. Autonomy and input. In a theme applicable to learners of all ages— from toddlers to university levels and beyond—Karplus recognized the importance to successful instruction of a combination of student autonomy with input from the teacher. Intellectual Development Karplus focused early on the intellectual development of individual students as a key factor in science education and one that could explain the differential progress of different students, even in the same class. Jean Piaget’s writings, based on Piaget’s extensive interviews of children of various ages as they confronted and grappled with various puzzles Piaget set for them, helped clarify and advance Karplus’ understanding of the development of reasoning. Intellectual development became a central theme that

184   R. G. FULLER and B. K. HARTLINE

integrated and informed the science education research, teaching, and teacher-training efforts Karplus led. To study students’ reasoning, Karplus and his colleagues replicated Piaget’s interview approach. In addition, they expanded the approach to include more children by using written puzzles, which could be administered to an entire class. In the written approach, the puzzle problem is described on a sheet of paper. The student writes his or her solution on the same piece of paper, along with an explanation describing how he or she arrived at the answer. Each puzzle was designed to focus on a specific reasoning skill that is important to scientific inquiry. Such skills include proportional reasoning, probabilistic reasoning, correlational reasoning, logic, and the control of variables. Karplus was particularly interested in finding out when and how these components of formal reasoning emerged in children of different ages. Like in Piaget’s interviews (and in the truck example at the beginning of this profile), the child’s explanation was key to understanding what and how he or she was thinking and essential for classifying reasoning strategies. Sometimes the process of explaining, both in the interview and for the written puzzles, caused a student to rethink and revise the answer. Initially Karplus worked with children in the United States, and later he and colleagues expanded the study to include students from seven countries (Karplus, Karplus, Formisano, & Paulsen, 1977). To analyze the students’ responses to the written puzzles, he used the technique of having multiple readers review and categorize each solution and explanation, then discuss their categorization to arrive at a consensus classification of the student’s reasoning. In this way, the research team self-calibrated and could minimize distortions from reader bias affecting the conclusions. Today qualitative methodologies similar to those pioneered by Piaget and extended by Karplus are mainstays of science education research. With a deep knowledge-base of the work of Jean Piaget on the development of thought among young adolescents, coupled with a strong desire and commitment to help understand and improve students’ reasoning in science and mathematics, Robert Karplus and his colleagues believed both in the merits of qualitative-based research and in the significance of including classroom teachers in the process of research to improve students’ learning. In a paper presented in 1980, Elizabeth Karplus described the significance and the challenges of this avant-garde action-research perspective. She said, “For the practicing teacher, every classroom can be a research laboratory in which new data about individual students’ learning . . . [must] be analyzed daily. This research must be carried out under conditions that would stun a scientist or a mathematician: the vast number of variables, the lack of controls, the many subjects to be observed at the same time, the variations among the observers’ knowledge and sensitivities, are combined with the ‘noise’ of random careless

Robert Karplus (1927–1990)     185 errors and slips . . . that occur frequently.” To help teachers observe and assess their students’ reasoning, and to help teachers learn how to act upon their observations in order to enhance their students’ mathematical and scientific reasoning, R. Karplus and his colleagues (1977) developed and conducted a sequence of five inservice teacher-development programs titled, “Science Teaching and the Development of Reasoning.” (Khoury, 2002, p. 98)

The Learning Cycle With funding from the pioneering 1959 NSF grant, which was a precursor to SCIS, Karplus and his colleagues prepared and taught three elementary science units, titled “Coordinates,” “Force,” and “What Am I?” Although the experience was interesting and fun, analysis of what students had learned revealed serious misconceptions and other weaknesses. Faced with this evidence, Karplus raised a key question: “How can we create a learning experience that achieves a secure connection between the pupil’s intuitive attitudes and the concepts of the modern scientific point of view?” (Lawson, 2002, p. 53). During the spring of 1960, using twice-weekly science lessons presented in his three oldest children’s first, second, and fourth grade classrooms, Karplus started to become familiar with the points of view children at these ages have about natural phenomena. He also began to formulate preliminary answers to his question and test them against his classroom experiences. The following year, during a physics sabbatical in Europe, he visited Piaget’s research institute in Geneva and gained from the Swiss psychologist what would prove to be key insights about learning. By the fall of 1961, when Karplus returned to Berkeley, he had decided to try a novel instructional approach. The students would gain hands-on experience with and make their own observations of a phenomenon. Then he, as teacher, would help them interpret their observations in a more analytical way than they could without assistance. J. Myron Atkin, then a professor of education at the University of Illinois and one of the researchers on an early NSF grant to improve high-school science instruction, visited Berkeley to share his views on teaching with Karplus. Together, Atkin and Karplus formulated a method of “guided discovery,” which was implemented in subsequent trial lessons (Atkin & Karplus, 1962). The guided discovery method mimicked the way scientists invent, use, and refine new concepts to explain nature. In their 1962 paper, Atkin and Karplus wrote about how early humans observed the motions of the sun and planets, and used these observations to develop and refine an understanding of the solar system. The first explanations were mythological. Later, humans made the conceptual “invention” of placing earth at the center

186   R. G. FULLER and B. K. HARTLINE

of the solar system. The heliocentric theory was an even later invention. People then attempted to “discover” other phenomena that could be understood using the invention. If successful, these attempts helped reinforce and refine the invention. Unsuccessful attempts revealed limits of the invention or, sometimes, led to its replacement. During this time, Karplus resumed discussions he had started over 10 years earlier with Berkeley faculty member, Thomas Kuhn,4 when they were both students and junior faculty at Harvard. Kuhn was a physicist, whose career had transitioned into the history and philosophy of science, and he shared with Karplus the pre-publication manuscript of his famous book, The Structure of Scientific Revolutions, published in 1962. Kuhn’s ideas on scientific thought provided a strong foundation for the discovery and invention components of the learning cycle and Atkin’s and Karplus’ 1962 paper. Atkin and Karplus distinguished between the teacher’s introduction of a new concept (the invention phase) and its subsequent test and extension by the students (which they called the discovery phase). Karplus and Atkin assumed that children are not generally capable of inventing (or constructing) the modern concepts of science. Therefore, the teacher must help the children by explaining the concepts, connecting the explanation with the students’ previous observations and experience, and helping them refine their interpretations of these observations, if necessary. After the “invention,” the teacher gives the children opportunities to discover other observations or phenomena that can be interpreted using the concepts. The “discovery” experiences allow students to extend and apply their new understanding, and also achieve the “equilibration” and self-understanding described theoretically by Piaget. Karplus accepted an invitation to work with the Elementary Science Study in Cambridge, MA during the summer of 1962. There he realized that children needed first to explore an experimental system at their own pace, making observations—sometimes inconsistent with what they expect—and wrestling with their own preconceptions. Only after such an unscripted “exploration” does it make sense for the teacher to bring the explanation or “invention” to their attention. Armed with this new insight, Karplus went back into the classroom in public schools near the University of Maryland, where the SCIS program was temporarily headquartered during the 1962–1963 school year. He tried with considerable success the three-stage sequence of preliminary exploration, invention, and discovery. Note that the invention of “learning cycle” as a name first appeared in SCIS Teacher’s Guides around 1970. Without the name, the conceptual foundation was described clearly in the 1967 book, A New Look at Elementary School Science, co-authored with Thier. As Karplus was going through multiple cycles of exploration, invention, and discovery regarding how students learn science and teachers can best teach them, Chester Lawson, a geneticist at Michigan State

Robert Karplus (1927–1990)     187

University, was doing similar work aimed at improving high school biology instruction. In his biology sourcebook, Lawson (1958) identified a thought pattern he described as “Belief—Expectation—Test,” which is important to learning. This pattern essentially matches Karplus’ and Atkin’s invention and discovery. Thus, the same pattern of instruction had been independently “invented” by Atkin and Karplus and by Lawson. When Karplus, the physicist, needed a biologist to assist in developing the life science half of SCIS, he called on Lawson. A two-week consultation in the summer of 1965 culminated with Lawson spending ten years as director of life sciences for SCIS. The learning-cycle concept of exploration, invention, and discovery continues to be foundational in science education, guiding curriculum development, work with science teachers, and the hands-on exhibits at discovery centers, such as Berkeley’s Lawrence Hall of Science (LHS). Many people have built on Karplus’ learning cycle and have modified it to meet a wide range of needs. Among other changes, the names of its phases have evolved. Already in the mid-1970s, Karplus and his colleagues noticed that teachers were having difficulty understanding what the terms “invention” and “discovery” meant. In a series of 1977 papers Karplus changed the names of those phases to “concept introduction” and “concept application” (e.g., Karplus, Karplus, et al., 1977). The Biological Sciences Curriculum Study (BSCS, 1992) version of learning cycle-based instruction has “Five E” phases: engage, explore, explain, elaborate, and evaluate. Essentially Karplus’ original exploration phase encompasses both “engage” and “explore,” while “explain” maps onto concept introduction, and “elaborate” maps onto concept application. “Evaluation,” which includes observing students and assessing their knowledge or skills, had an implicit though unmentioned ubiquitous presence in the Karplus learning cycle. Karplus and colleagues worked very hard to align how students were evaluated with the SCIS curricular goals and learning cycle and to emphasize self evaluation by the students (Piaget’s equilibration stage). A Research-Based Curriculum-Development Feedback Loop and the Science Curriculum Improvement Study By 1969, after several years in science education and with SCIS well advanced, Karplus was still convinced “that there is no satisfactory theory of instruction or of learning which leads unambiguously to a teaching-learning experience, once an educational objective has been specified” (Karplus, 1969b). Since entering science education, he had tackled the challenges of curriculum development using the same systematic scientific approach that he had used to advance physics in areas where knowledge was non-existent

188   R. G. FULLER and B. K. HARTLINE

to sketchy and scientific theories were lacking. He performed exploratory experiments and made observations of teaching and learning to try to find potentially predictable relationships. He read widely to explore the foundational and latest research and theories about the psychology, sociology, and neuroscience of learning, and to try to make sense of his experiments and observations. And he collaborated with others who brought complementary expertise and perspectives, especially teachers. For SCIS, Karplus assembled and mentored a large and highly diverse team, including over 50 scientists, teachers, science educators, educational psychologists, equipment designers and inventors, illustrators, typists, film makers, and business staff—each of whom was valued for his or her special expertise. The main team, headquartered at Berkeley’s then-new LHS, took each SCIS activity and module through a series of rigorous development hurdles and classroom trials in local schools. The project also included a carefully selected network of five geographically distributed, universitybased, semi-autonomous Trial Sites, where the curriculum modules could be tested extensively in nearby urban, suburban, and rural schools by real teachers, whose classrooms sampled the diversity of the nation. This network comprised Teachers College in New York City, Michigan State University, University of Oklahoma at Norman, University of California at Los Angeles, and University of Hawaii at Honolulu. Professors at each of those campuses directed the Trial Site and served on the SCIS coordinating team. Karplus and the SCIS team cycled through a feedback-loop process to develop → field test → revise → re-test activities, modules, materials, and teacher’s guides—collecting and digesting mountains of feedback in the process. The field testing was done in real classrooms, and, as the development progressed, increasingly with regular teachers (after some specialized training, described in the next section) in the diverse range of settings, geographic locations, and communities encompassed by the network of Trial Sites. Lessons, materials, and teaching approaches had to excel against every one of the following criteria: “scientifically accurate, educationally relevant to schools and society, and intrinsically interesting and understandable to children in the classroom. In this latter case the ‘gold standard’ was that children had to be able to explain a key concept in their own words” (Peterson, 2002, p. 14). Karplus’ experimental approach to curriculum development was designed to produce a new science curriculum, whose success in the classroom could be predicted based on the extensive experimentation and field testing. This approach relied on three major elements: 1) Combining the expertise of scientists, teachers and science educators in the creation of a curriculum, 2) A formal process for the initial testing of ideas for the new science curriculum, and 3) Extensive replication under the various conditions in which the curriculum was to be used, that is, by conduct-

Robert Karplus (1927–1990)     189 ing field experimentation using SCIS Trial Editions at nationally distributed Trial Sites. (Peterson, 2002, p. 13)

1. Combining Expertise. The model involved pooling expertise by forming a dedicated authoring team for each unit. Teams usually consisted of a scientist with expertise in the topic, a science educator, and an elementary school teacher. They collaborated on the development of SCIS activities, using actual students in the classroom to try curriculum ideas. When authoring teams tried their activities in the Berkeley classrooms, one member of the team introduced the science activity, while a second team member took extensive notes, and the third team member walked among the students along with the regular classroom teacher, watching and listening to children’s reactions and responses to activities. After introducing a science activity in the classrooms of two or three Berkeley-area schools, the team returned to LHS to pool their evidence of each activity’s successes and failures. Observations provided clear and sometimes disappointing evidence of children’s responses to activities and materials. Children’s written responses and drawings were examined for evidence of their understanding. Direct observations of children brought clarity to how some activities could be improved. Revision and re-teaching of each candidate SCIS activity in local elementary school classrooms continued until a decision was made to include it in a Trial Edition, or to “archive it.” The SCIS Archive was a vaultlike room at the LHS, containing shelves of boxes and binders documenting SCIS operations, including a huge collection of every potentially good idea that never made it to publication. 2. Testing Initial Ideas About Science Curriculum. Karplus had established a rigorous experimental environment and high expectations for the SCIS curriculum. This approach meant that activities had to be tried and revised many times before they achieved the multiple criteria that allowed publication. [E]very idea about the choice of science content, the language to be used with the teacher and the students, the instructional materials, and teaching methods, would be tested experimentally—in the reality of the classroom. The decision-rule for whether a single science lesson or a complete unit would “go to publication” was based upon multiple sources of replicable evidence, reviewed by teams of experts who produced the curriculum, and through extensive field testing with children and by teachers and school administrators who represented the potential consumers of SCIS. (Peterson, 2002, p. 14)

3. Replication: SCIS Trial Editions and Trial Sites. When units were considered successful enough for sending to the Trial Sites, they were formatted into very professional-looking sets of teacher’s guides, student booklets (one per student), and classroom kits of hands-on materials. At summer

190   R. G. FULLER and B. K. HARTLINE

workshops, teachers from elementary schools partnering with each Trial Site were introduced to the SCIS program and philosophy. They worked with the kits, like their students would, and learned the science behind the activities. Karplus met personally with many of these groups of teachers and encouraged them to give honest feedback about the effectiveness of every science activity in each unit. This feedback and review of the completed student booklets led to further revisions and refinements in activities and units that ultimately made it to the final edition of Teacher Guides, accompanying Student Booklets, and durable, manufacturable kits for the exploration and discovery activities. Compared with the prevailing model for writing science textbooks, Karplus’ approach to curriculum development featured several major innovations. SCIS used the exploration/invention/discovery learning cycle to guide the design of materials, lessons, and curriculum. Karplus and scientist colleagues identified the few fundamental ideas and concepts children needed to understand to begin to think like scientists. They selected and sequenced the topics and activities in partnership with teachers, science educators, and educational psychologists to both match and advance the children’s intellectual development. They included teachers as full partners in the development and implementation, providing extensive training and resources. And they used extensive data collected from and by observing students and teachers in the classroom. “This methodology yielded a set of products that was to ultimately transform the landscape of elementary science education and classroom teaching” (Bowyer, 2002, p. 234). By the mid-1970s SCIS had produced and turned over to commercial production an integrated and coherent K–6 life-science and physical-science curriculum (Table 9.1) based on learning cycles. These units increased in conceptual difficulty from grade to grade, and systematically prepared students for the explorations, inventions, and discovery in the following unit. In addition to the efforts of Karplus, Thier, Peterson, and Chester Lawson, key staff members Anton Lawson, Jack Fishleder, Robert Knott, Carl Berger, Table 9.1  SCIS Units Grade Life Science K 1 2 3 4 5 6

Beginnings Organisms Life Cycles Populations Environments Communities Ecosystems

Physical Science

Material Objects Interaction and Systems Subsystems and Variables Relative Position and Motion Energy Sources Scientific Theories

Robert Karplus (1927–1990)     191

and Marshall Montgomery made substantial contributions. Beyond Berkeley, Mary Budd Rowe, Stanford Davis, John Renner, Albert Carr, and Glenn Berkheimer were critical to the development and dissemination effort as coordinators of the Trial Sites. Topics, themes, and approaches from SCIS have been emulated and adapted by essentially all of the hands-on science curricula that have emerged in the subsequent decades. Moreover, SCIS-based classroom kits continue to evolve and are available commercially and in classroom use in 2012. A Paradigm Shift in Elementary School Teacher Development in Science In my opinion, of all the challenges Karplus faced in his combined roles as science curriculum researcher/developer, the creation of a major paradigm shift in the teacher’s role was the most important and the most difficult. And, the stakes were high! If Karplus was not successful in identifying and addressing the major issues involved in science teacher development, the goal of achieving scientific literacy for all students would fail. In retrospect, it proved to be easier for students to learn the major concepts and processes of science using the SCIS curriculum (when it was optimally taught) than it was for teachers, particularly elementary teachers, to make the transitions necessary to teach this material. The reasons for this are numerous and important to understand if we are to fully appreciate the scope and enormity of Karplus’ work. (Bowyer, 2002, pp. 230–231)

In the 1960s, teaching science to elementary school children starting at age five or six was a radical idea. Parents, teachers, and school administrators believed that such young children needed to concentrate on learning reading, writing, and arithmetic. In addition, elementary teachers had extremely limited science content knowledge and little familiarity with the nature of science or scientific inquiry. Moreover, they were unaware of the research on the development of reasoning. Their college courses on pedagogy and the psychology of learning that they referenced did not emphasize hands-on approaches or the learning cycle. Finally, Karplus was an outsider in terms of any experiential or intellectual knowledge of the classroom environment or teacher development. These were some of the challenges he faced in teacher development for SCIS. His challenges were rivaled and perhaps even exceeded by those facing the teachers. To be successful, teachers would have to learn science content, learn the methods of science, and master new pedagogical strategies—all while continuing to teach the children in their schools five days per week.

192   R. G. FULLER and B. K. HARTLINE

With Thier as the deputy director for SCIS, Karplus had an experienced and expert invention partner for teacher development. Their relationship produced bold and visionary solutions to the huge challenges of teacher development. Early in their collaboration, they realized that it was vital for SCIS teachers not only to understand science content and inquiry and how to foster student learning, but also to enjoy science. Only then could they share enthusiasm for science with their students. Karplus used the same learning cycle to learn about teacher development as he and his colleagues had developed for teachers to use in teaching science to their students. In the exploration phase, Karplus went into the classroom and immersed himself in its culture. He questioned everything about classroom phenomena and tirelessly pursued answers. When he observed the teaching of a new candidate SCIS lesson, he would walk over to a group of students and ask thoughtful questions concerning their data or designs. He would listen intently to the students’ answers. He was visibly pained when teachers described frustrations. He was elated, beaming in fact, when a teacher joyfully discovered a student who understood a concept because of a particular experiment he/she had conducted. Karplus was open to the disequilibrated feelings associated with not knowing, being totally puzzled, and desperately wanting to understand. (Bowyer, 2002, pp. 233–234)

In the invention phase Karplus endeavored to find out through reading and discussions what the experts had to say. Questions from his explorations focused this study. Karplus was influenced by the writings of Johann Heinrich Pestalozzi, John Dewey, Jerome Bruner, Vygotsky, and Piaget in his thinking on teacher development. The third phase, discovery, came from applying knowledge gained from both earlier phases. His methodology consisted of teaching, getting feedback, reflecting and revising, and then re-teaching. This approach produced the SCIS teacher-training curriculum based, not surprisingly, on the learning cycle. The teachers were guided through exploration, invention and discovery phases, in which concepts, processes, and pedagogical practices were organized, experienced, and presented, along with their underlying psycho-social and pedagogical basis. At teacher institutes, teachers practiced using the hands-on equipment and materials in the extensive SCIS kits. They learned about the SCIS program and philosophy, and the importance of students using the materials to explore the phenomena and construct new knowledge for themselves. They were also familiarized with the simple but powerful idea of “wait time,” developed by Rowe, who had learned that answers improve and more children participate, if the teacher waits a few seconds, rather than calling on the first child to raise a hand or providing the answer when students are slow to respond.

Robert Karplus (1927–1990)     193

To strengthen teacher development, Karplus and his colleagues created and used numerous films featuring model lessons being taught in the classroom. With Rita Peterson, he made a film of students performing various Piagetian reasoning tasks, for example, formal reasoning patterns, to help teachers realize the importance of matching their teaching to student reasoning abilities. Overall, the development experience was designed to maximize the opportunities for and depth of teacher understanding of science content, inquiry, and effective pedagogy. SCIS Teacher Guides also help to develop the teachers’ knowledge and skills. These booklets provide science background information relating to the science concept, so the teachers would know more than their students are expected to learn. The guides include examples of open-ended questions teachers could use to stimulate their students’ thinking. The lesson plans follow the learning-cycle model, to help the teachers help their students construct knowledge and understanding. Evaluation supplements give teachers techniques and information to gauge their students’ understanding of the concepts. Instructions to teachers, science coordinators, and administrators on how to acquire, distribute, and organize the handson materials are provided. Even after the SCIS curriculum had been completed, Karplus continued to emphasize teacher development. Working with others, he created and offered a series of teacher workshops on science teaching and the development of reasoning. These experiential workshops helped middle school and high school teachers and college faculty explore and understand the work of Piaget and Karplus and the implications for student learning in the science classroom—from elementary school through university levels (Karplus, Lawson, et al., 1977).5 Karplus never stopped thinking about children and teachers. He continually tried to figure out how they might come to a better understanding of the basic scientific constructs. He truly hoped they would retain an appreciation of scientific explorations throughout their lives. (Bowyer, 2002, p. 235)

Autonomy and Input In February 1982, only a few months before the heart attack that left him disabled and ended his career, Karplus exchanged correspondence with Nobel laureate Joshua Lederberg,6 then president of Rockefeller University. Lederberg had written to ask for advice about what it would take to be successful in college biology instruction. Karplus’ reply provides an excellent ending for this chapter.

194   R. G. FULLER and B. K. HARTLINE I feel very strongly that all instruction should combine instructional input from the teacher with opportunities for autonomous activities by the students. . . . The reason I bring up this point is that most teachers, especially individuals who are very expert in their field, tend to concentrate on the input—what they have to offer the students. They tend to overlook the students’ need for autonomy, for pursuing questions or interests arising from within themselves, but inspired by the teacher. My second point would be to look for an instructor who is an expert in his field and who is at least as bright as the students. He or she should be genuinely interested in the development of the students, more so than in displaying his or her ability and knowledge. Third, I would look for a set of related topics in which the students, at whatever level they may be, could engage in some independent study or investigation even though they have very limited specialized knowledge. Some instructors have chosen readings from original source materials in the history of their sciences. Others have used more of a laboratory orientation, with either the students’ homes or a scientific laboratory serving as the setting for investigation. Your letter distinguished between two groups of students, I believe—those who are genuinely interested in their studies of biology and others who are required to study the field, but do not have a genuine interest. It is clearly much more difficult to work with the latter group of students, and it is here that a gifted instructor can be particularly effective. (Karplus, personal correspondence, 1982)

Student autonomy with input from an instructor who is genuinely interested in the development and learning of the students remains a formula for success in science education today. Science For Everyone Karplus never thought of school science as an elite subject, suitable for only the brightest students. In his earliest papers on science education he emphasized science for everyone. Now “science for all” is a mantra for the United States’ national science goals. In the 1960s, Karplus and his coworkers tried to make sure that their science materials were, indeed, for all students. They tried the SCIS lessons with diverse children, including those with disabilities and those without English fluency. Drawings and photographs in student materials and Teacher’s Guides included girls and boys from many racial and ethnic groups—far ahead of the national attention to diversity. Only lessons that worked for all students made it into the final set of SCIS activities and materials.

Robert Karplus (1927–1990)     195

Robert Karplus struck his co-workers as an infectiously happy and caring person. Those who knew and worked with him testify to the immense influence he had on their lives and careers. He loved discovering new things himself. He wanted everyone to know the joy of discovery. “Don’t tell me, let me find out.” Notes 1. Asking students to explain their thinking became one of the stock tools in his education-research toolbox, much like focusing in on puzzling observations and measurements is a standard tool for the research scientist. 2. 1965 Nobel laureate in physics (with Richard Feynmann and Sin-Itiro Tomonaga) for fundamental work in Quantum Electrodynamics (QED). 3. Orinda is a suburb just east of Berkeley. 4. With a PhD in physics (Harvard, 1949), Kuhn was at the time a professor of the history of science at Berkeley, affiliated with both the history and philosophy departments. 5. In 2009 R.G. Fuller, T.C. Campbell, D.I. Dykstra, Jr., & S.M. Stevens published College Teaching and the Development of Reasoning (Information Age Publishing), updating the material from the 1970s and broadening it from science to all subjects, to provide 21st century college faculty with insights about the development of reasoning and the importance of matching teaching strategies and materials to the students’ intellectual development. 6. 1958 Nobel laureate in Physiology and Medicine.

References Alvarez, L. W. (1953). Letter to Berkeley Physics Department Chairman Raymond T. Birge, December, 1953. Atkin, J. M., & Karplus, R. (1962). Discovery or invention? The Science Teacher, 29, 45–47. Bethe, H. A. (1953). Letter to Berkeley Physics Department Chairman Raymond T. Birge, December, 1953. Biological Sciences Curriculum Study (BSCS). (1992). Science and technology: Investigating human dimensions. Dubuque, IA: Kendall/Hunt. Bowyer, J. (2002). Science literacy and the teacher development gap: Karplus’ challenge. In R. G. Fuller (Ed.), A love of discovery: Science education—The second career of Robert Karplus (pp. 230–235). New York, NY: Kluwer Academic/Plenum Publishers. Fuller, R. G. (2002). A love of discovery: Science education—The second career of Robert Karplus. New York, NY: Kluwer Academic/Plenum Publishers. Hamlin, D. A., Karplus, R., Vik, R. E., & Watson, K. M. (1961). Mirror and azimuthal drift frequencies for geomagnetically trapped particles. J. Geophys. Res. (USA), 66(1), 1–4.

196   R. G. FULLER and B. K. HARTLINE Karplus, R. (1969a). Introductory physics: A model approach. New York, NY: W. A. Benjamin, Inc. Karplus, R. (1969b). Some thoughts on science curriculum development. In E. W. Eisner (Ed.), Confronting curriculum reform The Cubberly Conference, Stanford University (pp. 56–61). Boston, MA: Little, Brown and Company. Karplus, R., Karplus, E., Formisano, M., & Paulsen, A. C. (1977). A survey of proportional reasoning and control of variables in seven countries Journal of Research in Science Teaching, 14(5), 411–417. Karplus, R., & Klein. A. (1952a). Electrodynamic displacement of atomic energy levels I. Hyperfine structure. Phys. Rev., 85, 972–984. Karplus, R., & Klein. A. (1952b). Electrodynamic corrections to the fine structure of positronium. Letter in Phys. Rev., 86, 257. Karplus, R., & Klein. A. (1952c). Electrodynamic displacement of atomic energy levels III. The hyperfine structure of positronium. Phys. Rev., 87, 848–858. Karplus, R., & Kroll, N.M. (1950). Fourth-order corrections in quantum electrodynamics and the magnetic moment of the electron. Phys. Rev., 77, 536–549. Karplus, R., Lawson, A. E., Wollman, W. T., Appel, M., Bernoff, R., Howe, A., Rusch, J. J., & Sullivan, F. (1977). Workshop on science teaching and the development of reasoning [Presentation]. Berkeley, CA: Lawrence Hall of Science. Karplus, R., & Luttinger, J. M. (1954). Hall effect in ferromagnetic. Phys. Rev., 95, 1154–1160. Karplus, R., & Neuman, M. (1951). The scattering of light by light. Phys. Rev., 83, 776–784. Karplus, R., & Ruderman, M. A. (1955). Applications of causality to scattering. Phys. Rev., 98, 771. Karplus, R., & Thier, H. (1967). A new look at elementary school science. Chicago, IL: Rand McNally and Co. Khoury, H. A. (2002). Central role of students’ reasoning. In R. G. Fuller (Ed.), A love of discovery: Science education—The second career of Robert Karplus (pp. 97– 99). New York, NY: Kluwer Academic/Plenum Publishers. Kittel, C. (1957). Letter to Berkeley Physics Department Chairman A. Carl Helmholz, October, 1957. Lawson, A. E. (2002). The learning cycle. In R. G. Fuller (Ed.), A love of discovery: Science education—The second career of Robert Karplus (pp. 53–60). New York, NY: Kluwer Academic/Plenum Publishers. Lawson, C. A. (1958). Language, thought and the human mind. East Lansing, MI: Michigan State University Press. Oppenheimer, J. R. (1954). Letter to Berkeley Physics Department Chairman Raymond T. Birge, January, 1954. Peterson, R. W. (2002). A Love of Discovery. In R. G. Fuller (Ed.), A love of discovery: Science education—The second career of Robert Karplus (pp. 7–19). New York, NY: Kluwer Academic/Plenum Publishers. Teller, E. (1957). Letter to Berkeley Physics Department Chairman A. Carl Helmholz, October, 1957. Thier, H. D. (2002). It was a great time . . . In memoriam . . . Let us continue, Robert Karplus, 1927–1990. In R. G. Fuller (Ed.), A love of discovery: Science educa-

Robert Karplus (1927–1990)     197 tion—The second career of Robert Karplus (pp. 320–323). New York, NY: Kluwer Academic/Plenum Publishers. University of California Berkeley, Committee on the Status of Women and Ethnic Minorities. (1981). Workshop on faculty diversification. (Members: Y. George, S. Humphreys, R. Karplus, N. Kreinberg, A. Saragoza, M. Wake).

198   R. G. FULLER and B. K. HARTLINE

Chapter 10

The Nature and Development of Scientific Reasoning My Career in Science Education Anton E. Lawson Arizona State University, Emeritus

Introduction In 1969 I was a graduate student at the University of Oregon studying marine biology. The Vietnam War was still raging, and I was about to lose my student deferment and be drafted into the Army—a most unappealing prospect. Fortunately, deferments were still being granted to school teachers. So after receiving a master’s degree at the end of the fall semester, I applied for and was hired to teach eighth grade mathematics and science at Ralston Intermediate School in Belmont, California. My teacher preparation involved all of one week sitting in the back of a classroom observing an experienced teacher. The teacher was retiring at the end of the week, and I was to take over his classes the following week.

Going Back for Our Future, pages 199–216 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

199

200   A. E. LAWSON

When my week of “teacher training” was over, the teacher handed me a copy of the class textbook and proudly announced that he had covered more than half of the text during the first semester and my job was to cover the rest before the end of the school year. Being inexperienced and naive, I was more than happy to take up the challenge. So I enthusiastically grabbed the textbook and started using it to develop lessons. Unfortunately, it soon became distressingly apparent that my students seldom learned anything of value along the way. And if meaningful learning did occur, it was certainly not because of anything that I had done. Hard as I tried, my lessons failed time and time again. My so-called “teaching” job quickly became one of simply trying to make the little rascals behave. Learning How to Teach While struggling to learn how to teach, my father (Chester Lawson, a biology professor from Michigan State University) and his colleague Robert Karplus (a physics professor at the University of California at Berkeley) were hard at work across the San Francisco Bay at the Lawrence Hall of Science in Berkeley developing a K–6th grade science program called the Science Curriculum Improvement Study (SCIS). Together they, and others at SCIS, crafted a teaching method called the learning cycle. And they were using the learning cycle method to develop a science curriculum (Atkin & Karplus, 1962; SCIS, 1970, 1973). When I told my dad about my teaching frustrations, he was quick to share several of the SCIS lessons and teacher guides. Many of the lessons involved concepts that I was trying to teach, so I was more than happy to give them a try. As you may know, learning cycles start with explorations. Putting interesting materials in students’ hands and asking them to explore was a “no brainer.” Exploration not only made great sense to the scientist in me but to my students as well. Pulling off the next phase of learning cycles (the invention or term introduction phase), however, was more challenging. In spite of my lack of teaching experience, I knew quite a lot about the science involved in the lessons. So I didn’t pay much attention to the teacher guides. This turned out to be a big mistake. For some reason, unknown to me at the time, I simply could not get the students thinking along the lines that I wanted. I would ask question after question and mostly get blank stares. So one day, out of sheer frustration, I put my ego aside and said to myself: “OK. No matter what I say, I can’t make these lessons work. So I will ask questions exactly when the teacher guide tells me to and I will phrase them exactly how they are phrased in the guide.” To this day, I can still recall going to the guide and writing the guide questions on a small piece of paper, which I then concealed in my hand. Then when the time came, I glanced at

The Nature and Development of Scientific Reasoning    201

the paper and read off the questions just as they were written in the guide (e.g., “What happened to the four grass setups? What do you think now about what light does for plants? Yes, light is important for plants. So how might we explain why the plants grew in the dark for a while?” And so on.). When I did this, to my utter amazement, something wonderful happened. The students and the classroom suddenly became alive! The students were tossing out all sorts of ideas. And just like real scientists, they began debating those ideas in terms of the evidence and their past experiences. At last, I had discovered the proper bait and had hooked my first big fishes! Needless to say, I was also hooked—hooked by the knowledge that SCIS was really on to something and that teaching could be fun and engaging. Clearly, I was never going back to my initial way of teaching. My next task was trying to figure out what SCIS with its learning cycle approach knew about teaching that I didn’t. Back to Graduate School: Testing Piaget’s Theory After a few years of middle school teaching, the Vietnam War was ending, and I was able to go back to graduate school. But instead of reenrolling in a PhD program in marine biology, my teaching experiences convinced me that a greater intellectual challenge lay in trying to figuring out what makes students, rather than marine organisms, tick. I might add that before going back to graduate school, my father advised me to learn everything I could about Jean Piaget’s theory of intellectual development with the primary goal of finding out what was wrong with it! Much of the SCIS approach was based on Piaget’s work. By the way, my father not only provided me with this particular challenge, he provided similar challenges while I was growing up. For example, when I was about eight years old, I recall riding in the back seat of our car on a family trip to Pennsylvania (I grew up in Michigan). The sky was full of clouds, some brilliant white, others dark, almost black. So I asked my father why the clouds were different “colors.” His reply was typical: “Good question Tony—what ideas do you have?” I do not recall the rest of the conversation except to say that I am certain that he did not tell me the answer, which I am sure he knew. I suspect that growing up on a steady diet of these sorts of exchanges taught me to enjoy thinking about, and trying to explain things. This lesson was reinforced several times during my PhD program by John (Jack) Renner, my doctoral committee chair at the University of Oklahoma. Jack liked to say that giving students the right answers stops, rather than starts, thinking.

202   A. E. LAWSON

My doctoral research tested an important prediction from Piaget’s theory. I learned that the theory concerned the development of “how to” knowledge (procedural knowledge) and its importance in the acquisition of “know that” knowledge (declarative knowledge). For example, one needs to know how to count to know that there are ten marbles on the table. In biology, one needs to know how to sort, classify and seriate to know that species diversity increases from the poles to the equator. And one needs to know how to test theories to know that evolution has occurred as opposed to special creation. At the time, data suggested that many high school students lacked the procedural knowledge presumably needed to acquire much of the declarative knowledge that teachers typically try to teach. In Piagetian terms, many high school students had not yet developed the formal stage reasoning patterns presumably needed to understand the more complex and abstract concepts that teachers were trying to teach. I coined the term “formal concepts” to refer to those concepts and defined them as concepts derived from the postulation of unseen theoretical entities such as atoms, photons, and genes. Formal concepts, or theoretical concepts as I have also referred to them (e.g., Lawson & Karplus, 1977), played, and still play, a prominent role in the high school science curriculum. In short, my results supported this aspect of Piaget’s theory, as a strong correlation was found between students’ developmental stages and their concept understanding. Further, none of the “concrete operational” students demonstrated any understanding of formal concepts (Lawson & Renner, 1975). Thus, if teachers want their students to truly understand what they are trying to teach (as opposed to merely memorizing words), it would seem that the development of procedural knowledge (i.e., reasoning patterns) needs to become a primary goal. Importantly, in Piaget’s theory, intellectual development occurs as a consequence of both physical and social experience, neurological maturation, and an internal, largely subconscious, mental process called self-regulation. Thus, an early goal of mine was to better understand how these factors operate and the role teachers can, or cannot, play in their students’ intellectual development. An early focus was on self-regulation, which presumably occurs when prior ideas and/or behaviors are contradicted. In theory, contradictions not only lead to new ideas and new behaviors, but also to improved reasoning. Thus, a related theoretical implication is that the really important aspects of science and mathematics literacy, such as effective reasoning and problem-solving skills, cannot be directly taught. Instead, they are the products of self-regulation. Interestingly, although people generally know if and when they have learned a specific piece of declarative knowledge, they seldom know if and when their procedural knowledge developed. This means people who lack higher-order reasoning patterns do not

The Nature and Development of Scientific Reasoning    203

realize their deficiencies, while people who have developed such reasoning patterns assume incorrectly that everyone else has developed them as well! Not surprisingly, a number of problems result, not the least is that, just as I had done on my first teaching job, many teachers, administrators, test developers and policy makers ignore procedural knowledge and focus solely on teaching and testing declarative knowledge. However, because the pace of intellectual development lags in so many students, a huge portion of what we try to teach junior high and high school students (and even many college students) is missing the mark. Instead, it simply “goes in one ear and out the other.” The good news is that once the problem is understood, there is a lot that teachers can do to help students develop better reasoning patterns and construct understanding of centrally important scientific concepts and theories—but more about that later. Post-doctoral Research After finishing my PhD in 1973, I spent a year as a post-doctoral researcher at Purdue University and then moved to the University of California at Berkeley to work with Robert Karplus at the Lawrence Hall of Science on several research projects in student reasoning and intellectual development. For those who may be unfamiliar with Karplus’ many contributions to science education, I strongly recommend A Love of Discovery: Science Education—The Second Career of Robert Karplus (Fuller, 2002), which contains a collection of his works. My father was still working at the Lawrence Hall of Science when I arrived there in 1974, and we happily spent many lunch hours on a hillside overlooking the San Francisco Bay eating sandwiches and discussing the latest research findings and theoretical topics. Because developmental stage proved to be such a good predictor of science achievement and students’ ability to learn science concepts, among other things, I decided to try to help teachers discover which of their students may be lagging in their intellectual development. So I developed what is now called The Classroom Test of Scientific Reasoning (Lawson, 1978). I am pleased to say that the test has subsequently been shown to be very valid and reliable, has been translated into several foreign languages, and remains in widespread use (e.g., Bao et al., 2009; Coletta & Philips, 2005; Cracolice, Deming, & Ehlert, 2008). In fact, I still receive requests for its use almost every day. While at Berkeley, I teamed up with Warren Wollman for another key research project in which we attempted to teach fifth and seventh graders how to use an important reasoning pattern commonly referred to as the identification and control of variables. The project was designed to better understand how this advanced (i.e., formal) reasoning pattern develops by

204   A. E. LAWSON

trying to teach it to concrete operational students who have not yet acquired it, and who, in Piagetian theory, may not be able to do so. Indeed, while at Purdue, in spite of using several hands-on materials and the best learning cycle approach we could think of, we had little to no success in teaching 10th grade students how to identify and control variables (Lawson, Blake, & Nordland, 1975). However, Warren and I had considerable success with these younger students. This was partly due to the use of one-on-one training sessions, as opposed to the classroom approach used at Purdue, and a more effective way of provoking self-regulation. A key aspect of the training was allowing students to generate their own strategies that ended up being contradicted, thus provoking students to quickly abandon their initial strategies and hunt for better ones (Lawson & Wollman, 1976). Another key difference between the Purdue and Berkeley approaches was that instead of introducing the scientific terminology to refer to the reasoning pattern in question (i.e., identifying and controlling variables), we used the more familiar and intuitively understood terms “fair tests” and “unfair tests.” By repeating the use of these familiar terms in varying contexts, students were soon able to understand their meaning and were able to control variables in both familiar and unfamiliar contexts. Using Classroom Instruction to Provoke the Development of Reasoning After moving to Arizona State University in 1977, my colleagues and I continued researching several aspects of the nature of advanced reasoning, including how classroom instruction could provoke contradictions, selfregulation, and continued intellectual development. An article written with Cecil Lewis (a graduate student) and James Birk (a colleague in the Department of Chemistry) discusses a classic example of how classroom instruction can provoke the development of improved reasoning—in this case the use of hypothetico-deductive reasoning (Lawson, Lewis, & Birk, 1999). The phenomenon in question involves a lit candle sitting in a pan of water. When an inverted glass is placed over the candle and into the water, the flame goes out and the water rushes up into the glass. Many students initially believe that the water rises because the flame “consumes” the oxygen trapped under the glass, so the water is “sucked” in to replace the now empty space. This explanation contains two commonly held student misconceptions. First, flames do not consume anything in the sense that matter is not destroyed. Instead, flames convert oxygen gas to carbon dioxide gas. Second, suction (which students think of as a pulling force) does not exist. Instead, the relatively greater air pressure outside the glass pushes the water up inside the glass.

The Nature and Development of Scientific Reasoning    205

Helping students understand the accepted scientific explanation for why the water rises, and understand why the scientific explanation is accepted instead of their more intuitively appealing alternative, is no small matter because acceptance requires not only understanding kinetic-molecular theory, but also knowing how to generate and test alternative hypotheses, in this case hypotheses involving unseen theoretical entities (i.e., atoms and molecules). Nevertheless, the best instructional approach encourages students to first explore the puzzling phenomenon, raise causal questions, generate several possible explanations, and then attempt to test them experimentally. Students are usually quite capable of generating hypotheses. But few understand the need to test them by deducing explicitly stated predictions. However, provoking students to realize that their initial explanation(s) must be wrong—because their explanation(s) lead(s) to predictions that are contradicted by further evidence—helps them begin to see the light. For example, as mentioned, most students initially believe that the water rises due to the consumption of oxygen (and the resulting pull caused by suction). However, if this is true, and we repeat the experiment varying the number of burning candles, then the water should continue to rise to the previous level (more burning candles should consume the available oxygen faster, but should not consume more oxygen). Alternatively, if the water rises because the air has been heated, has expanded, and some has escaped, then increasing the number of burning candles should cause water to rise higher (more burning candles will heat and drive out more air, thus further reducing the internal air pressure). As it turns out, more burning candles do in fact cause much more water to rise, thus providing strikingly contradictory evidence to their initial hypothesis. Student misconceptions can be used to provoke contradictions and self-regulation in other contexts as well. For example, biology teachers are often confronted by what we might call the “special creation” misconception, which for some students can be as persistent as the suction misconception. Dealing with the special creation misconception (or more recently the “intelligent design” misconception) can be even more difficult because its roots lie in an often emotionally charged religious belief. For some students, accepting the scientific explanation for species diversity means rejecting part of their dominant and guiding religious worldview. Nevertheless, the teacher’s role is essentially the same as above, which is not to tell students what to believe, but to help them learn how to come to a belief. And in a science class, this means that we again propose alternative explanations and then test them. Fortunately, with respect to the alternative theories of evolution and special creation, there are several ways this can be done (e.g., Lawson, 1999, 2004; Nelson, 2000). With respect to the fossil record, for example, several observations contradict special creation theory and support evolution theory. For example, if special creation theory is

206   A. E. LAWSON

correct, and we compare fossils from the older/lower rock layers to those from younger/higher layers and to present-day species, then a) species that lived in the remote past (lower layers) should be similar to those living today; b) the older layers should be just as likely to contain fossils similar to present-day species as the younger layers; c) the simplest as well as the most complex organisms should be found in the oldest layers containing fossils, as well as in more recent layers; and d) a comparison of fossils from layer to layer should not show gradual changes in fossil forms. In other words, intermediate forms should not be found. Abundant evidence contradicts all of these predictions. The key point is that teachers need to allow students to explore phenomena, make puzzling observations, and then generate and test multiple explanations. Teachers need to be open to all explanations—most especially those that can be found to be wrong. Essentially, the explanations, once generated, must be used to deduce predictions. Evidence must then be gathered and compared with prior predictions. If the evidence and predictions match, then the explanation has gained support. However, if they do not match, then the explanation has been contradicted. Importantly, scientific beliefs are formed after consulting the evidence and comparing it to predictions. Students need to learn this, but they won’t when teachers simply tell them which ideas are right and which are wrong. Thus, improved reasoning patterns develop when they are used, and they are certainly used when alternative ideas are generated, tested, and contradicted. Contradictions force students to not only reflect on what they initially believed, but also on their reasons (and reasoning) for those beliefs and the reasons (and reasoning) for rejecting them. The point is that arguments about which ideas are right or wrong, and why they are right or wrong, provide the motivation for reflecting on and eventually abstracting the reasoning patterns, the forms of argumentation, that are used in learning (for a relatively complete list of those reasoning patterns and forms of argumentation see Lawson, 1995, pp. 50–53). What Role Do Analogies Play? A related aspect of my research concerned analogy, or analogical reasoning, as it appeared to play a key role in science during the invention of explanations, or hypotheses and theories (e.g., Lawson, 1993). Consider the historical example of Elie Metchnikoff. In 1890, Metchnikoff watched starfish larvae under his microscope as he tossed a few rose thorns among them and noticed that the larvae quickly surrounded and dissolved the thorns. This seemed to him to be similar to what happens when a splinter gets stuck in a finger. Pus surrounds the splinter, which via analogical rea-

The Nature and Development of Scientific Reasoning    207

soning, Metchnikoff hypothesized consists of tiny cells that attack and eat the splinter. So through the use of analogical reasoning, Metchnikoff was on the road to the eventual discovery of the bodies’ main defense mechanism—namely mobile white blood cells that swarm around and engulf materials such as splinters and invading microbes. Charles Darwin’s conceptualization of natural selection can also be traced to an analogy. In Darwin’s case, the analogy was between artificial selection of domestic plants and animals and the selection process that he imagined occurs in nature (i.e., natural selection). Other examples of analogical reasoning are numerous in history of science. Kepler borrowed the ellipse from Apollonius to describe planetary orbits. Mendel borrowed algebraic patterns to help explain hereditary patterns. Kekulè borrowed the image of a snake biting its tail (in a dream) to create a molecular structure for benzene. And Coulomb borrowed Newton’s patterns of gravitational attraction to describe the electrical forces that exist between sub-atomic particles. Indeed, as predicted, our research has found that both the use of physical analogies and hypothetico-deductive reasoning are positively related to the understanding of theoretical concepts such as genes, molecular polarity, bonding and diffusion (i.e., Baker & Lawson, 2001; Lawson, Baker, DiDonato, Verdi, & Johnson, 1993). Again, the classroom implication is that students need freedom to explore nature in the lab and in the field to discover puzzling observations. Students should then be encouraged to use analogical reasoning to creatively generate multiple explanations for the puzzling observations. To make sure that many ideas are freely generated, none should be criticized during this initial brainstorming period. But once several plausible explanations, and perhaps some not so plausible explanations, have been generated, they need to be tested. In this way, students learn new science concepts and theories in a way analogous to the way scientists initially invented them. A key point is that most of the concepts that lie at the heart of modern scientific thought are theoretical in the sense that they are about non-perceptible entities and processes (e.g., atoms, DNA, photons, biogeochemical cycles, natural selection, protein synthesis). Thus, for scientists to have invented the concepts and theories in the first place, they had to use analogical reasoning. Likewise, students must do the same. For students to get some sense of what DNA is like, we can help by suggesting that is like (i.e., analogous to) a twisted ladder. And to help them understand natural selection, we can have them participate in a simulation in which they play the role of birds capturing and eating mice (colored paper chips) in various habitats (pieces of colored fabric) (e.g., Maret & Rissing, 1998; Stebbins & Allen, 1975). Yet teachers need to keep in mind that although analogies and simulations should be sought and used as often as possible, their use-

208   A. E. LAWSON

fulness is limited by the students’ ability to understand not only how the analogue and the theoretical target concept are similar, but also how they differ. After all, DNA is not really a twisted ladder! Neural Research Supports Learning Cycle Instruction Another goal of my research was to understand the neurological basis of learning. This goal was motivated in part by a book my father titled Brain Mechanisms and Human Learning (Lawson, 1967). In that book, he developed a theory of learning and development based on the neural mechanisms that were known at the time. Although the book was incomplete in terms of brain functioning, it pointed readers, including me, in the right direction. The book also introduced a teaching model. That model helped form the basis of the SCIS learning cycle approach after my father moved from Michigan State University in 1967 to co-direct the SCIS project at UC Berkeley. The central issue behind my interest in neural physiology was trying to find out if the brain functions in a hypothetico-deductive way. After all, if people learn by generating and testing hypotheses, which occurs in a hypothetico-deductive way, then it follows that that is also the way the brain should function. What we now know is that the brain learns when previously non-functional synapses become functional. And this happens in one of two ways. One way is through sheer repetition and/or via emotionally charged contexts. Repetition and emotion “burns” new input into one’s synapses (one’s long-term memory), essentially by boosting pre-synaptic activity to a high enough level to create functional connections. Unfortunately, this rote way of memorizing information produces knowledge of very limited value because it remains disconnected to what one already knows. The second more effective way to form new functional synaptic connections involves linking new input with prior ideas. When neural activity is simultaneously boosted by new input and by prior ideas, the resulting pre- and post-synaptic activities combine to create new functional connections. This second way of learning produces useful transferable knowledge because the new knowledge is connected to what one already knows. Thus, the brain learns by linking new input with prior ideas, and—as I suspected—it does so in a hypotheticodeductive way (Lawson, 2006). Consider vision. Most people would guess that the brain processes information, including visual input, primarily in an “inductive” way—that is we look and we look again, and perhaps look still again, until we eventually “induce” an idea about what we are looking at. But as we now know, this is not

The Nature and Development of Scientific Reasoning    209

what happens. Instead, based on the initial look, the brain spontaneously and subconsciously generates a hypothesis of what might be out there and then uses subsequent looks to test its initial hypothesis. For example, suppose Karen, who is extremely myopic, is looking or rummaging around the bathroom and spots the end of an object that appears to be a shampoo tube. The nature of the object’s end and its location prompt Karen’s brain to generate a shampoo-tube hypothesis. Based on this initial hypothesis, as well as knowledge of shampoo tubes stored in her associative memory, when Karen looks at the other end of the object, she expects to find a cap: If it really is a shampoo tube (hypothesis), and I look at the other end (planned test), then I should see a cap (prediction). Thus, Karen shifts her gaze to the other end (actual test). And upon seeing the expected cap (result), she decides that the object is in fact a shampoo tube (conclusion). Hearing also occurs in this hypothetico-deductive way. For example, I was in my front yard early one morning raking up pine needles. I knew that it was going to be a very hot day, so I was trying to finish raking before it got too hot. As I was raking, a neighbor pedaled by on her bike and called out, “I see you are out raking while it is still cool.” But that is not what I heard. To me, and to my diminished hearing, I thought she said, “I see that the kids are still in school.” But this made no sense to me as school was no longer in session. And my kids have long since graduated from college. So I quickly rejected my brain’s initial hypothesis of what she said (i.e., I see that the kids are still in school), and gave it some additional thought, and came up with another hypothesis (i.e., I see you are out raking while it is still cool), which makes considerably more sense (i.e., is more consistent with the facts). The point here is that because the brain learns best by generating and testing hypotheses, it follows that the most effective way to teach is by encouraging students to do the same. Of course, the hypotheses students generate and test in science classes are not of the visual or auditory sort just discussed. Instead they are primarily causal in nature. Nonetheless, the hypothetico-deductive learning pattern remains the same. And as mentioned, students, and people in general, are often quite good at generating causal hypotheses. Yet they are generally lousy at testing them. Indeed, when the initial facts are consistent with a spontaneously-generated causal explanation, most people are satisfied that they have the correct explanation. But this is often an illusion. To really put the explanation to the test, additional predictions and evidence must be sought. In other words, the explanation in question must be used to deductively generate predictions, which must then be compared with new evidence.

210   A. E. LAWSON

The Pattern of Scientific Reasoning and Its Basic Inferences In addition to linking neural theory to instruction, another of my career goals was to identify how scientists make discoveries in terms of their basic inferences (e.g., Lawson, 2000a, 2000b, 2002a, 2002b, 2003a, 2003b, 2004, 2005, 2009, 2010a). The main conclusion of that research is that science is an enterprise that attempts to explain nature and does so utilizing a cyclic seven step pattern with four basic inferences called abduction, retroduction, deduction, and induction, as depicted in Figure 10.1 and as described below. 1. First, scientists undertake explorations that lead to puzzling observations (e.g., Using his newly invented telescope, Galileo observed three unexplained points of light near Jupiter). 2. Then, thanks to prior declarative knowledge, the scientist uses abduction (analogical reasoning) to generate a hypothesis—a tentative explanation (e.g., Galileo thought that perhaps the three points of light were fixed stars). 3. Next, the scientist uses retroduction to subconsciously test the initial hypothesis against prior observations (e.g., If the three points of light are fixed stars, then their relative positions around Jupiter should be random like other fixed stars. But unlike random fixed stars, the points of light appear along a straight line across the middle of Jupiter. Therefore, via retroduction, the fixed-star hypothesis is contradicted). 4. Then the scientist once again uses abduction to generate another hypothesis (e.g., Perhaps the three points of light are moons that are orbiting Jupiter). 5. The scientist then plans a test and uses deduction to generate one or more expectations (i.e., predictions) about what subsequent observation(s) should be made assuming that the hypothesis is correct and the test is conducted as planned (e.g., If the points of light are orbiting moons, and we observe their positions on subsequent nights, then sometimes they should appear to the right of Jupiter and sometimes they should appear to the left. But they should always appear on a straight line across the middle of Jupiter). 6. The scientist then conducts the test as planned and makes the subsequent observations. 7. Finally, the scientist compares his predictions with the subsequent observations and uses induction to draw a conclusion about veracity of the tested hypothesis. In Galileo’s case, his subsequent observations matched the predictions derived from his orbiting-moons hypothesis. Therefore he had found support for the hypothesis. He was then able to proclaim that he had “discovered” Jupiter’s moons.

The Nature and Development of Scientific Reasoning    211

Figure 10.1  A model of the elements of If/then/Therefore argumentation used during the generation and test of proposed explanations.

Viewed in this way, scientific reasoning and discovery consist of undertaking novel and sometimes random explorations that lead to puzzling observations. The puzzling observations are then subsequently explained by the cyclic and repeated use of abduction, retroduction, deduction, observation, and induction. Science is a collective enterprise that seeks to explain nature based the open generation and test of tentative explanations. Although science does not lead to proof or disproof, its collectiveness and openness help insure that mistakes are corrected. Consequently, science leads to useful knowledge—in the sense that reliable predictions about future events can be made. Of course scientists are not the only people who reason this way. As mentioned, all adults do, at least in some contexts some of the time. The problem is that this pattern of hypothetico-deductive reasoning has remained subconscious for virtually all adults. And without a conscious guide to effective reasoning, all sorts of subconscious biases and mistakes creep in to derail effective reasoning and produce faulty conclusions. These subconscious biases and mistakes have names such as cherry picking, confirmation bias, anchoring, outcome bias, wishful thinking, affect bias, and premature closure (e.g., Kahneman, 2011). For an example of how these biases can derail physicians from correctly diagnosing illnesses, see Lawson and Daniel (2011).

212   A. E. LAWSON

The Bottom Line from Research The bottom line from my research and that of others who have tackled similar educational issues is that meaningful learning is a “constructive” process. Thus, the most effective approach to teaching is an “inquiry” learning cycle approach based on the following four basic findings of educational research: 1. learning is a natural process in which students are inherently curious and motivated to understand their world; 2. students have distinctive experiences, interests, beliefs, emotional states, stages of development, talents, and goals that must be taken into account; 3. learning occurs best when students are actively engaged in creating new understandings and making new connections with prior knowledge, and this occurs when students confront puzzling observations and attempt to explain them by generating and testing their own ideas; and 4. learning occurs best when students’ ideas and efforts are appreciated and respected. All of this means that effective teaching takes teachers off center stage and puts student-generated questions, hypotheses, tests, evidence, arguments and conclusions on center stage. Closing Remarks Very few science educators, and even fewer teachers, knew about the learning cycle method of teaching back in 1968 when I began my career. Since then, however, I have been pleased to see that, at least among the science education community, the learning cycle has become the generally accepted method of teaching science. Yet there remains the significant problem I noted earlier in this chapter because the method has failed to gain widespread use among curriculum developers and classroom teachers—in science and beyond. Additionally, I have been pleased to see that my early research focus on reasoning, along with a similar focus by others such as Chester Lawson, John Renner, Robert Karplus, Warren Wollman, and Arnold Arons, caused a significant movement among the science education community in that direction during the 1970s and 1980s. Although interest lagged somewhat during the 1990s and early 2000s, the research pendulum seems to be swinging back. For example, organizations such as the National Institutes of Health are planning on funding a new wave of research on student reasoning, and as evidenced by several recent publications and many

The Nature and Development of Scientific Reasoning    213

requests for use of my reasoning test from researchers worldwide, a resurgence of interest in measuring and improving student reasoning patterns appears to be taking place. On the other hand, a recent troubling trend is taking place in which some science educators have become vocal in claiming that there is no scientific method. While it is true that there can be no “lockstep” method that reliably leads to scientific success, there can be no doubt that the pattern of hypothetico-deductive reasoning that I have described in this chapter lies at the heart of doing science and makes science the only tool for generating useful, reliable knowledge (e.g., Lawson, 2010b). Tossing out a focus on this hypothetico-deductive pattern and advocating an “anything goes” alternative is like tossing the baby out with the bath water. Clearly, many science educators and even some scientists fail to grasp the nature of hypotheses and predictions and the critical role played by hypothetico-deductive reasoning in doing science. In a recent review of science education research, I identified several research papers published in the Journal of Research in Science Teaching in which the authors, and apparently the reviewers and editors as well, failed to correctly differentiate hypotheses from predictions and results from conclusions (Lawson, 2010c). Indeed many, dare I say most, science educators fail to understand the central role played by hypothetico-deductive reasoning in scientific research. Yet to become truly scientifically literate, our students need to not only learn how this pattern leads to useful, reliable knowledge, but they also need to know how to use it in the service of their own lives, in life of their society, and indeed in the very maintenance of life on this planet. Unfortunately, this is not likely to happen until the science education community fully grasps the nature and power of the hypothetico-deductive scientific “method.” References Atkin, J. M. & Karplus, R. (1962). Discovery or invention? The Science Teacher, 29(5), 45. Baker, W. P. & Lawson, A. E. (2001). Complex instructional analogies and theoretical concept acquisition in college genetics. Science Education, 85, 665–683. Bao, L., Cai, T., Koenig, K., Fang, K., Han, J., Wang, J., . . . Wu, N. (2009). Learning and scientific reasoning. Science, 323, 586–587. Coletta, V. P. & Philips, J. A. (2005). Interpreting FCI scores: Normalized gain, preinstruction scores, and scientific reasoning ability. American Journal of Physics, 73(12), 1172–1182. Cracolice, M. S., Deming, J. C., & Ehlert, B. (2008). Concept learning versus problem solving: A cognitive difference. Journal of Chemical Education, 85(6), 873– 878.

214   A. E. LAWSON Fuller, R. G. (2002). A love of discovery: Science education—The second career of Robert Karplus. Dordrecht, The Netherlands: Kluwer. Kahneman, D. (2011). Thinking, fast and slow. New York, NY: Farrar, Straus and Giroux. Lawson, A. E. (1978). The development and validation of a classroom test of formal reasoning. Journal of Research in Science Teaching, 15(1), 11–24. Lawson, A. E. (1993). The importance of analogy: A prelude to the special issue. Journal of Research in Science Teaching, 30(10), 1213–1214. Lawson, A. E. (1995). Science teaching and the development of thinking. Belmont, CA: Wadsworth. Lawson, A. E. (1999). A scientific approach to teaching about evolution and special creation. The American Biology Teacher, 61(4), 266–273. Lawson, A. E. (2000a). How do humans acquire knowledge? And what does that imply about the nature of knowledge? Science & Education, 9(6), 577–598. Lawson, A. E. (2000b). The generality of hypothetico-deductive reasoning: Making scientific reasoning explicit. The American Biology Teacher, 62(7), 482–495. Lawson, A. E. (2002a). What does Galileo’s discovery of Jupiter’s moons tell us about the process of scientific discovery? Science & Education, 11(1), 1–24. Lawson, A. E. (2002b). The origin of logical reasoning: Does a cheater detection module exist? Journal of Genetic Psychology, 163(4), 425–444. Lawson, A. E. (2003a). The nature and development of hypothetico-predictive argumentation with implications for science teaching. International Journal of Science Education, 25(11), 1387–1408. Lawson, A. E. (2003b). Allchin’s shoehorn, or why science is hypothetico-deductive. Science & Education, 12(3), 331–337. Lawson, A. E. (2004). T. rex, the crater of doom, and the nature of scientific discovery. Science & Education, 13(3), 155–177. Lawson, A. E. (2005). What is the role of induction and deduction in reasoning and scientific inquiry? Journal of Research in Science Teaching, 42(6), 716–740. Lawson, A. E. (2006). On the implications of neuroscience research for teaching and learning: Are there any? Cell Biology Education, 5, 111–117. Lawson, A. E. (2009). On the hypothetico-deductive nature of science—Darwin’s finches. Science & Education, 18(1), 119–124. Lawson, A. E. (2010a). Basic inferences of scientific reasoning, argumentation, and discovery. Science Education, 94(2), 336–364. Lawson, A. E. (2010b). How many scientific methods exist? The American Biology Teacher, 72(8), 334–336. Lawson, A. E. (2010c). How “scientific” is science education research? Journal of Research in Science Teaching, 47(3), 257–275. Lawson, C. A. (1967). Brain mechanisms and human learning. Boston, MA: Houghton Mifflin. Lawson, A. E., Baker, W. P., DiDonato, L., Verdi, M. P., & Johnson, M.A. (1993). The role of physical analogues of molecular interactions and hypotheticodeductive reasoning in conceptual change. Journal of Research in Science Teaching, 30(9), 1073–1086.

The Nature and Development of Scientific Reasoning    215 Lawson, A. E., Blake, A. J. D., & Nordland, F. H. (1975). Training effects and generalization of the ability to control variables in high school biology students. Science Education, 59(3), 387–396. Lawson, A. E. & Daniel, E. S. (2011). Inferences of clinical diagnostic reasoning and diagnostic errors. Journal of Biomedical Informatics, 44, 402–412. Lawson, A. E. & Karplus, R. (1977). Should theoretical concepts be taught before formal operations? Science Education, 61(1), 123–125. Lawson, A. E., Lewis, C. M., Jr., & Birk, J. P. (1999). Why do students “cook” lab data? A case study of the tenacity of misconceptions. The Journal of College Science Teaching, 29(3), 191–198. Lawson, A. E. & Renner, J. W. (1975). Relationships of concrete and formal operational science subject matter and the developmental level of the learner, Journal of Research in Science Teaching, 12(4), 347–358. Lawson, A. E. & Wollman, W. T. (1976). Encouraging the transition from concrete to formal cognitive functioning—An experiment. Journal of Research in Science Teaching, 13(5), 413–430. Maret, T. J. & Rissing, S. W. (1998). Exploring genetic drift & natural selection through a simulation activity. The American Biology Teacher, 60(9), 681–683. Nelson, C. E. (2000). Effective strategies for teaching evolution and other controversial topics. The creation controversy & the science classroom. Arlington, VA: NSTA Press. Science Curriculum Improvement Study (SCIS). (1970). Environments: Teacher’s guide. Chicago, IL: Rand McNally. Science Curriculum Improvement Study (SCIS). (1973). SCIS omnibus. Berkeley, CA: Lawrence Hall of Science. Stebbins, R. C. & Allen, B. (1975). Simulating evolution. The American Biology Teacher, 37(4), 206.

216   A. E. LAWSON

Chapter 11

A Half-Century Effort to Create a Theory of Education to Guide the Improvement of Teaching and Learning1 Joseph D. Novak Cornell University, Emeritus Florida Institute for Human and Machine Cognition

Introduction I recall when I was in second grade at Van Cleve School that I brought to class various jars that my mother used for canning to illustrate relationships between pints, quarts, and gallons. Our principal, Miss Lohman, was so impressed with my presentation that she had me repeat it in all the other classrooms. I was happy to have the chance to visit other classes, but I did not think what I had to say was that important. Of course, I did not know that Piaget (1926) and others had shown that only about 10% of high school seniors can consistently conserve volume of cylinders, let alone provide specific relationships between them. I also recall that in my 4th grade

Going Back for Our Future, pages 217–248 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

217

218    J. D. NOVAK

class with Miss Bartz some of the children had difficulty recognizing that 3 × 4 was the same as 4 × 3 and similar number relations. I remember thinking, “what is wrong with these kids that they have trouble with such simple ideas. Isn’t there a way to teach them so they could all understand simple ideas and relationships?” Most of my 12 years in public schools were filled with boredom, and my report cards showed my lack of interest and effort to learn what we were being taught. The two exceptions were Sam Drage’s physics course and Miss Fisher’s English literature course, both of which stressed understanding the subject matter. Never during those 12 years did I think that I might become an educator and commit myself to seeking ways to make education more effective! When I began studies at the University of Minnesota in 1948, I was not sure how long I would survive what I assumed would be much more rigorous classes, nor if my job pressing clothes at Central Cleaners would be adequate to pay the bills. From the first quarter of college studies, I was impressed with how different my classes were from most high school classes and how much more the emphasis was on understanding the subject matter, not just memorization. I also began reading many paperback books on various subjects just to learn more about history, sociology, anthropology and science. For the first time in my life I found school to be exciting and challenging. By my sophomore year I decided I would like to prepare for a career in teaching. Working 30 to 40 hours per week at Central Cleaners limited my efforts to excel in class work; however, I was able to reduce my work time and became a serious student by my senior year. Most of my education classes I found lacking in substance and, in contrast to the sciences, also lacking in solid principles and theories. Subsequent graduate studies in education also contrasted sharply from the science courses I continued taking that were characterized by solid theoretical and philosophical foundations. Why, I began to wonder, can’t education become more like sciences? Thus began my early efforts to work toward a science of education. This chapter will describe the journey that I, my students, and supportive colleagues have taken to build a theory-based science of education. New Positions Offered New Opportunities It was my good fortune to be offered a teaching assistantship in the Botany Department at the University of Minnesota upon completion of my BS degree in science and mathematics education. Although I had taken a concentration in botany for my BS degree, I was in the College of Education, not the College of Arts and Sciences. This opportunity undoubtedly arose from a research project I had chosen to pursue in my senior year in the

Half-Century Effort to Create a Theory of Education     219

laboratory of Professor Frankel, who subsequently also hired me as his parttime research assistant. Thus while I was a graduate student in education, I also had the opportunity to teach botany and biology courses and to conduct research in plant physiology as Frankel’s research assistant. In my fifth and final year in graduate school, I was offered a half-time instructor position and placed in charge of introductory botany laboratories. This gave me the opportunity to better manage instruction to allow for my PhD thesis research where I studied the effect of a six-week intensive research project in botany done by an “experimental section” and compared achievement with “traditional” laboratory groups (Novak, 1958, 1961). This was my first experience in modifying science instruction and assessing the effects on student learning. My experiences at the University of Minnesota opened the door to an assistant professor position in the Biology Department at Kansas State Teachers College at Emporia, with a primary responsibility to conduct research on biology teaching and to supervise MS students’ research on biology teaching. The Biology Department at Emporia State encouraged and supported participation in state and national professional organizations, and this led to me becoming friends with some of the leading science educators in the U.S. When a joint position in biology and education opened up at Purdue University in 1959, I was nominated by the Dean of Education at the University of Kansas and joined the Purdue faculty in August. With excellent support from the Biology Department, and good MS and PhD students, it was possible to begin to build a science education research program that I tried to pattern along the lines of research groups in the sciences. In fact, the Biology Department at Purdue University later added a PhD program option in biology teaching as one of the departmental options. Strong as the support was for my work at Purdue University, the joint appointment in biology and education sapped my energies and made it difficult to pursue the kind of educational research effort I saw as necessary. The opportunity came to join Cornell University in 1967 with the principal responsibility of building a world-class research program in science education. I saw this as the ideal position for me. This proved to be an excellent place to pursue building the kind of theory-based educational research program I thought was needed to improve education significantly. Cornell University was a great place to work, and it was with some misgivings that I chose to retire in 1995 to pursue extending application of the theory, tools, and ideas we had developed to corporate and other organizational settings. This also proved to be fruitful, and I learned much about extending educational ideas and tools into other organizations.

220    J. D. NOVAK

The Search for New Theoretical Foundations In the 1950s, the Western world was awash in a sea of behavioral psychology. Behavioral psychology held that the only scientific way to study animal or human learning was to observe and record manifest behaviors of the organism. One must not speculate in the inner workings of the brain, since these were not directly observable. Hundreds of thousands of “experiments” were performed, mostly with animals, to discover “laws of learning.” The epistemology driving the behaviorist was positivism, espousing the idea that once true laws were discovered, they would remain the truths forever. The philosophy of science course I took as a graduate student was taught by Herbert Feigl, a world renowned logical positivist. I recall my arguments with him where I maintained that the way we did research in botany did not fit with logical positivist ideas. Of course I was no match for the erudite Professor Feigl. I did find the ideas of James Conant in his On Understanding Science (1947) to be very congruent with my thinking at that time. Conant’s writings and later his protégé Thomas Kuhn’s (1962) The Structure of Scientific Revolutions marked a turning point in thinking about the nature of knowledge and knowledge creation. These works and Toulmin’s (1972) book, Human Understanding, Volume One: The collective use and evolution of concepts, helped me to understand the nature of concepts, and his views on the creation and evolution of concepts made sense to me. These ideas later became the cornerstone of my developing theory of education. Having rejected behavioral psychology, the best ideas I could find to guide my PhD study as an acceptable learning theory was Norbert Wiener’s (1948, 1954) cybernetic model that saw the brain as a complex information storage and processing organ. Vygotsky’s (1926/1962) ideas on learning would have been much more powerful, but I did not learn about his work until the late 1960s. The data I gathered in my PhD study completed in 1957 could not be well explained using cybernetic theory, so I was searching for a better theory of learning to guide our research and instructional innovation. I recall how enthusiastic my graduate students and I were when we began to study David Ausubel’s assimilation theory of learning as presented in his 1963 book, The Psychology of Meaningful Verbal Learning. While his theory contains just seven major principles, understanding each principle requires some understanding of the other six, so one must proceed iteratively to build gradually an understanding of his theory. The fundamental question was: How can we help people become better learners? From early in my career to the present day, my belief has been that we can never make this a better world to live in unless we can develop better ways to help people “get smart.” It has been my conviction over the years that education, both formal and informal, can be dramatically improved—if we can make the study of education more like the study of sci-

Half-Century Effort to Create a Theory of Education     221

ence, in other words, guided by solid theory, principles, and productive methodologies. In short, I believed we needed new and more powerful paradigms, to borrow Kuhn’s (1962) term, to guide educational research and practice. This chapter is my autobiographical sketch of my search for better ways to educate. It is necessarily an abbreviated autobiography, and I cite only a few of the people who contributed to our work, but I hope it can serve to illustrate, especially to students and younger faculty in education, that it is possible to improve education significantly. Teaching Biology and Biology Teachers When I moved in 1959 to Purdue University with a joint appointment in biology and education, my principal work was instruction in methods courses and intern teacher supervision, plus supervision of MS and PhD students. I was surprised at how little conceptual understanding of biology and other sciences was evidenced in my certification candidates and also in many experienced teachers enrolled in our summer programs. From 1959 through 1962, the Department of Biological Science underwent a major curriculum overhaul, in which I participated as co-chairman of the Curriculum Committee. Most of the existing introductory courses were scrapped and a new six-semester sequence was developed, including Principles of Biology, Cell Biology, Developmental Biology, Genetics, and Ecology. All courses used examples from plants, animals, and microbes to varying degrees, and all placed major emphasis on understanding basic concepts. As teaching candidates emerged from the new program, there was a striking difference in their mastery of biology and related science concepts and their ability to use these in innovative independent research projects. Teaching candidates were now acquiring an understanding of basic biology concepts needed for effective teaching (Novak, 1963). My experiences at Purdue University taught me how important it was for teachers to develop a deep conceptual understanding of their subject. Over the years I have seen that this is not common for teachers, with most progressing through college memorizing subject matter rather than making conscious attempts to understand the subject matter. This remains a principal problem in the improvement of school teaching. Developing a Better Understanding of Human Learning As noted earlier, another important change in my thinking and the work of my graduate students occurred in the early and mid-1960s. Our research

222    J. D. NOVAK

studies were producing results that consistently pointed to the idea that information storage and information processing were not distinct brain functions, as cybernetic theory would suggest, but rather interdependent and related to the kind of learning approach utilized. As our “cybernetic paradigm” was beginning to crumble, David Ausubel’s (1963) cognitive learning theory was published, and our research group moved quickly to embrace his idea, as noted above. By 1964, all of our research projects and also our instructional improvement projects shifted to Ausubel’s “meaningful learning paradigm.” It was also in the early 1960s that I became familiar with Piaget’s (1926) work through seminars with a colleague, Charles Smock, who had studied with Piaget. While our research group found much of value in Piaget’s studies and developmental ideas, our data did not support his “developmental stages” ideas. Moreover, Piaget was a developmental theorist, and we saw learning theory as more fundamental to understanding both learning processes and developmental processes. Ausubel’s assimilation theory seemed to be more powerful and more parsimonious in explaining learning, learner success, and learner failures (Novak, 1977b). Ausubel’s 1968 book, Educational Psychology: A Cognitive View, further elaborated his ideas. I had the opportunity to work with Ausubel in the 1970s, and he invited me to co-author a revision of his Educational Psychology book in 1978, and this work offered me extensive opportunities to discuss learning issues with him. During the later 1960s and 1970s, there emerged what might be called the “battle of educational paradigms.” In the world of psychology departments, behaviorism or behavioral psychology held a very firm hegemony, and any “mentalistic” theory of learning or development was suspect. Cognitive learning ideas were virtually shut out from the academy, and Ausubel had great difficulty getting many of his papers published in leading psychology journals. Mager’s (1962) behavioral objectives were the only valid foundation for instructional planning in the eyes of psychologists and also in the eyes of many educators, who failed to see the limiting psychological and epistemological foundations of behavioral objectives. In the world of science education, there was overwhelming acceptance of Piaget’s ideas by almost the entire community. Our group at Purdue University and later (1967) at Cornell University was almost alone in our critical views regarding Piaget’s theories and embraced instead Ausubel’s ideas. So rigid and so dominant were many of the proponents of Piaget’s ideas that none of the numerous research proposals I submitted to the National Science Foundation (NSF) or the U.S. Office of Education (USOE) were accepted for funding, at least not until 1978, when Mary Budd Rowe helped to steer one of my NSF proposals through her division. For two successive years, none of the research papers proposed by me nor by my students were accepted

Half-Century Effort to Create a Theory of Education     223

for presentation at meetings of the National Association for Research in Science Teaching, an organization in which I served as president in 1969. By the mid-1970s, the “cognitive revolution” had begun in some psychology departments and educational psychology programs. While behaviorism, and narrow interpretations of Piaget’s work, remain alive and well in school and corporate settings, these are dying paradigms in the scholarly studies focused on human learning. The climate for the kind of work we were doing 40 years ago is now much improved. In fact, in 1978 I received my first NSF grant for a study dealing with the feasibility and value of using concept maps and Vee diagrams in junior high school classes (Novak, Gowin, & Johansen, 1983). Throughout the 1960s and 1970s, when so much criticism of our work was prevalent from educationists and, to some extent, psychologists (although the latter with few exceptions ignored our work), my students and I were supported both intellectually and financially primarily by the science community. My students were often favored for teaching assistantships in science departments. We also received support from the U.S. Department of Agriculture Hatch program, the College of Agriculture and Life Sciences at Cornell University, and Shell Companies Foundation. So often I had occasion to be thankful for my affiliations with science departments and scientists, including Nobel Laureate scientists such as Roald Hoffmann and Kenneth Wilson. These associations provided much of the validation of our work during the 1970s and 1980s. My students, and a number of visiting professors, provided much of the caring and personal support anyone needs to pursue the difficult task of theory development to guide educational improvement. Creating a Theory of Education As our research and instructional improvement programs evolved, building heavily on Ausubel’s (Ausubel, Novak, & Hanesian,1978) assimilation theory of meaningful learning, several patterns began to emerge. First, it was clear that learners who developed well-organized knowledge structures were meaningful learners, and those who were learning primarily by rote were not building powerful knowledge structures and often showed numerous misconceptions (Novak, 2002). Second, while experience with what Ausubel called “concrete empirical props” and science educators call “hands on experience” was important, it was also important to carefully clarify the meanings of words (or concept labels) and propositional statements. Much of this could be done by didactic or reception instruction, provided that it was integrated with appropriate experience. In agreement with Ausubel, our work showed the importance of distinguishing between learning approach and instructional approach. With regard to instruction, either re-

224    J. D. NOVAK

ception instruction or inquiry (or discovery) approaches can be very rote or very meaningful learning experiences. This is illustrated in Figure 11.1. The rote-meaningful learning continuum is distinct from the receptiondiscovery continuum for instruction. Both reception and discovery instruction can lead to rote learning or meaningful learning, depending on whether or not the learner chooses to seek to integrate new knowledge with their prior relevant knowledge and the adequacy of that prior knowledge. School learning needs to help students move toward high levels of meaningful learning, especially in reception instruction that is still the most common. In the last 20 years there have been intensive efforts to engage students in “hands on inquiry activities.” While I applaud some of these efforts, students who are not guided to develop the conceptual knowledge needed to make sense out their inquiry activities may end up learning very little (Mayer, 2004). “Inquiry” activities in science and mathematics can be just as meaningless as rote learning information to pass multiple-choice tests. This has been a running battle for me with people associated with Project 2061 of the American Association for the Advancement of Science (AAAS). So far I see little progress with this and similar groups. There is also the problem of inordinate delay in instruction on basic science concepts such as the nature of matter and energy that should begin in the primary grades.

Figure 11.1  Ausubel helped to clarify the distinction between the learning continuum, from rote to highly meaningful, and the instructional continuum, from reception to autonomous discovery.

Half-Century Effort to Create a Theory of Education     225

Another pattern that was evident in our research was that the learners’ approach to learning was somewhat related to their epistemological ideas, albeit the nature of the relationship even today remains problematic and is one of our continuing research concerns. We observed some tendency for those students who were most positivistic in their epistemology ideas to favor rote learning approaches, and those who hold more constructivist ideas tended to favor meaningful learning strategies. In any case, it was evident in our research that philosophical issues, and especially epistemological issues, needed careful consideration in education. Finally, it became increasingly evident that in educating, we “reap what we sow.” Instruction and evaluation emphasizing or favoring rote learning strategies lead to little improvement in learner’s usable knowledge structures, whereas the reverse was the case when meaningful learning strategies were encouraged or favored (Mintzes, Wandersee, & Novak, 1998). Parallel to the evolution in psychology of learning in the 1960s and 1970s was an evolution in philosophy and epistemology. As already noted, some of this was sparked by Conant and later Kuhn, but then a cascade began, stimulated by people such as Toulmin (1972) and Brush (1974), and an unstoppable rush away from positivistic epistemologies occurred. More recent philosophers, such as Feyerabend (1988) and Miller (1989), argue for “realists” epistemology, and von Glasersfeld (1984) and others argued for “constructivist” epistemologies. These parallel developments in psychology and philosophy encouraged me to take a try at synthesizing a theory of education. The dream 1 had as a graduate student in the early 1950s, that education could become more like science, guided by theory and principles, seemed within reach. Given a sabbatical leave in 1973–1974, I had the time needed to attempt the synthesis that led to publication in 1977 of A Theory of Education (Novak, 1977a). I had already organized a course called Theories and Methods of Education that afforded me the opportunity to debate and share my ideas with students and visiting faculty, including insights and application of the theory. In my view, progress in our research and instructional development programs accelerated significantly after we had a viable theoretical foundation to work from. Perhaps the most significant work that helped to build my theory of education derived from a 12-year longitudinal study described below. A 12-Year Longitudinal Study of Science Learning Given the “battle of paradigms” that was going on in the 1960s and 1970s, it seemed to our research group essential to determine what basic science ideas young children could understand and whether or not this un-

226    J. D. NOVAK

derstanding would facilitate future learning, as predicted from Ausubel’s theory. However, my work with lower elementary school teachers as I was writing and testing science ideas in the early 1960s for our elementary science series, The World of Science (Novak, Meister, Knox, & Sullivan, 1966), indicated that most of these teachers did not understand or could not teach basic concepts of atoms, molecules, energy, energy transformations, and so on. Therefore, I decided to adopt an audio-tutorial strategy we had developed with college botany students at Purdue University (Postlethwait, Novak, & Murray, 1964, 1972, Chapter 6) and use this strategy to teach primary school children. A sabbatical at Harvard University in 1965–1966 afforded me the time and resources to develop, evaluate, and refine several audio-tutorial lessons working with first grade (6- to 7-year-old) children. When I returned to Purdue University, we obtained some U.S. Department of Education funding through the West Lafayette Public Schools and began further development, testing, and refining of audio-tutorial lessons. When I moved to Cornell University in 1967, we obtained similar support through Ithaca Public Schools and also support from Shell Companies Foundation. By 1971, we had developed some 60 audio-tutorial science lessons designed for children in grades one and two. Incidentally, when these lessons were placed in school learning centers rather than primary grade classrooms, even fifth- and sixth-grade students (10–12 years old) worked with the lessons with enthusiasm. Most teachers who took the time to work through the lessons found them “very enlightening.” Parenthetically, our work with audio-tutorial instruction in botany courses grew out of the labeled photomicrographs and other materials I developed for a study guide for my PhD research. Postlethwait began using these photos and drawings to supplement taped comments for students who had missed lectures. Student response was so positive that he began to add lab demonstration materials, and within a year, the traditional laboratory and much of the lecture material was dropped in favor of self-paced, audio-tutorial study in a new “learning center” that was established in the biology building. Hundreds of visitors came to Purdue to see the learning center in operation, and Postlethwait’s approach was replicated in many schools and universities around the world. At Cornell University, my graduate students helped to develop and evaluate an introductory audio-tutorial physics course in 1969 (See http://www.physics.cornell.edu/academics/courses/undergraduatecourses/). Other students helped to develop an audio-tutorial introductory biology course (See http://biology.cornell.edu/advising/courses.html). Both course continue to be offered today. Application of this instructional strategy in elementary school classrooms proved to be very successful, but we found we had to discourage classroom discussion led by the teacher on some topics to avoid introduction of teacher misconceptions, so prevalent at the elementary school level. From our

Half-Century Effort to Create a Theory of Education     227

pool of 60-plus audio-tutorial lessons, we selected 28 for administration to children on a schedule of roughly one new lesson every two weeks. The lessons were carefully sequenced to provide for early introduction of basic concepts and progressive elaboration of these concepts in later lessons. Some modifications in lessons were made as we progressed over a twoyear period. Figure 11.2 shows a child working with materials in a carrel unit. Most of the equipment we needed to afford hands on experience to illustrate concepts was designed by our team and replicated at low cost in a sheltered workers program in Ithaca. A group of 191 children were given this instruction during 1971–1973; we called this group the “instructed” group. Children were interviewed every four to six weeks to record their understanding of concepts taught. A similar group of 48 children enrolled in the same classrooms with the same teachers during 1972–1974, one year later, did not receive the audio-tutorial science lessons; we called this group the “uninstructed” group. By the end of the second year of the study, when “instructed” and “uninstructed” students could be compared, it was evident that instructed students were benefiting from the lessons. After grade two, all students received the same instruction in science as delivered in Ithaca public schools. We did periodic interviews of both instructed and uninstructed children throughout their tenure in Ithaca schools, although for

Figure 11.2  A six-year-old child working in an audio-tutorial carrel on a lesson dealing with electric energy transformation during our 12-year longitudinal study. The apparatus shown in the picture was developed by our team members to illustrate transformation of electrical energy (from dry cells) into light, heat, and kinetic energy.

228    J. D. NOVAK

various reasons we had to limit subsequent interviews to concepts dealing with the particulate nature of matter, energy, and energy transformations. During our first year of interviewing, we recognized the problem of interpreting interview tapes and transcriptions. Discerning patterns of changes in children’s conceptual understandings from these tapes and/or transcriptions was overwhelming. We began to accumulate file drawers full of interview transcripts. It became obvious that we needed a better method or tool for representing children’s cognitive structures and changes in cognitive structures. Reviewing again Ausubel’s ideas and his theoretical principles, we decided to try to represent children’s ideas as hierarchically organized frameworks of concepts and propositions. In short, we invented a new knowledge representation tool: concept maps. These were constructed by our staff from interview transcripts to represent the knowledge evidenced by the child. Figure 11.3 shows two concept maps constructed for one of our students; the upper map represents the student’s knowledge in Grade 2, and the lower map represents the student’s knowledge in Grade 12. It is evident that this child was learning science meaningfully and was elaborating her conceptual frameworks. Although both her Grade 2 and her Grade 12 concept maps show some misconceptions and missing concepts (e.g., the concept of space is missing in grade 12), it is evident that Cindy has been building and elaborating her knowledge of the nature of matter and energy. From the time our first children received instruction and were interviewed until the time when the last uninstructed students were interviewed in Grade 12, a period of 13 years had elapsed. Many graduate students participated in the study, and both MS and PhD students used some of the data for their thesis work. Finally, Dismas Musonda compiled data for the entire span of the study, and his PhD thesis served as the primary database for a publication on the study. Because of the extraordinary nature of the study and data obtained, it was necessary to revise and resubmit the paper three times before it was accepted for publication (Novak & Musonda, 1991). In response to editorial requests, graphic presentation of the data was dropped, but one of the original figures is reproduced here in Figure 11.4. Several remarkable things are shown in Figure 11.4. First, it is obvious that the instructed students showed fewer “naïve” or invalid notions and more valid ideas than the uninstructed students. This difference was highly statistically significant and practically very important. We see that instructed students had less than half as many invalid notions and more than twice as many valid notions as uninstructed students. Furthermore, Figure 11.4 shows that instructed students became progressively better, both with regard to valid and invalid notions, but this was not the case for uninstructed students, especially from Grade 7 forward, when formal instruction in science began in Ithaca schools. These data strongly support the theoretical foundations for the study, namely that when powerful anchoring concepts

Half-Century Effort to Create a Theory of Education     229

Figure 11.3  Two concept maps drawn from interview transcripts for Cindy in grade two (top) and grade 12 (lower), illustrating her developing conceptual understanding.

are learned early in an educational program (Ausubel calls these “subsuming concepts”), they should provide a foundation for facilitation of later learning. Obviously this occurred. Since the instructed and uninstructed samples (by Grade 12) did not differ significantly in ability as suggested by

230    J. D. NOVAK Instructed

Uninstructed

10

Mean Number of Valid Notions

9.73

8.75

9 8

7.47

7 6 5

4.63

4.5

4.13

3.85

4 3 1.75

2 1 0

7

2

12

10

Grade Level

Instructed

Uninstructed

7.7

8

Mean Number of Invalid Notions

7

6.4

5.9

6

6.25

5 4 3

2.57

2.11

2

1.25

1 0

2

7

10

1.07

12

Grade Level

Figure 11.4  Bar graphs showing the frequencies with which “instructed” students (black bars) and “uninstructed” students (clear bars) showed valid notions about the structure of matter and nature of energy (top graph) and the number of invalid notions held (lower graph). Note that only the instructed students show continuous improvement over the years.

grade-point averages and SAT scores (Novak & Musonda, 1991), the only factor that could account for the huge differences observed was the power concept learning from the early audio-tutorial science lessons. The content of the science lessons offered in Grades 1 and 2 was relatively abstract and certainly not commonly taught in these grades. The fact that these lessons had a highly facilitating effect on future learning indicates that they were learned meaningfully, at least for a substantial percentage of the students. Clearly these results cannot be explained with either behavioral theory (there was very limited reinforcement and rehearsal of ideas) or Piagetian developmental theory, which posits little success with learning abstract sci-

Half-Century Effort to Create a Theory of Education     231

ence concepts at ages six to eight (see, for example, Shayer & Adey, 1981). The results are consistent with results from more recent studies such as those of Matthews (1980), Carey (1985), Gelman (1999), and Keil (2011). Virtually all cognitive psychologists now recognize that children’s learning capabilities have been grossly underestimated. However, recent AAAS (2011) Benchmarks and National Research Council (NRC, 2011) Standards curriculum recommendations still fail to recognize this capability. Our 12-year-study, the only one of its kind reported to date, clearly indicated that our program was on the right track, both theoretically and in terms of practical consequences for science education. It has been common in the sciences that the necessity for developing new tools for making or processing records obtained in research have led to new practical applications. For example, development of oscilloscopes to study and record electromagnetic wave patterns led to the development of television. In an analogous manner, the concept mapping tool, as we developed it, proved to be a useful tool in many education applications. First, making concept maps from interviews has permitted us and others to observe specific changes in learners’ cognitive structures with relative ease and a degree of precision that was not possible by reviewing interview recordings or transcripts alone (Coffey, Hoffman, & Novak, 2011). They are useful for identifying and remediating misconceptions or “alternative conceptions” (Novak, 2002; Novak & Abrams, 1993; Novak & Musonda, 1991; Wandersee, Mintzes, & Novak, 1994; Mintzes, Wandersee & Novak, 1997). Subsequently, we and others have found that interviews in any knowledge domain can be better interpreted using concept maps, including interviews with consumers, patients, or counselors. Second, concept maps can help learners organize subject matter and facilitate learning and recall of any subject matter (Novak & Gowin, 1984; Novak & Wandersee, 1990). Third, concept maps can be a valuable evaluation tool—one I have used extensively in my own classes (Novak & Ridley, 1988). Fourth, they can be useful for teachers and other educators for organizing and planning instructional material (Symington & Novak, 1982; Novak, 1991, 1995). Fifth, they can identify cognitive-affective relationships that can be enormously helpful in counseling settings (Mazur, 1989). Sixth, they help in team building and reaching consensus on project goals and objectives (Edmondson, 1995). Seventh, they can enhance creative production, since creativity requires well-organized knowledge structures and an emotional proclivity to seek new interrelationships between diverse domains of knowledge (Novak, 2010, in press). These and other applications of the concept mapping tool have had a major impact on our programs, including recent applications in corporate settings (See Moon, Hoffman, Novak, & Cañas, 2011 for many examples). Many other examples of the use of concept maps can be found in the five Proceedings of International Conferences on Concept Mapping at http://cmc.ihmc.us.

232    J. D. NOVAK

Development of the Vee Heuristic Much of our research dealt with instruction in science laboratories. A continuing problem observed was that students often proceeded through the laboratory work doing what was prescribed in the laboratory manual, but often not understanding why they were doing what they were doing, nor could they explain what the meaning of their records, graphs, tables, charts had for understanding better the science they were studying. My colleague, Bob Gowin, had found this to be the case in other disciplines as well. Gowin’s PhD work was in philosophy, and he maintained a strong commitment to understanding the structure of knowledge and the process of knowledge creation. Gowin (1970, 1981) first proposed five questions that need to be answered to understand the structure of knowledge. In 1977, Gowin conceived a new way to represent twelve elements involved in knowledge creation and to suggest their interrelationships. He created the “Knowledge Vee.” This gradually underwent some modifications, and Figure 11.5 shows our current form of the Vee heuristic as used in our current work and definitions for 12 epistemological elements involved in the structure and creation of knowledge. We proceeded to use concept maps and the Vee heuristic in many of our research studies and to aid in the design of instruction. While most students, teachers, and professors immediately see value in concept maps, it has been our observation that the Vee heuristic is more difficult to grasp. One reason for this may be that most of us are brought up in patterns of thinking about knowledge and knowledge discovery that are primarily positivistic in character. The highly fluid, complex process of knowledge creation represented in the Vee can at first be overwhelming. With time and effort, however, we have found that the value of the Vee heuristic is recognized, and concept maps are seen as useful to represent some aspects of the Vee, namely the structure of concepts and principles guiding the inquiry (on the left side) and the structure of knowledge and value claims (on the right side). I know of no other heuristic tool for representing the structure of knowledge, and I would predict that in 20 to 30 years it may become more widely used in schools and corporations. Learning How to Learn While writing the draft of A Theory of Education, I organized a new course, “Learning To Learn.” Initially I thought the course would be appropriate for freshman and sophomore students, but mostly juniors and seniors enrolled. We had also found in earlier work with a special program for freshmen that very few Cornell University students think they have problems

Half-Century Effort to Create a Theory of Education     233

Figure 11.5  The Knowledge Vee showing key epistemological elements that are involved in the construction or description of new knowledge. All elements interact with one another in the process of constructing new knowledge or value claims or in seeking understanding of these for any set of events and questions.

learning. For the most part, they achieved predominantly “A” grades in their high school work and even in most courses taken as freshmen. It is not until their sophomore or junior years at Cornell University, when they take courses where rote memorization will not suffice to get “A” grades, that

234    J. D. NOVAK

students begin to realize they must either be less able than they thought they were before or they must be doing something wrong in their studies. At first, many of the students who enrolled in “Learning to Learn” were essentially seeking better tricks for memorizing, taking notes, and preparing for exams. They wanted a “quick fix” for their declining grade point averagees. They were not looking for a course that would help them develop new ways of learning and new insights into the nature of knowledge. The dropout rate the first year or two from my course was about 30%. Most of those students who persisted ended the semester with a new sense of empowerment over their own learning. It was not uncommon for students to tell me they never knew there was another way to learn other than rote memorization. As the merits of the course became better known, dropout rates dropped to near zero. They were grateful to me for “turning their lives around” as they gained skill and confidence in meaningful learning, aided, in part, by concept mapping and Vee diagramming tools. In 20 years of teaching this course, I never had a single student who completed the course who failed to gain skill and confidence in his or her ability to learn meaningfully. Some admitted they would continue to learn by rote in some courses, especially in courses that were poorly taught or courses that held little interest for them. Many in later years have written to tell me that their high success in graduate studies would have been unlikely without “Learning To Learn.” The sorry fact is that Cornell University students are among the very best high school graduates. And yet, most of them have done most of their previous learning largely by rote. Studies of student learning approaches in this country and abroad have shown that similar patterns prevail in other countries. The result is that students may graduate from both high school and university, and yet very little of what they memorized is functional knowledge. For example, in a widely circulated videotape, A Private Universe (Schneps, 1989), graduating seniors, graduate students, alumni, and faculty were asked to explain why we have seasons. Twenty-one of 23 persons interviewed at random could not give a satisfactory explanation—including one graduating senior who had just recently completed a course called “The Physics of Planetary Motion”! Obviously he had been learning almost totally by rote, as may have been the case for many others interviewed. The universality and pervasiveness of rote-mode learning patterns in schools and universities is astonishing, given our current knowledge of the limited value of such learning. As Pogo has said, “We have met the enemy, and the enemy is us!” Sixty years after my own disappointing school learning experiences, I find that, for the most part, not much has changed. Why? What can be done about it? These remain nagging questions. Although research on the value of the Vee heuristic for facilitating learning is not as extensive as that on the value of concept maps, our data indicates that secondary school and university students can benefit from using this tool

Half-Century Effort to Create a Theory of Education     235

(Novak, Gowin, & Johansen, 1983). An interesting research project might be to further assess the value of this kind of instruction in elementary, secondary, and/or college classroom settings. Many of the studies reported at the International Conferences on Concept Mapping, mentioned above, support the value of the concept mapping and Vee tools for encouraging meaningful learning. There is hope for improvement of education in the future. I take heart in the fact that Learning How To Learn at this writing has been published in nine languages, with a new Arabic translation in production. The ideas and strategies that were presented in this book are slowly taking hold, not only in the U.S. but also in countries around the world. The 350-plus graduate students and visiting professors who have done theses or worked with me are everywhere in the world, and their students and their students’ students (my “intellectual great grandchildren”) continue to multiply and add their efforts to the slow process of educational reform. Thousands of others who embrace similar ideas are pushing for educational change, changes that move learners from the disempowering effects of rote learning to the empowering consequences of rich, meaningful learning. For me, another source of optimism for future improvement of education is my belief that the theoretical foundations for education are improving. The growing consensus on the validity of constructivist epistemological ideas and cognitive learning principles suggest that the science education community and the education community in general are moving forward (Linn, 1987). My efforts to synthesize epistemological ideas and psychological ideas as they relate to the construction of knowledge were first published in 1987. Both disciplinary and personal knowledge is acquired through Human Constructivism (Novak,1993) . Although we have been working with professors to apply the Vee heuristic and concept mapping to their research work for more than two decades, our first systematic effort to study the value of the Vee and concept maps as tools to facilitate research and new knowledge creation began in 1993. We were fortunate to enlist the cooperation of Professor Richard Zobel, who headed the “Rhizobotany group,” a research group at Cornell University focused on the understanding of roots and root functions. Some of the preliminary findings were that even experienced researchers in the group had only a sketchy knowledge of the overall intellectual activities of the group. Concept maps and Vee diagrams helped the group see where individual projects contributed to broader research questions. Some of the researchers evidenced little knowledge of, or interest in, epistemological issues. A preliminary report on this work has been published (Novak & Iuli, 1995). I began a revision of A Theory of Education during my sabbatical in 1987– 1988, but pressures of other work required that I set this aside. During my sabbatical in 1994–1995, I resumed work on this manuscript; however, our

236    J. D. NOVAK

research studies between sabbaticals and the initiation of work with corporations (described below) led me to restructure substantially the earlier work and to write essentially a new book with the title Learning, Creating, and Using Knowledge: Concept Maps as Facilitative Tools in Schools and Corporations (Novak, 1998a). This new work served as the foundation for much of our work with corporations and other organizations such as NASA, the Department of the Navy, the National Security Administration, and other governmental and corporate groups. The latter work also funded further refinement of CmapTools (described below), enhancing the potential of this software for capturing and archiving expert knowledge, solving problems, and enhancing creativity and learning in many organizations. Some of this work is summarized below and in two recent books (Novak, 2010; Moon, Hoffman, Novak, & Cañas, 2011). My Third Career—Helping Corporations Learn For some four decades, I have had occasional students with experience in business and had participated in meetings and seminars with persons from the business world. While I found individuals who saw the relevance of our work to business, they were predominantly from lower ranks in business organizations, and their efforts to introduce the ideas and tools generally met with quick dismissal by upper-level management. In 1991, Alan McAdams, professor in the Cornell Johnson School of Business and Management, invited me to co-teach a course with him in the business school. For some years he had been an enthusiast of concept mapping, and he wanted to learn more about education. I was eager to learn more about business and the corporate world. We enjoyed good success with our new course (it received the highest student ratings for any course in the Johnson School), and we taught it again in 1992, 1993 and 1994. Most of the students were MBA candidates with four to eight years of experience in the business world. As part of the course project activities, we worked with for-profit and not-for-profit organizations, using concept mapping and Vee diagrams as tools to better understand the structure and function of the organizations. The principal data gathered derived from intensive interviews with key persons in the organization and processing the interviews using concept maps and sometimes also using Vee diagrams. What we found was that all of the organizations studied had some serious problems in one or more of the following areas: limited personnel knowledge about the “mission” or strategic plan of the organization, functioning of individual units, barriers to communication, and failure or limited ability to learn as an organization. Both McAdams and I were struck by the energy and enthusiasm of our students for the work we were doing, but

Half-Century Effort to Create a Theory of Education     237

also by the resistance of most of the organizations to accept the implications of the findings from our studies. One thing became eminently clear to me: nonprofit and for-profit corporations could benefit significantly by applying ideas from A Theory of Education, Learning How To Learn and, later on, from Learning, Creating, and Using Knowledge. The problem was finding an organization with the right leadership to demonstrate the value of the tools and ideas. There is the saying, “Nothing is more unstoppable than an idea whose time has come.” Something very important has happened in the business world in the past two decades—globalization. While it is true that worldwide trade has been part of the business world at least since prehistoric times, new transportation, communication, and other technologies have evolved to the point where almost any product can be made almost anywhere and shipped everywhere at relatively low cost. Suddenly, U.S. and other nations’ businesses have found themselves head to head in competition with businesses all over the globe. In 1988, Prestowitz observed, in Trading Places: How We Are Giving Our Future To Japan and How To Reclaim It, that although the United States was once the economic giant of the world, we were rapidly giving this position to Japan. Now we appear to be ceding dominance to China and perhaps India. While Prestowitz placed much of the blame on poor trade policies with Japan and other countries, there were other problems facing corporate America. Peter Senge (1990), Marshall and Tucker (1992), and Peter Drucker (1993) were among the business sages who were saying business will not get better until American corporations become better at learning and better at creating new knowledge. Nonaka and Takiuchi’s (1995) book, The Knowledge Creating Company: How Japanese Companies Create the Dynamics of Innovations, has been “required reading” for many corporate executives. Fortuitously, I had the opportunity in June of 1993 to meet an executive of a major U.S. corporation who was seeking “better tools” to facilitate research and development work. I held my first meeting with a research team at Procter and Gamble in late December, 1993. The consensus was that concept mapping and the ideas I presented could be of value to their corporate research and development work. Another meeting was held in April, 1994, and gradually the number of scheduled meetings increased. I was faced with a difficult decision: Should I continue in the secure position of full professor at Cornell University and pursue only token efforts to apply our ideas in the corporate world, or should I resign from Cornell and free myself to pursue my hunch that a better way for me to improve education in schools and universities may be to improve education and knowledge creation in corporations? I chose in July, 1995 to pursue my hunch. My consulting work with Procter and Gamble increased to nearly full-time during 1995–1998, and we saw that tools and ideas we had developed to improve education could be applied successfully to corporate work in research and

238    J. D. NOVAK

development, understanding administrative problems, developing new marketing strategies, and strategic planning. During the late 1990s I also expanded my work with the Institute for Human and Machine Cognition (IHMC) in Pensacola Florida. With multimillion dollar contracts from the Department of the Navy, NASA, National Security Administration and other federal and private organizations, IHMC was able to enormously improve software called CmapTools that was designed explicitly for creating concept maps of the type we recommended. This software also permits easy collaboration between individuals or groups, efficient capturing and archiving of experts’ knowledge, facilitating individual and group creativity, and many other important knowledge engineering functions. Alberto Cañas and some 40 colleagues created this software largely with grants from taxpayer-supported organizations, so the software has been made available to any user at no cost. It can be downloaded at: http://cmap.ihmc.us. The availability of excellent free concept mapping software has led to a world-wide increase in the use of the concept mapping tool and related ideas. Some 40,000 copies of the software are downloaded from the IHMC servers each month, and unknown numbers of additional copies are downloaded from servers located in schools, corporations and other groups’ servers. Many corporations are now using CmapTools to enhance employee creativity and productivity. While most corporations and other organizations choose not to share their successes with this tool, a growing number are publishing their successes. A recent book by Moon and his colleagues (Moon, Hoffman, Novak, & Cañas, 2011) presents some 30 examples of successful work using CmapTools in corporate settings. Current Status of My Theory of Education All of the above activities have been guided by and have contributed to my evolving theory of education. By 1998, the theory consisted of five fundamental elements: (1) the Learner, (2) the Teacher, (3) Knowledge, (4) Context, and (5) Evaluation. Each of these five elements is involved to some degree in every successful educative event. Each needs to be considered independently, but all interact with each other in the process of an educational event. The nature and function of these elements was described at length in my book Learning, Creating, and Using Knowledge: Concept Maps as Facilitative Tools in Schools and Corporation (Novak, 1998a). Also at that time I summarized the theory with this comprehensive statement:

Half-Century Effort to Create a Theory of Education     239 Meaningful Learning underlies the constructive integration of thinking, feeling and acting leading to empowerment for commitment and responsibility. (Novak, 1998a, p. 15)

Over the years, Ausubel’s theory of meaningful learning has been elaborated by our research group and with contributions from other cognitive psychologists and practitioners. Meaningful learning is truly a profound concept, as are the concepts evolution in biology and entropy in the physical sciences. One can work for a lifetime and still acquire additional and refined meanings for these concepts. After over 50 years of working with the idea of meaningful learning, I still feel that I am refining my understanding of this concept as I continue to study educative events. However, the basic ideas as expressed in my 1998 book cited above (Novak, 1998a) have remained essentially the same, although I hope I have added clarity to these ideas in my latest book (2010). This book spells out how each of the five elements of education operate in effective teaching, management, and learning, illustrating how meaningful learning functions to enhance each element and in turn is maximized when each element is operating optimally. Further discussion of the application of the theory in corporate setting is included in the 2010 edition. A detailed discussion of these ideas is beyond the scope of this chapter. A New Model for Education With the enormously enhanced capabilities of CmapTools that occurred over the past two decades, plus the greatly enhanced availability of high speed Internet and enhanced computer capability at lower cost, it became possible to implement what Alberto Cañas and I call a New Model for Education (Novak & Cañas, 2004). The New Model integrates these learning activities: 1. Use of expert skeleton concept maps to scaffold and guide initial learning and to create Knowledge Models, 2. Use of CmapTools to organize and integrate the complete spectrum of learning activities, 3. Use of collaborative learning facilitated with CmapTools, 4. Extensive use of the Internet and other knowledge sources to be incorporated into knowledge models, 5. Individual and group projects that contribute to building knowledge models, and 6. Archiving knowledge models to scaffold future learning.

240    J. D. NOVAK

Figure 11.6 shows a representation of the New Model. We begin with a Focus Question, “What is required for maintaining good health?” The question is partially answered by the expert skeleton concept map. A team of students then proceeds to engage in some of the other activities shown around the central oval to gather information to be incorporated into the skeleton map. For example, a two to four member team of students might search the Internet for information on the nutritional values of various foods they eat and prepare a table of these values. They might use this information to modify their diets and record the improved diets in a table. Other health related information can also be found on the Internet. The students might interview friends or parents about their nutritional knowledge and show this information in tables, graphs, or as concept maps. Digital recordings of audio or video records of interviews can be added to the evolving knowledge models. Figure 11.7 shows a representation of the evolving model for Good Health, with icons that open some form of digital resource when clicked. CmapTools provides the option of creating files of concept maps and digital resource in this knowledge model; that is, all of these items are connected to a basic concept map, and all of these digital files will be transferred to another server when the knowledge model is transferred. The

Figure 11.6  Schematic illustrating our New Model for education. Beginning with a small expert skeleton concept map, a two to four student team works collaboratively to gather resources of various kinds leading to the construction of a knowledge model for this domain, i.e., “Good Health.” Shown in ovals around the central concept are the various kinds of activities that might contribute to building the knowledge model, as shown in Figure 11.7.

Half-Century Effort to Create a Theory of Education     241

Figure 11.7  A developing Knowledge Model showing added concepts and icons that can be clicked to reveal additional digital resources pertinent to good health.

most comprehensive knowledge model now available for viewing is one created by Cañas and colleagues at NASA. The many digital resources in the file can be seen at: http://spaceexp.ihmc.us. To move educating in the direction of our new model requires that schools or corporations have available high speed Internet access and adequate numbers of computers for each team member to have access for collaborative work on knowledge models. For most schools, this remains a significant limitation at this time, but the picture is improving even in developing countries. For example, Argentina is now in the process of purchasing three million laptops for high schools and providing high speed Internet access. In Panama, we worked with all 4th, 5th, and 6th grade teachers from 1000 schools from 2005 to 2009 to introduce them to CmapTools and meaningful learning strategies. Of course we had varying degrees of success, but all were successful in using computers (most had not done so previously), and many were successful in implementing at least some aspects of our New Model. For example, one of the elementary schools created a knowledge model dealing with the native Cuna Indians, including photos of various aspects of Cuna lives and videos of Cuna dancing. Figure 11.8 shows a knowledge model created by the students, with inserts showing video frames of dancers. Clicking on icons on concepts gives access to video clips and other resources stored in the knowledge model file.

Figure 11.8  A representation from a knowledge model dealing with the Cuna Indians in Panama prepared by a team of fifth grade students.

242    J. D. NOVAK

Half-Century Effort to Create a Theory of Education     243

Promising as the Panama project work was, when a new government took over in 2009, the new government chose to shut down the project, an action that is all to common in Latin countries where political considerations outweigh educational needs. There is much more to be done. At best, a very small fraction of learners in our world are engaged in predominantly meaningful learning practices, whether in schools, universities, or workplaces. While I remain skeptical of data that suggest American students are improving in their “critical thinking skills,” there is a growing recognition that something is wrong with the way many students are being taught and the learning patterns that result from this. See, for example, the many research papers on student learning available via the Internet at http://cmc.ihmc.us. Probably no educator in touch with the research believes that to be effective, all a teacher (or textbook) needs to do is to present the “facts.” The extent to which learner and teacher-held epistemologies influence the quality of learning is still recognized by only a minority of educators, but this situation appears to be improving rapidly. Seldom does a day go by where I do not receive one or more e-mail requests for information on concept mapping or other aspects of our work from researchers and educators in the U.S. and dozens of other countries. So positive changes are occurring. I believe that if corporate America, and the corporate world in general, move to employ new ideas on knowledge creation and knowledge utilization, the entire process may accelerate enormously. For one thing, corporations may have the incentive— good profits at best and survival at the least—to move ahead in applying new educative tools and ideas. They could bring enormous resources and their own examples to bear on these problems—resources measured in the hundreds of billions of dollars! These long-term self-interests require that they help schools and colleges improve their educational practices. These are revolutionary times in the business world, and this will, in due course, require and help to bring about revolutionary changes in education. This, I predict, shall happen in the next 20 years or so—less than half of the time of my career in education. These are exciting times for educators and learners! We cannot, however, rely on corporations to do the basic research needed to understand more effective ways to teach science and to educate science teachers. Most likely, these studies will be done in schools, colleges, and universities. There are now tremendous opportunities for young scholars who choose to pursue such work. It is for these people, especially the “new generation” of scholars, we hope this book can serve as a “handbook” of fundamental ideas and research approaches.

244    J. D. NOVAK

Note 1. Based in part on Novak, J. D. (1998b), The Pursuit of a Dream. In J. J. Mintzes, J. H. Wandersee, and J. D. Novak (Eds), Teaching Science for Understanding and Novak. J. D. (2004), Reflections on a Half Century of Thinking in Science Education and Research: Implications from a Twelve-year Longitudinal Study of Children’s Learning.

References American Association for the Advancement of Science (AAAS). (2011). Benchmarks. Retrieved from http://www.project2061.org/publications/bsl/online/index.php?txtRef=http%3A%2F%2Fwww%2 Ausubel, D. P. (1963). The psychology of meaningful verbal learning. New York, NY: Grune and Stratton. Ausubel, D. P. (1968). Educational psychology: A cognitive view. New York, NY: Holt, Rinehart and Winston. Ausubel, D. P., Novak, J. D., & Hanesian, H. (1978). Educational psychology: A cognitive view (2nd ed). New York, NY: Holt, Rinehart and Winston. Brush, S. G. (1974). Should the history of science be rated X? Science, 183(4130), 1164–1172. Carey, S. (1985). Conceptual change in childhood. Cambridge, MA: The MIT Press. Coffey, J. W., Hoffman, R. R., & Novak, J. D. (2011). Applications of concept maps to web design and web work. In R. W. Proctor & K.-P. L. Vu (Eds.), Handbook of human factors in Web design (2nd ed.; pp. 211–230). New York, NY: Taylor & Francis, CRC Press. Conant, J. B. (1947). On understanding science. New Haven, CT: Yale University Press. Drucker, P. F. (1993). Post-capitalist society. New York, NY: HarperCollins Publishers, Inc. Edmondson, K. M. (1995). Concept mapping for the development of medical curricula. Journal of Research in Science Teaching, 32(7), 777–793. Feyerabend, P. (1988). Against method. London, UK: Verso. Gelman, S. A. (1999). Dialog on early childhood science, mathematics and technology education: A context for learning. Concept Development in Pre-school Children. Retrieved from http://www.project2061.org/publications/earlychild/ online/context/gelman.htm Gowin, D. B. (1970). The structure of knowledge. Educational Theory, 20(4), 319–328. Gowin, D. B. (1981). Educating. Ithaca, NY: Cornell University Press. Keil, F. C. (2011). Science starts early. Science, 331, 1021–1022. Kuhn, T. (1962). The structure of scientific revolutions. International Encyclopedia of Unified Science (2nd ed.). Chicago, IL: University of Chicago Press. Linn, H. (1987). Establishing a research base for science education: Challenges, trends and recommendations. Journal of Research in Science Teaching, 24(3), 191–216. Mager, R. F. (1962). Preparing objectives for programmed instruction. San Francisco, CA: Fearon.

Half-Century Effort to Create a Theory of Education     245 Marshall, R., & Tucker, M. (1992). Thinking for a living: Education and the wealth of nations. New York, NY: Basic Books. Matthews, G. B. (1980). Philosophy & the young child. Cambridge, MA: Harvard University Press. Mayer, R. E. (2004). Should there be a three-strikes rule against pure discovery learning? American Psychologist, 59(1), 14–19. Mazur, J. N. (1989). Using concept maps in therapy with substance abusers in the context of Gowin’s theory of educating Unpublished Master’s thesis, Cornell University, Department of Education, Ithaca, NY. Miller, R. W. (1989). Fact and method: Explanation, confirmation and reality in the natural and the social sciences. Princeton, NJ: Princeton University Press. Mintzes, J., Wandersee, J., & Novak, J. D. (1997). Meaningful learning in science: The human constructivist perspective. In G. D. Phye (Ed.), Handbook of academic learning (pp. x–y). Orlando, FL: Academic Press. Mintzes, J. J., Wandersee, J. H., & Novak, J. D. (1998). Teaching science for understanding. San Diego, CA: Academic Press. Moon, B. M., Hoffman, R. R., Novak, J. D., & Cañas, J. J. (2011). Applied concept mapping: Capturing, analyzing, and organizing knowledge. New York, NY: CRC Press. National Research Council (NRC). (2011). A framework for K-12 science education. Retrieved from http://www7.nationalacademies.org/bose/Standards_Framework_Homepage.html Nonaka, I., & Takiuchi, H. (1995). The knowledge creating company: How Japanese companies create the dynamics of innovations. Oxford, UK: Oxford University Press. Novak, J. D. (1958). An experimental comparison of a conventional and a project centered method of teaching a college general botany course. Journal of Experimental Education, 26(21), 7–230. Novak, J. D. (1961). The use of labeled photomicrographs in teaching college general botany. Science Education, 45(3), 122–131. Novak, J. D. (1963). What should we teach in biology? NABT News and Views, 7(2), 1. Reprinted in Journal of Research in Science Teaching, 1(3), 241–243. Novak, J. D. (1977a). A theory of education. Ithaca, NY: Cornell University Press. Novak, J. D. (1977b). Analternative to Piagetian psychology for science and mathematics education. Science Education, 61(4), 453–477. Novak, J. D. (1991). Clarify with concept maps. The Science Teacher, 58(7), 45–49. Novak, J. D. (1993). Human constructivism: A unification of psychological and epistemological phenomena in meaning making. International Journal of Personal Construct Psychology, 6, 167–193. Novak, J. D. (1995). Concept mapping to facilitate teaching and learning. Prospects, 25(1), 79– 86. Novak, J. D. (1998a). Learning, creating, and using knowledge: Concept maps as facilitative tools in schools and corporations. Mahwah, NJ: Lawrence Erlbaum & Associates. Novak, J, D. (1998b). The pursuit of a dream. In J. J. Mintzes, J. H. Wandersee, & J. D. Novak (Eds.), Teaching science for understanding (pp. 3–27). San Diego, CA. Academic Press. Novak, J. D. (2002). Meaningful learning: the essential factor for conceptual change in limited or appropriate propositional hierarchies (LIPHs) leading to empowerment of learners. Science Education, 86(4), 548–571.

246    J. D. NOVAK Novak. J. D. (2004). Reflections on a half century of thinking in science education and research: Implications from a twelve-year longitudinal study of children’s learning. Canadian Journal of Science, Mathematics, and Technology Education, 4(1), 23–41. Novak, J. D. (2010). Learning, creating, and using knowledge: Concept maps as facilitative tools in schools and corporations (2nd ed.). New York, NY: Routledge, Taylor and Francis. Novak, J. D., & Abrams, R. (Eds.). (1993). Proceedings of the third international seminar on misconceptions and educational strategies in science and mathematics (August 1–4). Published electronically. Retrieved from http://www2.ucsc.edu/mlrg Novak, J. D., & Canas, A. J. (2004, September). Building on new constructivist ideas and the CMap tools to create a new model for education. Closing Lecture, First International Conference on Concept Mapping: Theory, Methodology, Technology, Pamplona, Spain. University of Navarra. Novak, J. D. (in press). Meaningful learning is the foundation for creativity. Quirriculum. Novak, J. D., & Gowin, D. B. (1984). Learning how to learn. Cambridge, UK: Cambridge University Press. Novak, J. D., & Iuli, R. I. (1995). Meaningful learning as the foundation for constructivist epistemology. In F. Finley, D. Allchin, D. Rhees, & F. Fifield (Eds.), Proceedings of the third International History, Philosophy and Science Teaching Conference (Vol. 2, pp.  873–896). Minneapolis, MN: University of Minnesota. Novak, J. D., Meister, M., Knox, W. W., & Sullivan, D. W. (1966). The world of science series (Books One through Six). Indianapolis, IN: Bobbs-Merrill. Novak, J. D., & Musonda, D. (1991). A twelve-year longitudinal study of science concept learning. American Educational Research Journal, 28(1) , 117–153. Novak, J. D., & Ridley, D. R. (1988). Assessing student learning in light of how students learn. Washington, DC: The American Association for Higher Education Assessment Forum. Novak, J. D., & Wandersee, J. (1990). Perspectives on concept mapping. Journal of Research in Science Teaching, 27(10). Novak, J. D., Gowin, D. B., & Johansen, G. T. (1983). The use of concept mapping and knowledge Vee mapping with junior high school science students. Science Education, 67(5), 625–645. Piaget, J. (1926). The language and thought of the child. New York, NY: Harcourt Brace. Postlethwait, S. N., Novak, J. D., & Murray, H. (1964). The integrated experience approach to teaching botany. Minneapolis, MN: Burgess. Postlethwait, S. N., Novak, J. D., & Murray, H. (1972). The audio-tutorial approach to learning through independent study and integrated experience (3rd ed.). Minneapolis, MN: Burgess Publishing Company. Prestowitz, C. V., Jr. (1988). Trading places: How we are giving our future to Japan and how to reclaim it. New York, NY: Basic Books. Schneps, M. (1989). Private Universe Project. Cambridge, MA: Harvard University. Senge, P. M. (1990). The fifth discipline: The art & practice of the learning organization. New York, NY: DoubIeday. Shayer, M., & Adey, P. (1981). Towards a science of science teaching: Cognitive development and curriculum demand. London, UK: Heinemann Educational Books.

Half-Century Effort to Create a Theory of Education     247 Symington, D., & Novak, J. D. (1982). Teaching children how to learn. The Educational Magazine, 39(5), 13–16. Toulmin, S. (1972). Human understanding, Vol. 1: The collective use and evolution of concepts. Princeton, NJ: Princeton University Press. von Glasersfeld, E. (1984). An introduction to radical constructivism. In P. Waxlawick (Ed.), The invented reality (pp. 17–40). New York, NY: Norton. Vygotsky, L. S. (1962). Thought and language. E. Hanfmann (Ed.) and G. Vakar (Trans.). Cambridge, MA: MIT Press (Original work published 1926) Wandersee, J . H., Mintzes, J. J., & Novak, J. D. (1994). Learning: Alternative conceptions. In D. L. Gabel (Ed.), Handbook on research in science teaching (pp. 177– 210). A project of the National Science Teachers Association. New York, NY: Macmillan. Wiener, N. (1948). Cybernetics. New York, NY: Wiley. Wiener, N. (1954). The human use of human beings (2nd ed.). Garden City, NY: Doubleday.

248    J. D. NOVAK

Chapter 12

Pinchas (Pini) Tamir A Long-Distance Runner Across and Beyond Science Education Avi Hofstein The Weizmann Institute of Science, Rehovot, Israel Hanna J. Arzi Independent Scholar, Tel Aviv, Israel Anat Zohar The Hebrew University of Jerusalem, Israel

How do we start and structure an essay on Pinchas Tamir’s career? Tamir (or Pini, as his colleagues and students alike call him) did it all: He was a school teacher, curriculum developer, teacher educator, leader of systemwide educational change, and an incredibly productive researcher. His versatile activities were intertwined, feeding and extending each other, reflecting his wide perspective and integrative approach to research and practice. Most of Tamir’s studies and projects were carried out in the context of biology education in Israel and are rooted in his previous school teaching of

Going Back for Our Future, pages 249–267 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

249

250   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

biology and horticulture. Yet, as we describe and discuss below, the impact of his work has crossed boundaries, reaching other science disciplines, international science education, and the wider arena of education. The quantity of Tamir’s work is particularly amazing in view of the fact that he embarked on doctoral studies after 15 years of high school teaching and got his first academic appointment at the age of 40; furthermore, a cruel disease that he fought courageously forced him to slow down earlier than he had wished. The quality and versatility of his academic work can be exemplified through major honors conferred upon him by prestigious professional associations who attended to different aspects of his research: the National Association for Research in Science Teaching (NARST) recognized in 1992 his lifetime achievements by the highest Award for Distinguished Contribution to Science Education Through Research, while the American Educational Research Association (AERA) gave him in 1977 the Palmer O. Johnson Memorial Award for his work on cognitive preferences, which is not specifically related to science education. It is noteworthy that these two prestigious awards are given by American-based organizations and in both Tamir was the first non-American recipient. The versatility of Tamir’s career is also reflected by the backgrounds of the authors of this chapter. We come from both the physical and the life sciences, our interests and career tracks vary, and each one’s acquaintance with Tamir occurred in a different context. While all of us have felt privileged for the opportunity to write about Tamir, the quantity, quality, and intertwined nature of his work posed a challenge, which brings us back to the opening question above: How to structure an essay that would convey a whole picture? Neither a linear chronological career account nor a topic by topic description seems appropriate. In an attempt to weave all, we provide a frame of dates and milestones in Tamir’s life (see Appendix) and will go back and forth between a chronological order and a presentation of pioneering research and development contributions. From Healing Plant Wounds, Through Teaching, to Science Education Having wondered how to tell Tamir’s story in science education, perhaps a way to start is through quoting his own opening words of a research-oriented autobiographical account written at the middle of his academic career (Tamir, 1985a): My first experience in doing research is associated with my Master’s thesis as a student of horticulture. My research problem was to find or develop an improved wound dressing which would facilitate the healing of pruning wounds.

Pinchas (Pini) Tamir    251 I actually invented an improved dressing using chalk, oil, bee wax and plant hormones. However, from 1951, the year in which I completed my MSc. thesis at the Hebrew University, up to 1966, the year I started my Ph.D. studies in science education at Cornell University, U.S.A., I did not do any systematic research, and consequently, I did not publish any research papers (other than three papers based on my MSc. dissertation). Although I did not do systematic publishable research in both of my occupations, namely teaching biology and horticulture to high school students and managing the fruit farm of a secondary agricultural school, I have been open to experiments and to innovative approaches. My interest in trying new approaches led me to become a member of the selected group of high school biology teachers who adapted the innovative American BSCS (Biological Science Curriculum Study) program for use in Israeli schools. When in 1966 I was awarded with a scholarship for doctoral studies in the U.S.A., it was only natural that my Ph.D. dissertation would focus on long-term effects of high school biology with special reference to the BSCS program. Ever since I received my Ph.D. in 1968, the study of high school biology has been a major focus of my research. . . . Major topics of my research have been: teaching and learning in the school laboratory; facilitating meaningful and long-lasting learning, as well as preferences and interests of students and teachers. (pp. 1–2)

The major topics highlighted in the 1985 autobiographical notes reemerged years later in Tamir’s self-choice of his most significant published work. This choice was made in response to Peter Fensham’s (2004) request to do so as part of a worldwide interview study with science educators on the evolution of science education as a field of research. It is beyond the scope of our chapter to cover all of Tamir’s accomplishments. Thus we chose to focus on selected aspects of areas that he himself had highlighted, as noted above, mainly (a) on the holistic approach in his work on high school biology that formed the basis for all his contributions in science education and beyond, and (b) on the making of inquiry and assessment integral parts of teaching and learning in the school laboratory. With the benefit of a time perspective, we will draw patterns across Tamir’s work that does or does not subscribe to strict definitions of “research,” as well as across non-publishable achievements, including curriculum development, teacher education, reform leadership, synthesis and dissemination of knowledge. This entire work was based on, or driven by, research, often generated new research, and always linked to practice. But before we delve into Tamir’s work as a science educator, we should have a glimpse into the preceding 15 years of high-school teaching—an invaluable experience that underlay his perspective on education and nurtured his thinking throughout his career.

252   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

Where It All Started: Tamir—the High School Teacher One of us, Avi Hofstein, is in a unique position to provide first-hand evidence of Tamir’s years as a high school teacher, since he has personal recollections from two perspectives: of a high school student and of a staffroom colleague. In his reflections, Hofstein adds his current perspective as a science educator and juxtaposes traditional school teaching that prevailed in the 1950s–1960s with how his teacher actually did it very differently: In the late 1950s Pini was my biology teacher in the Pardes Hanna Agricultural High School. Later on, after graduating from the Hebrew University of Jerusalem when I came back to the same school as a chemistry teacher, he became my colleague and model teacher. There is no doubt that as a biology teacher he instilled in me the love for nature, environment, agriculture (especially horticulture), and practical work mainly conducted as studentcentered inquiry. In those days, the biology taught in high school was characterized by collection of facts and phenomena, and memorizing names of specimens, animals, and plants was one of the main skills that students had to learn and develop. The laboratory was mainly confirmatory in nature, demonstrating the facts and principles that were taught in the biology classroom. Resulting from that, the assessment of students’ progress and achievements by their respective teachers, as well as in final examinations (matriculation) administered in Israel by the Ministry of Education, were highly based on memorizing facts and not on what is called today high level thinking skills or inquiry. Many issues related to the teaching and learning of science that are cornerstones in most of the current science curricula, such as issues related to the nature of science, were not included in the science curricula that prevailed in Israel in general and in biology teaching and learning in particular. Pini, a pioneering teacher, was ahead of his time and clearly one could say that he was a reflective practitioner and open minded to pedagogical approaches with the goal of making the learning of biology more interesting and motivating to most of the students. These included project-based learning (long before the acronym PBL was established), field trips, integration of subjects, and critical thinking. I had first experienced these approaches as Pini’s student. Later on, when I started teaching chemistry and became Pini’s colleague, he introduced me to these ideas from a teacher’s perspective. I recall vividly project-based learning related to the topic Blood in Human Beings, in which chemistry and biology were taught by the two of us in an integrated approach. Once again, in the early 1960s this was an unusual approach in science classrooms.

Pinchas (Pini) Tamir    253

Tamir’s Holistic Approach in His Tour de Force in Biology Education in Israel The reform in science education in the U.S. during the 1950s and 1960s, fondly called “the golden age of science education” despite hindsight criticism, echoed also in Israel across all school science subjects. Tamir was then a leading figure among a group of biology teachers who were dissatisfied with the existing programs that lagged far behind the progress in science and related fields, such as agriculture and technology, and were in need of reform in both content and pedagogy. Eventually, these teachers, together with university scientists, education experts, school administrators, and government officials, were involved in a long-term iterative process of development and implementation of a new biology curriculum in Israel. While clearly building on the earlier emergence of the ground-breaking American curricula during the 1950s and 1960s across all science subjects, the story of school biology in Israel was different. The most salient characteristic of this reform was its holistic approach that viewed curriculum development, implementation, and teacher professional development as one interactive, long-term, and coherent process that involved collaboration of multiple actors, including teacher participation from the very beginning. Tamir had a key role in this unique approach, as we highlight below. Our description is based on articles that were published over the years on different aspects in different stages since the 1960s. We draw largely on the more recent articles by Tamir (2004, 2006). The reform in biology education in Israel actually started in 1964 with the decision to build on an existing exemplary curriculum—the inquiryoriented “Yellow” version of BSCS (1963)—and to adapt it to the context of the Israeli school system. It was also decided to do the adaptation through full collaboration of teachers with academic scientists and teacher educators. The leading figure in the establishing team was Alexandra Poljakoff— a prominent plant physiologist from the Hebrew University of Jerusalem who was deeply interested in education (Mayer, 1999). Tamir was in the first group of 25 selected teachers and his contributions to the collaborative efforts were recognized and appreciated. And thus, when he returned from his doctoral studies at Cornell University, Tamir was invited by Poljakoff to assume the leading position in the process of changing biology education to an inquiry-based approach, emphasizing scientific reasoning and meaningful learning. Formally, in 1968 Tamir became the director of the BSCS Adaptation Project in Israel, also known as the Israel High School Biology Project. Since that time until retirement in 1996, his fingerprints left strong marks on the Israeli biology curriculum. Such sustainability is uncommon in attempts at system-wide educational reforms. It is therefore worthwhile to

254   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

take a deeper look at the implementation process in order to learn from it about the nature of that process and the person who had led it. Right from the beginning, it was clear to the group of educational innovators led by Tamir that the way to create long-lasting educational change must be holistic. Rather than sufficing with the translation and adaptation of a textbook that was successful in another country (i.e., the BSCS Yellow version), it was assumed that the route to success must pass through working simultaneously in all areas, mainly curriculum materials, professional development, and assessment. As already mentioned, the development of curriculum materials had started in 1964 and initially involved 25 teachers who learned, wrote, trialed in their classrooms, revised, and instructed new teachers to join in gradually. When Tamir’s status changed from team member to team leader, the initial production of materials had been practically completed and the balance of the simultaneous work shifted to what we now refer to as implementation. It is noteworthy that the term implementation appeared in the literature around 1970 and became common only after the related review by Fullan and Pomfret had been published in 1977. Hence, Tamir’s focus on implementation procedures was not self-evident. Within implementation, professional development included integration of the ideas and methods of the new curriculum into pre-service programs, and intensive in-service teacher learning that took place all over the country in recurring one-day conferences and in longer courses during school vacations. Special care was given to the provision of technical conditions required for smooth implementation, including upgrading of school laboratories, establishing a center for supply of laboratory consumables and biological materials, securing an appropriate budget, and more. Harnessing the Power of Assessment to Reform Most of the instructional techniques found in the Hebrew adaptation of the BSCS program can be found in inquiry-oriented curriculum materials in other countries, and there are examples of curriculum reforms accompanied by changes in assessment intended to achieve content consistency. However, as part of the change process in Israel, the biology team under Tamir’s leadership made radical changes in the methods of assessment that were innovative in the 1970s. They understood that a lack of consistency between the inquiry goals of the new curriculum and the assessment will put the reform effort at risk, because teachers will continue to “teach for the test” in terms of content, rather than follow the inquiry-oriented learning environment. Instead of fighting the existence of the biology matriculation examination taken at the end of high school (Bagrut, in Hebrew), the examination was transformed to reflect the goals of the new inquiry-based cur-

Pinchas (Pini) Tamir    255

riculum. In effect, Tamir and his team changed the traditional examination into what by today’s terminology can be defined as “alternative assessment.” The rationale for, structure of, and research on, this examination was described over the years by Tamir and his colleagues (e.g., Tamir, 1974, 1985b, 2006). In the following paragraphs we provide a brief description of the main inquiry-oriented components as they were constructed in the 1970s. The new matriculation examination in the 1970s consisted of varied means of assessment: a written test, a school-based ecological research project accompanied by an oral test, and a school-based laboratory test. Since the examination was intended to reflect faithfully the goals of the new curriculum, assessing students’ inquiry skills was an important component across all parts of the examination: 1. In the written test, one section consisted of multiple-choice items assessing factual knowledge and basic understanding, while other sections addressed higher order thinking skills (through constructed responses requiring explanations and justifications) and inquiry skills (through a task presenting a previously unseen scientific research that students were required to analyze in a critical way). 2. The ecological project was inquiry-based, students had to summarize their work in writing and to defend it in an oral test. 3. The laboratory test presented a scientific phenomenon that students had not encountered as part of their biology studies. Students were expected to formulate a research problem related to the phenomenon, to raise hypotheses, to plan an investigation, to carry out an investigation using equipment and materials that had been prepared for them in the school laboratory, to record and report their results using graphs and tables, to analyze the results, to draw conclusions, and to suggest implications and applications. The test took place in schools all over the country on the same day. Ten percent of the grade was given at the time of the test for the quality of students’ technical work. The remaining 90% assessed inquiry through the quality of students’ open-ended constructed responses to the laboratory-based test questions, using a specially-designed instrument that is close to what is currently called a “rubric.” The laboratory test was the main innovation of the assessment framework, which in subsequent years served as a model for the development of alternative assessments worldwide. Developing that type of assessment in the early 1970s, before the proliferation of the literature concerning new assessment types—such as performance assessment, qualitative assessment, or rubrics—had obviously required deep understanding across all areas of education, a vision, and a creative mind. Implementing that type of assess-

256   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

ment on a national scale had required acquaintance with the system from the teacher level all the way to government officials, persistent hard work, leadership, and charisma. The engagement in, and study of, this process positioned Tamir as an influential international expert in the fields of assessment and of laboratory-based inquiry learning. We revisit Tamir’s ideas on these issues later in this chapter. Reform Sustainability In general, it is difficult to implement educational reforms that pertain to deep changes in the quality of instruction and learning; long-term sustainability of such large-scale changes is even more difficult (Arzi, 2012). It is therefore remarkable that many of the main facets of the inquiry-based biology curriculum reform in Israel are still practiced 40 years after it was initiated. Let us point to two pieces of evidence demonstrating the longterm influence that Tamir’s work had on biology education specifically and on education in general within Israel. The first piece of evidence is taken from a study on science teachers’ pedagogical knowledge in the context of teaching higher order thinking (Zohar & Schwartzer, 2005). This study compared the pedagogical knowledge of biology, chemistry, and physics teachers and found that the knowledge of the biology teachers was significantly higher than that of both the physics and the chemistry teachers. These findings are explained by the long-term tendency in Israeli schools, beginning with the reform described earlier, to teach biology with a strong emphasis on inquiry, whereas such a reform had not yet taken place in physics and chemistry when the study was conducted in 2002. The second example of the long-term influences of Tamir’s work refers to the Israeli educational system at large, not just in biology. In 2007, the Israeli Ministry of Education formulated a national policy aimed at fostering higher order thinking across the curriculum (“Pedagogical Horizon”—Teaching for Thinking). The model for implementing this policy was strongly influenced by the successful implementation process of the biology inquiry-based curriculum that took place years earlier and followed a similar design of introducing simultaneous changes to curriculum development, professional development, and assessment (Zohar, 2008). Laboratory and Inquiry: The Heart of Tamir’s Thoughts and Actions The chapter on science education in AERA’s Second Handbook of Research on Teaching, published in 1973, was written by Lee Shulman and Pinchas Tamir

Pinchas (Pini) Tamir    257

and drew much attention within and beyond science education. This truly seminal review saw the school science laboratory as a distinctive feature of science education: The laboratory has always been the most distinctive feature of science instruction. . . . With the shift of emphasis from acquisition of knowledge to other objectives which stress the processes of science, the laboratory acquired a central role, not as a means for demonstration and confirmation but rather as the core of the science learning process. (pp. 1118–1119)

Having a deep belief in the centrality of the laboratory, Tamir in collaboration with his colleagues and students researched laboratory instruction comprehensively and intensively. This acclaimed work was carried out mainly in the context of teaching and learning biology in Israel, as described earlier. Why was Tamir’s work on the school laboratory so important? A partial answer is provided in the paper How Are the Laboratories Used? (Tamir, 1977) in which he questioned the match between the goals of the curriculum developers who design the laboratory activities, and the teachers’ classroom laboratory practices and interactions, mainly in regard to inquiry-type experiences. Most of Tamir’s work on this issue is described in the much cited review written in collaboration with Reuven Lazarowitz, a leading science educator from the Technion—Israel Institute of Technology (Lazarowitz & Tamir, 1994). In this review, they suggested that the laboratory is the only place in which students are provided with the opportunity to do science. They also posited that the laboratory, if designed properly, is the only place in school where certain kinds of important learning skills and understandings can be developed. In addition, they claimed that the science laboratory is a unique learning environment with unique pedagogical interventions and related assessment methods. This claim is highly based on Tamir’s work conducted in the 1970s (Tamir, 1972, 1974, 1975b) in the context of the development and implementation of the adapted version of the BSCS curriculum. Based on these studies, he managed to present clear evidence that three distinct modes of performance can be identified in biology teaching and learning: analytical, constructive, and practical. Tamir and his research group developed a form of practical test in a laboratory setting that measures the practical mode, which cannot be assessed by traditional paper-and-pencil tests. Subsequently, the inquiryoriented laboratory test was implemented in Israel as part of the biology matriculation examination mentioned earlier in this chapter. In a major article on this test that presents detailed descriptions of the test construction and administration along with data analysis, Tamir (1974) argued that this type of test reflects the inquiry objectives of the BSCS philosophy and goals, providing a valid and reliable measure of problem solving ability in a practical laboratory learning environment. These developments and re-

258   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

search responded to the claim made at that time (e.g., Grobman, 1970) that measurement of practical abilities by teachers, as well as by researchers, is difficult: It is, indeed, difficult, but Tamir showed that it is possible. Training Teachers to Teach Effectively in the Laboratory Among the key obstacles for effective implementation of such a demanding instructional change are the science teachers’ beliefs, classroom laboratory practices, and behaviors. In his paper Training Teachers to Teach Effectively in the Laboratory, Tamir (1989) cited the claim made earlier by Hofstein and Lunetta (1982) that studying science through laboratory-oriented programs fell short of expectations, and argued that “There is no doubt that improvement in the effectiveness of learning in the laboratory can be achieved only through substantial improvement in teacher preparation” (Tamir, 1989, p. 60). In this award-winning paper (see awards in the Appendix) he suggested several practical ideas for actually doing so within both the pre-service and in-service stages of teacher development. Among these were the following: • training teachers to analyze the content and skills that are developed in the laboratory by using the Laboratory Analysis Inventory (LAI) developed by Tamir and Lunetta (1981); and • training teachers to be flexible regarding the classroom laboratory learning environment, namely, to align the work with students’ abilities and characteristics and the nature of the experiments (e.g., grouping, individualizing, whole class activities). As part of their attempts to enhance laboratory instruction, throughout the years Tamir and his group developed reliable and valid schemes for assessing students’ achievement and progress in the biology laboratory (for detailed descriptions see, e.g., Tamir, 1974; Tamir, Nussinovitz, & Friedler, 1982). These schemes have been actually used successfully by teachers in their classroom laboratories as well as in the final matriculation examinations designed and administered centrally in the Israeli education system. Before concluding this section on Tamir’s work concerning the practical mode, we wish to note that his underlying work included all the relevant facets of a typical curricular cycle: intensive research into the nature of practical work, development of practical inquiry-type experiences, implementation in the biology classroom with major attention to teacher professional enhancement, and finally, development of usable and valid assessment tools. This

Pinchas (Pini) Tamir    259

reflects the holistic long-term approach that cuts across the research and development projects in which Tamir was extensively involved over the years. Beyond Reforming High School Biology and the Science Laboratory As we have described and discussed above, high school biology was Tamir’s main arena, with inquiry learning in the school laboratory at the center stage. The impact of this work has spread from Israel overseas, across the school science subjects and well beyond them. The prime example for the wide ranging impact is the spread of innovative methods that he devised with his colleagues and students for the evaluation of laboratory work in biology. We have already noted that it happened before the labels “alternative assessment” and “performance assessment” became common in the educational jargon, hence this was clearly a pioneering contribution. The following words by Stanford University professors Richard Shavelson and Lee Shulman attest that the impact was not confined to the school biology laboratory, nor to science education: Pini’s work on Israel’s biology matriculation examination was path breaking. . . . It still is the standard in my thinking. Both the performance tasks during the examination and the semester-long biology project . . . were creative and way ahead of their time. His work . . . has had broad impact far beyond science education. Indeed, I can trace his influence to today. (R. J. Shavelson, personal communication, November 17, 2010) How did one compete with the conservative power of the Bagrut [Israel matriculation examination] if one wished to introduce new approaches? Pini accomplished the impossible. . . .  This work not only had a significant impact on science education in Israel. It became well known all over the world and contributed to the development of “performance assessments” in science and other domains of the curriculum. Moreover, it taught reformers in other countries, myself among them, of the leverage to be gained by the re-invention of high-stakes assessments rather than always focusing on changing the curriculum directly. (L. S. Shulman, personal communication, November 11, 2010)

Life-Long Learning as Research Feeder and Generator From time to time, Tamir was intrigued by a topic that was not an integral part of his long-term research programs. As always, he explored his new interests seriously and produced high-quality work. We will present two examples for Tamir’s “diversions” from the core of his work that were offsprings of his curiosity and lifelong acquisition and integration of new

260   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

knowledge whenever an opportunity to learn crossed his way. The first example is within biology and emerged from his regular reading and amazing mastery of the literature, while the context of the second example is different. Our first example is linked to the paper entitled, When Is an Animal Not an Animal? by Beverly Bell (1981) that explored the concepts of “animal” held by students. Over the years, this paper has been much cited in the misconceptions literature; however, when it was published in the Journal of Biological Education, the author was still a graduate student in New Zealand, and this was her first article in an American journal. Tamir, an avid and reflective reader, immediately noticed the quality and potential of the study and wrote a letter to the journal, commending the article, raising some criticism, and concluding: “The article prompted me to look into the ‘animal issue’ more deeply in the Israeli context since we have two partially overlapping terms for an animal which are commonly used and may cause even greater confusion” (Tamir, 1982, p. 161). And thus, apart from the core issues of inquiry and reform in biology education to which we attended above, Tamir also explored and published on misconceptions, as well as other aspects of biology education, including student interest in studying plants and animals, problems of using animals in laboratory classes, and more. In the next section we exemplify Tamir’s lifelong curiosity and learning through his work on a subject not specifically related to biology that eventually established him as a world leader in yet another area. Cognitive Preferences and International Influences Tamir was a globetrotter, invited to speak from one research center to another and attending many conferences. These travels made his work known worldwide, but no less than that—he received feedback and learned all the time, to the extent that he described the impact of his international relations on his research as “enormous.” A major example is the notion of cognitive preferences to which he was introduced by Richard Kempa from the University of Keele: I was intrigued to cognitive preferences because this was a real novel idea . . . and because the rationale behind it matched my personal beliefs. The idea that learning science in a particular manner will not only affect achievement in terms of knowledge and skills, but can actually change the style and approach of a person . . . provoked my interest. (Tamir, 1985a, p. 4)

Tamir started studying cognitive preferences in 1973, and this subject occupied a significant part of his research endeavors through the 1980s. One of his early papers on cognitive preferences (Tamir, 1975a) received a pres-

Pinchas (Pini) Tamir    261

tigious AERA award (see Appendix), which enhanced the esteem for his scholarship outside the realm of science education. Based on Tamir’s writings and our personal conversations with him over the years, there were many more influences by colleagues and students in Israel and overseas; only a few are mentioned in this chapter. Standing out is the encounter with a career-long impact that occurred in 1967, in the midst of his doctoral studies, as the young scholar Joseph Novak had just arrived to assume an appointment at Cornell University. Tamir was introduced by Novak to Ausubel’s then recently published book (Ausubel, 1968); the ideas captivated him, he immediately applied them in his thesis, and he continued applying them and exploring meaningful long-term learning and retention ever since. Novak recalls: When I arrived at Cornell University in 1967, Pini was working on his PhD thesis. He sat in on my seminar on Ausubel’s cognitive theory of learning and immediately saw how the theory, that was new to him, could help to interpret the data he had gathered for his thesis. This was illustrative of Pini’s ability to quickly grasp new ideas and see how they could be used to improve science education and science education research. (J. D. Novak, personal communication, November 17, 2010)

Decades later, in the interview series conducted by Fensham (2004) with leading science educators, Tamir pointed to Ausubel’s 1968 book as the single publication that had been most influential on his career. Shortly after recognizing the potential of Ausubel’s ideas for science education, Tamir’s appreciation of the contribution of educational psychology to science education widened through another significant encounter—with Lee Shulman. In the next section we will take up this encounter. Communal Scholarship and Knowledge Synthesis Throughout the second half of the twentieth century Tamir continually carried out excellent research. He was certainly one of the best-known scholars in science education for the whole of that period. . . . There was not just high scholarship, demonstrated by his long list of research publications in major journals and numerous invitations to speak at universities and forums in diverse countries, but also his readiness to support other researchers in the best tradition of communal scholarship. (R. T. White, personal communication, November 19, 2010)

Communal scholarship, as noted above by Richard White from Monash University, a leading science educator who had known Tamir for many years, was a major component along Tamir’s life—not just learning but also contributing whenever and wherever possible.

262   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

The seminal review of research on science teaching and learning in AERA’s Second Handbook of Research on Teaching, to which we attended earlier (Shulman & Tamir, 1973), is a special case of communal scholarship, contributing to the research community and to each of its authors. Through creative synthesis of research that crossed the boundaries of science education, it advanced science education and education at large, and, at the same time, enabled its authors insights that underlay their own future work. It exemplifies synergic collaboration between two top scholars with complementing solid backgrounds: Shulman in philosophy and psychology, Tamir in science and science teaching. Tamir (1985a) said of his encounter with Shulman: “From my friend Lee Shulman I have acquired the inspiration and guidance to view broadly the interrelations between science education on the one hand and educational psychology and philosophy on the other hand” (p. 5). Years later, Shulman reflected: I first encountered Pini Tamir in 1969, when I arrived in Israel for a sabbatical leave as a visiting professor in the School of Education and the School of Medicine of the Hebrew University. He had just returned from completing his doctoral studies at Cornell University (this after a solid career as a practicing high school biology teacher), and I was so impressed with his talents that I invited him to co-author the chapter on “Research on Teaching in the Natural Sciences” for the Handbook of Research on Teaching. This chapter, primarily because of Pini’s contributions, became one of the central references in the field of science teaching for the next two decades. (L. S. Shulman, personal communication, November 11, 2010)

Four decades beyond its publication, this chapter is not a central reference as it used to be, but its importance persists, as noted by Shavelson: The Handbook chapter gave form to the field of science education, a field needing form in 1973 when the chapter was published. Importantly it clarified at the time debates about so-called inquiry teaching and other forms of teaching (discovery and what is now called “direct instruction”). The discussion has now proved to be extremely important as issues about formative assessment and learning progressions in inquiry teaching are front and center on the international science education stage. (R. J. Shavelson, personal communication, November 17, 2010)

Fruitful collaboration in research synthesis led to yet another exemplary product to which we have already attended: the chapter on the use of laboratory instruction, written by Lazarowitz and Tamir (1994) for the National Science Teachers Association (NSTA) Handbook of Research on Science Teaching and Learning. The collaboration between Lazarowitz and Tamir can be traced back to the 1960s when both were teachers in small localities in the north of Israel, trying to change tradition in biology class-

Pinchas (Pini) Tamir    263

rooms through Tamir’s local initiatives of teacher collaborations. Reflecting through the very long list of publications, Tamir chose to rate the two co-authored handbook chapters—with Shulman in 1973 and with Lazarowitz in 1994—among his four most significant publications (reported in Fensham, 2004). Tamir’s Scholarship as Quantity and Quality Tamir was an incredibly productive scholar during three decades of full academic activity (less so over the subsequent decade due to ailing health). Our notion of productivity refers to both quantity and quality. If we use the conventional academic measure of publication numbers, it adds up to over 200 articles in internationally refereed journals, over 300 if book chapters, books and other publications in both English and Hebrew are also counted, and very many more if the list also includes presentations in universities and conferences around the globe. Productivity as quality, however, is not as tangible as paper counting, but it can be inferred from the prestige of the hosting journals, publishers, awards, and inviting institutions. A major facet of quality is the pioneering nature of the work which we attempted to convey in our presentations of major themes. When the science curriculum reform era started in the 1950s and 1960s, Tamir was a school teacher; later he became the major driving force in biology education in Israel. Tamir’s work in school biology research, policy, and practice can be seen as both ends and means. It was primarily intended to push forward a comprehensive reform of biology education in Israel, and indeed it did so with sustainable long-term success, which subsequently affected biology education internationally. At the same time, the work in school biology served as a platform for research-based pioneering contributions well beyond it, particularly so in regard to the advancement of meaningful learning in the school science laboratory and inquiry-oriented instruction at large, performance assessment, and the harnessing of much maligned high-stakes tests in the implementation of reforms. Needless to say, not all could be conveyed in a single length-restricted chapter, and we urge readers to learn and form their own views by actually studying the original materials. We wish to conclude by drawing attention to time as a measure of quality across Tamir’s career: He was a lifelong learner; his development, policy, and research efforts on and off stage were continuous and long term, all eventually leading to unusually sustainable impact. Tamir can be therefore described as a long-distance runner in science education, not a track sprinter, but a road and hurdle long-distance racer who paved the way for others whom he trained to join in.

264   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR

Appendix Pinchas Tamir: A Few Dates and Milestones • Born in 1928 in the city of Tel Aviv where he completed primary school and started high school; at his request he moved to the rural Pardes Hanna Agricultural High School that was known for high academic achievements; graduated in 1945. In 1951 he returned to the same school for 15 more years, with a double appointment: teacher of biology and horticulture, and manager of the fruit farm.

• Immediately following high school, in 1945–1946 he volunteered to the Jewish Brigade that had been part of the British Army in World War II, and was deployed to Egypt and to Italy. He later served in the Israel Defense Forces. During military service he met Ruth Shomroni; they married and raised a family. • Studied at the Hebrew University, MSc in Agriculture received in 1951. • As a high school teacher, 1951–1966, he made innovations in instructional strategies, school-based curriculum, and teacher collaborations, and was an active member of the first team that in 1964 embarked upon the adaptation of the American program BSCS to the Israeli context, thus initiating the major reform in secondary school biology in Israel. In subsequent years, his academic R & D were invaluable to the longterm success of this reform, with consequent contributions to international biology education, science education, and education at large.

• A scholarship enabled him to travel to the U.S. for doctoral studies at Cornell University; in 1968 he received a PhD in science education and returned to Israel. • The Hebrew University of Jerusalem, his Alma Mater, was his academic home base from the first academic appointment in 1968 until official retirement in1996, and henceforth as Professor Emeritus. • Major awards include: –– 1977, by AERA (American Educational Research Association), Palmer O. Johnson Award, for his paper on cognitive preferences, published in the American Educational Research Journal (Tamir, 1975a); –– 1987, by AETS (Association for the Education of Teachers in Science, currently ASTE, Association for Science Teacher Education), Implications of Research for Educational Practice Award, for his paper on training teachers to teach effectively in the laboratory, published in Science Education (Tamir, 1989);

Pinchas (Pini) Tamir    265

–– 1992, by NARST (National Association for Research in Science Teaching), Distinguished Contribution to Science Education Through Research Award. He was the seventh recipient, first non-American, of this coveted award, probably the highest international honor for academic work in science education; –– 1998, by AETS/ASTE, Outstanding Longtime Service Award. • Died in 2012 at his home in Jerusalem. References Arzi, H. J. (2012). Change—a desired permanent state in science education. In B. J. Fraser, K. G. Tobin, & C. J. McRobbie (Eds.), Second international handbook of science education (pp. 883–898). Dordrecht, The Netherlands: Springer. Ausubel, D. P. (1968). Educational psychology: A cognitive view. New York, NY: Holt, Rinehart & Winston. Bell, B. F. (1981) When is an animal not an animal? Journal of Biological Education, 15, 213–218. Biological Science Curriculum Study (BSCS). (1963). Biological science: An inquiry into life (“Yellow” version). New York, NY: Harcourt, Brace & World.. Fensham, P. J. (2004). Defining an identity: The evolution of science education as a field of research. Dordrecht, The Netherlands: Kluwer. Fullan, M., & Pomfret, A. (1977). Research on curriculum and instruction implementation. Review of Educational Research, 47, 335–397. Grobman, H. (1970). Developmental curriculum projects: Decision points and processes. Itasca, IL: Peacock Publishers. Hofstein, A., & Lunetta, V. N. (1982). The role of the laboratory in science teaching. Review of Educational Research, 52, 201–217. Lazarowitz, R., & Tamir, P. (1994). Research on using laboratory instruction in science. In D. L. Gabel (Ed.), Handbook of research on science teaching and learning (pp. 94–128). New York, NY: Macmillan. Mayer, A. M. (1999). Alexandra Poljakoff-Mayber [In memoriam]. Seed Science Research, 9, 263–264. Shulman, L. S., & Tamir, P. (1973). Research on teaching in the natural sciences. In R. M. W. Travers (Ed.), Second handbook of research on teaching (pp. 1098–1140). Chicago, IL: Rand McNally. Tamir, P. (1972). The practical mode—a distinct mode of performance in biology. Journal of Biological Education, 6, 175–182. Tamir, P. (1974). An inquiry oriented laboratory examination. Journal of Educational Measurement, 11, 25–33. Tamir, P. (1975a). The relationship among cognitive preference, school environment, teachers’ curricular bias, curriculum, and subject matter. American Educational Research Journal, 12, 235–264. Tamir, P. (1975b). Nurturing the practical mode in schools. The School Review, 83, 499–506.

266   A. HOFSTEIN, H. J. ARZI, and A. ZOHAR Tamir, P. (1977). How are the laboratories used? Journal of Research in Science Teaching, 14, 311–316. Tamir, P. (1982). When is an animal not an animal? [Letter]. Journal of Biological Education, 16, 161. Tamir, P. (1985a, March). An autobiographical account focusing on research. Paper presented in a symposium at the annual meeting of the American Educational Research Association, Chicago, IL. Tamir, P. (1985b). The Israeli “Bagrut” examination in biology revisited. Journal of Research in Science Teaching, 22, 31–40. Tamir, P. (1989). Training teachers to teach effectively in the laboratory. Science Education, 73, 59–69. Tamir, P. (2004). Curriculum implementation revisited. Journal of Curriculum Studies, 36, 281–294. Tamir, P. (2006). Inquiry in science teaching and its reflection in biology teaching in Israel. in A. Zohar (Ed.), Learning by inquiry: An ongoing challenge (pp. 15–56). Jerusalem, Israel: Magnes Press. Tamir, P., & Lunetta, V. N. (1981). Inquiry related tasks in science laboratory handbooks. Science Education, 65, 477–484. Tamir, P., Nussinovitz, R., & Friedler, Y. (1982). The design and use of practical tests assessment inventory. Journal of Biological Education, 16, 42–50. Zohar, A. (2008). Teaching thinking on a national scale: Israel’s pedagogical horizons. Thinking Skills and Creativity, 3, 77–81. Zohar, A., & Schwartzer, N. (2005). Assessing teachers’ pedagogical knowledge regarding issues pertaining to instruction of higher order thinking. International Journal of Science Education, 27, 1595–1620.

Pinchas (Pini) Tamir    267

This page intentionally left blank.

Chapter 13

Shifting Paradigms in Science Education A Change Agent’s Life on the Edge Barbara S. Spector University of South Florida

You have always been ahead of your time. —Meta VanSickle, 2012

Prologue Change is the only constant. Everything is connecting. A holistic image is emerging. I have been dissolving boundaries among entities throughout my career. The boundaries insulating formal education institutions K–16 that separate them from the community were the first to go. Then came the boundaries separating each of the sciences, the sciences from technology, and both from society. Next, boundaries among all disciplines began to dissolve yielding a transdisciplinary image. Ultimately the boundaries between practice and research blurred. Calls for systemic thinking were coming from all directions. This is the story of my journey as a change agent in science education. It was a rocky start, on a long twisting path strewn with fallen trees and an Going Back for Our Future, pages 269–296 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

269

270    B. SPECTOR

occasional boulder. My story chronicles the increasing ways I involved the community as a resource to enhance science teaching, people who influenced me, innovations I developed demonstrating the viability of Science/ Technology/Society (STS) reform, lessons I learned about being a change agent, and my practice-research-theory cycle. The innovations were the subjects for my research to learn ways to improve environments for creative (out-of-the box) science teachers in schools and the learning of science. Regardless of the position I was in (to earn my paycheck), I used it as a platform to be a change agent to improve science education by shifting from the dominant, reductionist, mechanistic paradigm of transmission teaching to a holistic paradigm (STS) of constructivist learning. Earth’s last frontier, the ocean, its mysteries, its influence on humans became the perfect vector for science education change initiatives. It is even a great segue to today’s hot topic, climate change. Stepping Around a Big Boulder The door to medical school slammed shut. Now what? I was fresh out of college with a biology bachelor’s degree cum laud in hand. Disappointment and frustration reigned supreme. What to do for a plan B—teach? It was a socially acceptable profession for a woman in 1957, especially for a newlywed . . . teach for a few years until you have a baby and settle in as mommy and homemaker. From childhood my intent was to be a doctor to help people heal their bodies. At Erasmus Hall High School in Brooklyn, New York, I was an office assistant to my female chemistry teacher, Anna Stanko, and my male biology teacher and department chair, Thomas Lawrence. My discussions with them and success in all my science classes encouraged me to pursue medicine. When I graduated from college however, I was told by several medical school recruiters that as a married woman, I would have children and would never practice medicine. I would not be a good investment for any medical school. (Fortunately this attitude no longer exists.) Since I couldn’t go into medicine and make people’s bodies better, I decided I could help people improve their minds by going into teaching. I enjoyed informally tutoring my peers in elementary school, and my behind-the-teaching-scene experiences in high school were rewarding. My pre-med undergraduate major made teaching high school biology a logical place to start my career. Because New York state required a Master’s degree for permanent certification, I completed a Master’s degree in combined sciences. I began teaching at Nottingham High School in Syracuse, New York where my assignment included Regents biology classes and a basic biology class. I taught in the way I was taught by transmitting information but with a twist. I encouraged students to make 3-D

Shifting Paradigms in Science Education    271

interactive models to understand material and made up biology games and puzzles. Students had some choices as they freely roamed among laboratory stations and the class zoo. Along with the ubiquitous aquarium and assorted white rats and mice, my classroom housed two large turtles, a bed of crayfish, and a porcupine. All was well in the Spector class biome until the porcupine escaped and brought disaster upon his fellow zoo mates. When the AP Biology teacher retired, the principal offered me the opportunity to teach this senior biology elective. I said I would do it if I did not have to teach AP. I thought these seniors could learn freshman college biology when they went to college. There were numerous exciting things they could do and learn as high school seniors that they would not have the opportunity to do in college. When no one else was willing to take over that course, the principal agreed to let me do whatever I wanted as long as it kept the parents happy and off his back. Now, the door was opened for me to use this course as a continuing source of learning for me as well as my students. I liked the notion of students’ interests directing where we went with the learning (a generative classroom). This ensured students would perceive science as relevant to themselves and I would not be bored. My caveat was they needed to choose something with which I was not familiar. If students wanted to repeat what had been studied previously, I engaged them in self-directed investigations using materials (audio tapes, photographs, and other print matter) the previous classes had collected. I was available to answer their questions. Human pathology was a favorite. After their self-directed inquiry, we extended the topic by inviting different people from the medical school to present new topics. It was a special treat when students who had been in advanced biology and were now in medical school came back to present as expert guests. Social issues were high priority for us, such as the right to die law, extraordinary life prolonging technologies, controversial psychiatry issues, fluoridation of drinking water, aging, and so on. I used the community as the resource for us to study many ideas that I knew little or nothing about. My professional lens has always dominated my life. I saw science education opportunities in every aspect of my life and every place I went. Everyone with whom I interacted could be tapped to contribute to my teaching, whether in high school or at the university. It is not surprising that my teaching has always involved numerous ties to the communities in which I have lived and traveled. I just assumed schooling and community went hand in hand. I began by inviting scientists I met socially, or who were parents of my students, into my classroom to talk with my advanced biology students. When I found someone who welcomed us to his/her laboratory, I initiated site explorations away from school. In this affluent community, many seniors drove their own cars to school and could provide transportation to local sites. As time passed we went on overnight and week-long ex-

272    B. SPECTOR

cursions (e.g., to Woods Hole, MA & Andros Island, Bahamas). Of course, I had to negotiate with the other teachers in school to ensure they got something they valued by letting my students miss their classes. Creating these win–win situations was the beginning of my integrating science into other disciplines. For example, my students wrote the contents and produced a complete edition of the school newspaper for the English teacher, who was responsible for the paper. To say this course was a wonderful adventure is the understatement of the year! I co-authored a book titled Community Resources for Meaningful Learning (Spector & Barnes, 1988) to help teachers have the same enjoyable and productive experiences I had with community entities. I used this in my higher education career when I began teaching prospective and experienced educators to teach science. Later in my career in higher education, each time I told my daughter about a new course or program I developed at the university, she said, “There you go again reinventing advanced biology.” I did not see the pattern until she said that. It becomes visible throughout this chapter. I realize now that many of the things I did when teaching (for no reason known to me other than it seemed a logical thing to do) were, in fact, not the traditional way of approaching science teaching at the time. I assumed they were common and ordinary. I had no idea I was pioneering new territory. Today those things I started doing early in my teaching career are aspects of science education reform I keep refining. They have distinctive labels now and are part of the literature facilitating the paradigm shift from traditional transmission science teaching to the desired state for teaching described in the National Science Education Standards (1996), A Framework for K–12 Science Education: Practices, Crosscutting Concepts, and Core Ideas (2012), and other documents guiding science education reform. Some common labels in today’s science education literature that reflect what I have been doing most of my career are teacher as facilitator (instead of sage on stage); constructivist teaching/learning; enacting cooperative learning; Socratic questioning; developing communities of learners/of practice; using the community as a resource; interacting with scientists and other community stakeholders including informal educators; poking holes in the academic silos to implement transdisciplinary learning opportunities; implementing middle school philosophy; promoting reflection and metacognition leading to autonomous learning; and engaging in experiential learning (situationalbased learning, project-based, problem-based, or place-based learning). Time for a Change As much as I loved my job at Nottingham High School, I had, however, often felt thwarted as a high school biology teacher because there were numerous

Shifting Paradigms in Science Education    273

ideas I wanted to implement with all my students, but the structure and policies of the school and its administration made them too difficult to enact. I was always having to find ways around the system to make anything exciting and new happen. I succeeded with unusual ideas in my advance biology course, but not in the other courses, because of the school’s bureaucracy and traditional assumptions about science teaching. I kept a file titled “Death of Creativity in a Biology Teacher.” Many of the ideas in this file were later stimuli for strategies incorporated into my book Empowering Teachers: Survival and Development (1989). When the opportunity to do additional study surfaced at Syracuse University to learn to be an internal change agent, I jumped at it. I wanted to find a way to ensure other biology teachers who wanted to do creative things to enable youngsters to learn science could live in a school environment that supported risk-taking and rewarded a teacher’s creativity. They would not have to experience the frustrations I did. I was fortunate to be awarded a fellowship from Dr. William Ritz’s National Science Foundation (NSF) grant for curriculum implementation in science education. It was designed to enable teachers to be internal change agents in a secondary school. I remember telling an education management consultant in Syracuse how surprised I was to be awarded my graduate fellowship. She told me I had been preparing myself to be ready for such an opportunity all the while I was a high school biology teacher. I did not recognize I was doing that, but I can see it now. Contacts Count (Becoming an Agent of Change) My doctoral degree was to be in science education with Dr. Ann Howe as my committee chair and implementation of Dr. Marvin Druger’s Project Advance Biology Course as the subject for my dissertation, Dr. Arthur Blumberg (Art) and Dr. Robert Bogdan (Bob), however, were the two professors who most influenced my path as an agent of change for the rest of my career. Surprisingly, neither are science educators. This seems to be consistent with Joel Barker’s (1990) point that one should look for change ideas at the edges of enterprises where different disciplines intersect. Art was an expert in organizational development and sensitivity training and educated people to be agents for change in the Educational Administration and Supervision Department. He was the founding father of the Council of Professors of Instructional Supervision (CPOIS). I knew Art before graduate school because his children were my students at Nottingham High School. I told him I did not know what to call what he did, but I knew I wanted to be able to do it. Even though he had a reputation for being so rigorous that students would never graduate if they worked with him, I wanted to shadow him. I did spend most of my time interacting with

274    B. SPECTOR

him. Other professors even traded off required courses so I could work directly with him. Art taught organizational development (OD). There was only one book available that applied organizational development principles to education, so most of the OD examples we studied came from the business world where it had been developing. These examples served me very well later in my career when I linked community stakeholders with science education initiatives. The change agent skills I learned from Art also enabled me to turn higher education courses I taught into communities of practice. My knowledge of group dynamics, processes, and life cycles helped me be a facilitator, instead of a sage on stage. Building on those skills enabled me to enact the paradigm shift in science education as a change agent in numerous settings. After my introductory courses in educational research design during graduate school, I wasn’t excited about doing a quantitative study for my dissertation. I did not perceive science education research to that point as having much to say to me as a teacher. I wanted to understand why things were as they were, what things meant, not how many there were or test other people’s theories. The foundation for this was my childhood in an orthodox Jewish setting where the norm was questioning the Torah to construct new understandings and meanings year after year in light of new experiences and new contexts. While trying to decide what my study would be, I met with Art, now my primary mentor. I told him, “If I am going to have to put in the effort a dissertation needs, I want to find something interesting and meaningful. Just learning to do research is not a good enough reason to put in the time required. I have no intention of ever doing another research study in my life.” These turned out to be famous last words! When I was in Washington D.C. a couple of years later, I called Art at home from my office late at night. I was excited to share with him the numerous questions that were generating in my head and all the research studies I wanted to do as a result of being the Program Director for Education in the National Sea Grant College Program in Washington DC and my involvement in national initiatives. Art was a member of my doctoral committee. He told me about a professor named Dr. Robert Bogdan (Bob) on campus in the Special Education Department who did something called qualitative research. It enabled a researcher to find out why things were as they were and what meaning they had. That piqued my curiosity. Dr. Bogdan was well known for his research and subsequent activism that led to many reforms in mental health institutions. As I understood it, the research he did enabled him to be a very effective agent for change. Since I wanted to be an effective change agent in science education, I decided to find out more about what he did. I had already finished all my coursework for my PhD. I asked Bob if I could just sit in on his qualitative research

Shifting Paradigms in Science Education    275

classes to see what it was all about. He graciously agreed. My committee chair, Ann Howe, was also curious to know more about qualitative research. She often joined me sitting in on Bob’s courses. During those courses we were introduced to The Discovery of Grounded Theory: Strategies for Qualitative Research by Barney G. Glaser and Anselm L. Strauss (1967), in which they articulated the process of grounded theory. It described an emergent process ideal to study phenomena arising from complex causes where the emergent phenomenon is beyond the sum of the things that are known and is beyond what could be predicted by what is known. Emergent design can be used to describe the process of deriving grounded theory: One systematically gathers and analyzes data, and the analysis of the data direct where you go next in the inquiry. Theory emerges from the data. This theory provides us with relevant predictions, explanations, interpretations and applications (Glaser & Strauss, 1967). To this day the way I construct original ideas and products (from a middle school children’s camp to post-doctoral education opportunities), all the products I develop, the research I conduct on those products, and the way I interact with my students reflect emergent design. Bob taught me to ask the ultimate research question: What is going on here? This is my best tool as a change agent. It enables me to see things participants immersed in a situation may not see that are key to solving their problems or resolving issues. I never tire of asking this question in all the contexts in which I find myself. The Unknown I did not know that qualitative research was not a method accepted in the science education enterprise in the United States at that time (1973), nor that it was just being introduced to the social sciences. Education can be considered a social science. Qualitative research was a novel and controversial process to understand how people viewed themselves and their worlds. Bob was pioneering qualitative research when he and Steven Taylor wrote the book titled, Introduction to Qualitative Research: The Search for Meanings, which was foundational to the field in 1975. Revisions of this book continue to be a classic even though today there are journals, university courses, handbooks, and encyclopedias for qualitative research. So it was that I was to unknowingly become a pioneer for qualitative research in science education. The first mention of qualitative research in science education I ever knew about was a luncheon presentation by Wayne Welch at a National Association for Research in Science teaching (NARST) meeting in Grossinger’s hotel in 1981. His message (published in 1983) to the assembled group, as I understand it, was that there is this new approach to research being introduced in

276    B. SPECTOR

education called qualitative research, and science educators should look into it. This was six years after my dissertation study was accepted in 1975. My dissertation was a grounded theory qualitative study. The emergent theoretical model explained the way high school teachers embraced change in how they taught when they implemented Dr. Marvin Druger’s audio-tutorial college biology course in high schools (Druger & Spector, 1979). Almost a decade after I had actually gathered the data for my dissertation, I presented the emergent model for teacher change from my dissertation at NARST in 1983 during a session I shared with Dr. James Gallagher. This was the first time there had been a concurrent session at NARST in which qualitative research papers were presented. I got great feedback from the audience, and people seemed intrigued with my work. My findings were lauded as something current and new. I had to bite my tongue to not say, “I did this study 10 years ago. It is not new at all.” In the late 1980s there was much discussion and disagreement about what constituted qualitative research, or “interpretive” research, in NARST. Dr. Merton Glass (then my graduate student) and I sought to make sense of the conflicting language that appeared to be inherent in the disagreements. We wrote the chapter “What’s in a Label: The Vocabulary of Interpretive Research” (Spector & Glass, 1991), which was published in the NARST Monograph #4 Interpretive Research and Science Education (Gallagher, 1991) where Dr. Gallagher wrote Confusion about terminology is a frequent criticism leveled at interpretive research. In this chapter, Spector and Glass provide a clarification of some of the confusion pertaining to this field by providing an overview of the several research traditions that are encompassed within it. Their analysis will be valuable to researchers who are trying to understand the subtleties of interpretive research as a means of enriching their investigations and the reports made from them, and to people who analyze and review research reports as a way of improving their judgments about the quality and significance of interpretive research. Moreover, this chapter should engender reflection by and productive discussion among, practicing researchers and those entering this field of research. (p. 19)

Earth’s Last Frontier In 1979 I was invited to the nation’s capital on an Intergovernmental Personnel Act (IPA) loan to be Program Director for Education in the National Sea Grant College Office in the National Oceanic and Atmospheric Administration. My vehicle for change became a congressional mandate to infuse marine education in K–12 schools throughout the nation. My role in Sea Grant entailed designing strategies to infuse marine education in K–12 schools. I

Shifting Paradigms in Science Education    277

was to engage extension agents in the thirty Sea Grant Colleges across the U.S. who would work as marine educators with K–12 local and state education agencies. Here I was again linking the science community and schooling. My experiences had taught me change in school environments requires a systemic approach. This meant involving stakeholders in all sectors of a community, including higher education institutions’ science, engineering, and education faculty; K–12 schools, including teachers, administrators, and board members; informal education agencies; technical service providers; business and industry; civic groups; government agencies; and policymakers. My focus became inventing ways to engage all stakeholders in coordinated and collaborative initiatives to create a common vision and process to achieve changes facilitating marine literacy. This required a transdisciplinary approach that later would be inherent in the STS movement. Relationship building was key to influencing leaders in each stakeholder group to bring their constituents on board. Building these relationships foreshadowed things to come in all my change initiatives. The Road Less Traveled Pays Off When I moved into higher education in 1981 at Florida International University (FIU), I began educating in-service teachers in graduate classes and conducting research on their need for professional development. The data, which later appeared in the article Generating a Desired State for Master’s Degree Programs in Science Education Through Grounded Theory Research (Spector, 1985), were used by the Florida Department of Education as a framework to get their Eisenhower funds from the U.S. Department of Education (USDOE). I pursued my goal to change schools to be more friendly to teachers who wanted to work out-of-the-box via grants to develop innovative courses consistent with features of what was later labeled the science, technology, society reform movement. My approach was described in the chapter Inservice Science Teacher Preparation in STS: Perspective and Program (Spector, 1986). Social issues to make learning science relevant and modeling instructional features that currently appear in the various documents guiding reform were emphasized. At the time I began writing grant proposals at FIU, NSF and other funding agencies explicitly wanted to entice scientists and engineers to contribute to K–12 science education. I had been challenged to develop ways to successfully orchestrate communication cross-culturally among scientists, engineers, and educators when I was in Sea Grant. Each group had its own culture. Educators were at the bottom of the totem pole and few in number in the Sea Grant network. As I continued to work across these cultures for new grant initiatives, the number of challenges to effective communication

278    B. SPECTOR

and successful partnering seemed to increase the more I worked with representatives from these groups. Learning from these interactions became a focus of research to help me foster systemic change. Later in my career (2003), I discovered informal science had still another culture that needed to be bridged for successful systemic change. I was introduced to the label “androgogy” as a contrast to pedagogy while I was at FIU. This was my license to teach the in-service teachers in my classes as the experienced adults they were, to honor their knowledge, and communicate on an equal playing field. This is in contrast to the hierarchy I normally saw between professor and graduate student. Today, treating a classroom as a community of practice is beginning to come to the fore in higher education. This posture has reaped rewards for me and my students. I get consistent feedback from graduate students indicating their self-efficacy has increased because they perceive their ideas are visibly valued. At FIU apparently having a new assistant professor on a non-tenure earning soft money line writing multiple grant proposals immediately was not expected. Having just left my position as a program officer in the National Sea Grant College Program, where I was dispensing grant funds and encouraging other federal funding agencies to fund our K–12 marine education mission, it seemed perfectly normal to me. After four years and seven funded grants, the University of South Florida (USF) recruited me for a tenured faculty position. In my exit interview with the provost at FIU, he told me some folks at FIU perceived me to just be an opportunist writing for any grants that happened along. He suspected that was not the case and asked if I had some grand plan in mind. I did indeed have a plan! I had conceived a plan for systemic change in science education while I was in Washington, D.C. Each time a request for proposal (RFP) appeared, I wrote a proposal for another piece of my systemic reform plan. It did not matter which audience and task was called for in the RFP, because they were all pieces of my puzzle to engage in a systemic approach to reform. I developed each proposal to fit into my plan. The grants I pursued were to develop what later would be labeled STS reform based education initiatives so I could have something to study to learn how the holistic paradigm worked with learners. In retrospect, this plan has undergirded my leadership role in the STS reform movement throughout my higher education career.

Everybody Into the Pool My charge at USF in 1985 was to update their science education graduate programs. I collaborated with a highly experienced mathematics educator, Dr. E. Ray Phillips, to develop the Graduate Program of Excellence for

Shifting Paradigms in Science Education    279

Mathematics and Science Teachers (GpEMSt) (Spector & Phillips, 1989). Program development was supported by Florida Eisenhower funds at the time. We started by modifying three courses I had developed with USDOE Title III funds at FIU and ended with a series of nine sequenced and articulated courses making up the GpEMSt program (pronounced GEMS). Professors from a variety of departments and colleges on campus joined with us to develop and teach these transdisciplinary courses. To the best of my recollection, I signed contracts for about fifty different professors. Transdisciplinary education flourished once again. When Ray and I brought professors from different disciplines together to discuss ideas for this program, it was not unusual to hear a comment such as, “This is the most fun I’ve had on campus in years.” One of the highlights of work with the GpEMSt program for me was presenting it to Japan. I was part of a team from the United States, lead by Dr. Robert Yager, engaged in a three year study with a team from Japan’s science education leadership comparing issues of inservice education. It was also presented in Russia. The GpEMSt brochure advertising the program stated the following: GpEMSt is a state, nationally, and internationally recognized series of nine graduate courses including the World of Water program, designed to enable teachers to meet the goals and recommendations of national mathematics and science organizations and the Comprehensive Plan: Improving Mathematics, Science, and Computer Education in Florida. The World of Water is a summer residential middle school program which serves as a clinical site for the instructional, and administrative innovations taught in the courses. (Spector & Phillips, 1987, n.p.) Florida’s Educational Reform Act of 1983 funded programs providing the state’s pre-college students with summer learning opportunities in science, mathematics, and computers. The program was intended to encourage the development of creative approaches to teaching of these disciplines. Under this program, between 50 and 60 high achieving middle school students were in residence on the University of South Florida campus for 12 consecutive days of study in the World of Water (WOW) program. . . . Eight specially trained teachers were in residence with the students. Between 50 and 70 experts from the university, government, business, and industry interacted with students each year in an innovative science/technology/society (STS) program. (Spector & Gibson, 1991)

Most of the learning in the WOW was done at community sites in driving distance from USF, such as the Living Seas Oceanarium at Epcot, a waste disposal plant, a golf course, phosphate fields, Tampa Bay (via sailing), grass flats (via snorkeling), a mangrove beach (for transecting), and other physical locations. New sites were explored each day to study the science

280    B. SPECTOR

and technology present in order to contribute to students’ understanding of the role of water in our world. In a grounded theory qualitative study of WOW, my graduate student Charles Gibson and I identified factors influencing students’ intellectual growth. We were very excited about the theoretical model that emerged describing the nature and features of the learning process in WOW. I presented the manuscript titled, A Qualitative Study of Middle School Students’ Perceptions of Factors Facilitating the Learning of Science: Grounded Theory and Existing Theory (Spector & Gibson, 1991) to the head of the National Middle School Center located in our building and to a colleague who was a leader in the middle school movement. They were thrilled with it, because “It exemplified the way middle school philosophy could truly be implemented!” When I was told that what we described illustrated the emerging middle school philosophy perfectly, I was disappointed that someone had thought of these ideas already. On the other hand, I was delighted they were so excited to see a functioning example of the fact that what they were professing for middle school philosophy really worked. It turned out I had been pioneering the viability of middle school philosophy and did not know it. I used the model for educating middle school youngsters that emerged from nine years of WOW to lead the conversion of Monroe Junior High School to a middle school and as a jump off point to guide the transformation of Manatee County School District’s science program to the holistic paradigm. Teachers from Sugg Middle School in Manatee County and Monroe Junior High School were learners in the GpEMSt program. Dr. Meta Van Sickle, then my doctoral student, explicated the role of caring in relationships necessary to the success of these initiatives (Van Sickle & Spector, 1996). This foreshadowed what was to come in my future change initiatives. Success on a Bumpy Trail Qualitative research had become somewhat established in education beyond science by the time I moved to USF in 1985. I encountered an English education professor, Dr. Marie Nelson, who had been using grounded theory for a long time. She encouraged me to continue with this type of research that suited me so well. Later Dr. James King, a reading specialist, was doing qualitative research in the form of critical theory. He and I took up the gauntlet to lead the battle in the College of Education to make qualitative research acceptable for dissertations in the college and to teach students how to conduct qualitative research. We developed a two course sequence (8 credits) to co-teach any students in the college to do qualita-

Shifting Paradigms in Science Education    281

tive research. (There had already been qualitative research courses in the communications department in the College of Arts and Sciences, but not in education and not for science education students). Today qualitative research has grown in our college and as many as five different professors are teaching such courses. Students are also encouraged to take any of the numerous qualitative courses in the different colleges on the campus. Being Mentored Matters At USF my most influential collaborator for twenty years has been Dr. Paschal Strong, experimental psychologist, now retired, and all around Renaissance man. He contributes to developing and teaching most of my out-of the-box courses, serves as a sounding board and a support system, and is my most severe critic. If my ideas pass muster with him, I know they will work with any other scientists I want to involve in a task. He enables me to use a psychology lens to make sense of data. When I was frustrated by data from studying students in my methods course for elementary science after testing a variety of approaches with them over five semesters, he pointed out aspects of the culture of this population and how their culture was contrary to the culture of science (Spector & Strong, 2001). Understanding this culture clash made me better able to relate to these students. Humans learning about each other’s cultures to build relationships turned out to be key to people working with each other effectively in every setting in which I worked as a change agent. When Dr. William Rhodes (Bill), a world renown special educator, was brought to USF, I learned to use the language of the paradigm shift. Bill was famous for his introduction of a new paradigm to special education. I was getting a fair amount of criticism from the other science educators in my department about the way I was doing things. Dr. James Paul, Chair of the Special Education Department, knew about it. He told me I needed to meet Bill Rhodes. He said all the work I was doing in science education paralleled what Bill had been doing in special education for which he had been lauded. Bill introduced me to the idea of a paradigm shift to explain my work. He built on the idea from Thomas Kuhn’s Structure of Scientific Revolutions (1962). Bill pointed out how and why my work constituted a paradigm shift and how it was consistent with what he had introduced in special education. He convened a national special education meeting to talk about manifestations of the paradigm shift in special education and asked me to present what I was doing in science education. This I did. Building on Bill’s work, I sorted out the general characteristics of the competing paradigms in society visible in science education (see Table 13.1).

282    B. SPECTOR Table 13.1  Competing Paradigms Dominant Reductionist Paradigm

Holistic paradigm

There is one objective reality independent of a person that can become known to an individual. Truth is correspondent to the objective reality.

Reality is constructed by individuals within their own minds. Therefore there are multiple realities. Truth is what a group working in a field at a given time agree to call reality (socially constructed). The whole is greater than the sum of its parts. Pieces are altered when they interact to become part of the whole. Cause and effect relationships involve multiple factors, are complex, and may be difficult to distinguish. Networks dominate the organization of information, people, and things.

The whole is equal to the sum of its parts. Parts are discrete, each having their own identity. Cause and effect are linear and immediate.

Hierarchies are the prevailing model organizing information, people, and things. One can know the world by analyzing isolated smaller and smaller pieces. Science, using this reductionist approach, is the legitimate way of knowing.

One can know the world by examining the whole. Science is one of several equally valid ways of knowing. The wholeness of the person, the union of the physical, spiritual, intellectual, and emotional aspects of the individual, is acknowledged. Process is a product.

Source: Spector, 1993

The paper I eventually published giving specific examples in science education of the competing paradigms was titled, Order Out of Chaos: Restructuring Schooling to Reflect Society’s Paradigm Shift (1993). Bill help me construct a way to convince the skeptics in my college why the work I was doing made sense in the context of our changing society. The Chaos paper explicitly compared the traditional way of doing business in science education with the way consistent with changes in society and science education reform. The message was consistent with what was published years later as the “more emphases-less emphases” pages in the National Science Education Standards (1996). The Chaos paper is still effective in helping teachers sort out the conflicting messages they continue to get about what they should be doing in schools: for example, do they drill and kill to conform to demands of the No Child Left Behind Act (NCLB, 2002), or do they teach for meaningful learning and have youngsters learn to inquire? Another person who helped me explain my work was Dr. Joseph Novak. He spent his sabbatical with me at USF in 1994–1995 while he was writ-

Shifting Paradigms in Science Education    283

ing his book titled, Learning, Creating, and Using Knowledge: Concept Maps as Facilitative Tools in Schools and Corporations (1998). He wanted to test some of his writing with my students. I wanted him to help me answer my usual question of “ What is going on here?” “Here” being the various programs I developed, such as GpEMSt, WOW, and STS, which seemed to me to be different from each other, yet worked so well. I asked him to conduct a seminar course with me for some students and professors who had participated in the programs. The learning theory he developed explained why all the programs, from the ones with the middle schoolers to the ones with the graduate students, were so successful. Initiating and facilitating change, especially change in school climate and being able to make science learning relevant to students through use of social issues as I had done in advanced biology, was a consistent focus for me. What I wanted to accomplish was years later labeled Science/Technology/Society interaction (STS). The process to create change in school environments involved much more than I ever suspected. My post-doctoral education in organizational development and change came from working with Dr. Patricia S. Bourexis (Patti) on and off over many years. Patti is the leader of The Study Group, Inc. and works behind the scenes primarily in science education and special education. I first met Patti when I was a consultant to Project 2061 where she was working in its early days. It was she who recommended me for a National Institute of Science Education (NISE) fellowship and introduced me to Senta Raizen and Susan Loucks-Horsely. She is my ideal change agent! Her forte is planning and project design, coaching projects for success. She has commonly worked as a project evaluator on complex initiatives. (She is great at herding cats!) For example, she coached our team of about thirty disparate science educators studying science teacher education programs in nine universities to reach our potential while serving as evaluator for the Salish I project. I was awarded a NISE fellowship and a sabbatical year in 1998. These gave me opportunity to inquire into research initiatives to shift paradigms in science education throughout the nation. While I was gathering data to answer the question of how to bridge the gap between pre-service and in-service science and mathematics education to create an articulated continuum for teachers, I found evidence reinforcing much of the learning from my own change initiatives. The impact of the differences in cultures among scientists and educators was critical to all the initiatives studied. The Research Monograph No.17: Working Toward a Continuum of Professional Learning Experiences for Teachers of Science and Mathematics (Mundry, Spector, Styles, & Loucks-Horsley, 1999) categorized the vast cultural differences between scientists and educators and listed them this way: personal, professional, and institutional philosophies; mission; approaches to educating teachers; institutionally sanctioned uses of time, norms, work styles, status, reward systems, and use of knowledge.

284    B. SPECTOR

These categories are still used successfully by my doctoral students learning to be change agents to investigate ways to work across cultures with all stakeholder groups influencing science teacher education. Research and Development Cycle By now I expect you will not be surprised when I say my research umbrella has been how change occurs in education. My career-long inquiry began with my first research initiative, my dissertation. This experience led me to a career long cycle of practice-research-theory development-application back to practice. In addition to developing 31 university level courses, most of them at the graduate level, and a residential summer camp for middle school students that ran for nine years, I designed many other learning opportunities for various local and state education agencies across the U.S. in order to have something to study to understand how change occurs in education. To interpret change in the world of education, I commonly defaulted to the research lens I derived from working with Art Blumberg during graduate school: that is role theory, which addresses factors influencing the way people enact their roles in groups from their families to their work places. Teachers are Key Since teachers are key to changing education at all levels K–graduate school, many of my initiatives/interventions related to how to enable inservice teachers to change what they are willing and able to do with their students (Spector, 1986). This research focus on change first became especially relevant in the 1960s with the Sputnik era and later with the grassroots evolution of the STS reform movement in various places throughout the United States (Spector & Yager, 2011). My studies revealed there were multiple influences from outside the classroom and school on teachers’ willingness to change the way they enacted their roles. This, once again, suggested that change initiatives had to be systemic, involving all entities that in any way interfaced with schools or the learners who were products of schooling. Thus all sectors of a community were needed to facilitate meaningful and lasting change. Working as a consultant to Project 2061 with Dr. F. James Rutherford and several NSF state systemic grantees reinforced my findings. Thus audiences for my research with whom I innovated change initiatives had enlarged from high school science teachers of biology to include stakeholders from multiple areas in a community with potential to impact change in science teaching. Much of my research was influenced by the availability of grant money and consult-

Shifting Paradigms in Science Education    285

ing funds. When these sources of funding were not available, I focused my research on changes I could make in my own teaching in the university. The Mystique of the Sea Reappears The Center for Ocean Sciences Education Excellence–Florida (COSEE–FL) began at USF’s College of Marine Science in 2002. “The mission of the COSEE Network is to spark and nurture collaborations among research scientists and educators to advance ocean discovery and make known the vital role of the ocean in our lives” (COSEE Network, 2013). I had to exercise all my change agent skills while directing this initiative for five years. During the first community meeting where educators and scientists came together, it became painfully clear that people were talking past each other, even though they used the same or similar words. The vastly different cultures in which scientists and educators lived required convincing all of them it was necessary to learn about the other’s culture. This need was not intuitive for either target audience. Scientists were commonly not sensitive to human cultural concerns; educators often accepted the hierarchy in the academy and the rest of education as a given. Addressing cross-cultural activities was not an easy task. Among the multiple initiatives for which I used COSEE-FL funds was a five course graduate emergent sequence simply titled, Ocean Sciences Community Building 1, 2, 3, 4, and 5. The pilot cohort included highly accomplished scientists, educators, and other stakeholders exploring each other’s cultures, developing a common language, and creating a vision for changing education to focus appropriately on the ocean. Flashback—shades of Sea Grant—I was once again using marine education as a vehicle for change. My accomplishments in marine education, while using it as the vehicle for change in the science education enterprise, was part of the commentary when I was awarded the status of “Fellow of AAAS” by the American Association for the Advancement of Science (AAAS) in 2007. An outgrowth of COSEE-FL was a graduate certificate program meeting needs of informal science education institutions (ISI). Because the ISIs have freedom in the intellectual content and instructional design they use to educate multiple audiences, I currently see them as potential leaders for change in science learning in schools as well as for the public. Educating informal science educators to effectively interface with the formal education enterprise in support of the STS/STEM (Science, Technology, Engineering, and Mathematics) reform movement may be a light at the end of the reform tunnel. Although formal university courses are unusual structures for interventions by change agents, I developed them as a means of providing support over time for change (instead of short workshops) and to meet my university obligation to produce student credit hours and degreed graduate students.

286    B. SPECTOR

Reflecting back on the 31 courses I initiated for science teacher education over 30 years, each course modified the nature and the amount of community participation by increasing and diversifying aspects to which the community had input. Numerous change initiatives culminated in the most outof-the-box initiative, a model derived from the program I currently direct: The Informal Science Institutions Environmental Education Graduate Certificate Program. The model, titled Spector-Leard Community Engagement Model for Course/Program Development, is a model for course and program development that provides a higher education institution with a new option to fulfill its commitment to community engagement. (Many Association of American Universities [AAU] institutions are currently committed to community engagement and engaged scholarship.) The model involved stakeholders as equal actors in every aspect of the program from initiating the idea, to verifying the need, to conceptualizing course content and instructional design, to developing course materials, to pilot testing, to teaching in the pilot test, to conducting research and evaluation, to converting it to distance learning, to writing about it for publication (Spector & Leard, 2012, unplublished). My Focus on Social Issues My original aim to change schools to supportive environments for teachers to develop new and creative ways to help their students learn was embedded in the diverse programs I have described. As political winds shifted and different funding sources became available, I modified the labels I used for my change initiatives to fit the currently used label, but my trajectory remained on target to implement a holistic paradigm. Most recently the STS label has been changed to STEM, and professionals nationally are trying to figure out and agree on exactly what this means (R. Yager, personal communication, January 17, 2012). The focus on social issues to make learning science relevant has been a persistent thread throughout my change initiatives to shift paradigms. In an earlier autobiographical work, I addressed the question of how I came to use social issues in the teaching of science. The chapter was titled “Serendipity: A Change Agent’s Friend in Academia” (Spector, 2006) in the book Addressing Social Issues in the Classroom and Beyond (Totten & Pedersen, 2006). I traced events beginning with the context in which I was brought up at home and how it generated the belief my teaching of science needed to contribute to the betterment of the human condition, and the way this was expressed in the various professional positions and projects in which I was involved. That chapter told the chronological story of how I moved from introducing social issues in a high school biology class, where I was perceived as stealing time from my real job of teaching basic science facts of the biology curriculum, to my current posture

Shifting Paradigms in Science Education    287

that the interaction of science, technology, and society (STS) is the real job in any science classroom. The STS movement, as I understand it, is synonymous with the current reform movement that began in 1982 and continues today under the label of STEM—science, technology, engineering, and mathematics interactions. It is a synonym for constructivist (one makes sense of new data by adding, deleting, or rearranging information in one’s idiosyncratic cognitive framework) epistemology, learning theory, and teaching/learning approaches. The intended outcome for students is to learn how to learn, so they are capable and inclined to make reasoned decisions in our science and technology driven society. Learners become scientific and technologically literate through systematically inquiring into STS events, which are multifaceted issues that challenge society and the future of the planet. STS is used as an organizer for the entire curriculum, integrating many disciplines (Spector, 2009). My use of social issues in science education—the STS reform movement—evolved from a combination of experiences and opportunities: (a) those that came naturally when I was teaching in high school, including using the community as a resource; (b) applying my doctoral education in organizational development and grounded theory qualitative research methods; and (c) grant funds being available to test my ideas in new settings. These three things account for the way the practice-research-theory development-practice cycle evolved throughout my career. Driving Forces As far back as I can remember I believed that I had to contribute something to bettering society. In my family and the Jewish community in which I was brought up, it was taken for granted that one had to make a contribution to society that made life better for others in order to make one’s life worthwhile. This, combined with (a) my seemingly insatiable curiosity about why things are as they are, (b) my low threshold for repetition triggering boredom, (c) my need to continually learn something new, and (d) my need to share and do something with what I learned, are some forces that drive me. My need to be creative and constantly learn something new was overtly visible in WOW. At a staff meeting one night everyone was tired and wanted to have a short daily debriefing meeting. One person said, “Let’s tell Barbara we don’t have anything to analyze because everything went beautifully today so we can go to bed.” The program leader replied, “You don’t understand Barbara. If you tell her everything went well, she will say, ‘Good. Since we know that works, let’s figure out what new thing we can test tomorrow?’” I completed personality analysis tool that assessed affiliation, achievement, and power as human motivators (McClelland, 1958) when I began graduate school at Syracuse University was quite revealing. I scored very high on affili-

288    B. SPECTOR

ation and achievement, but next to nothing on power. Affiliation made me collaborative, and achievement made me persistent. The absence of power motivation influenced the strategies I used to accomplish goals throughout my career. It was difficult for me to learn ways to deal with people who were power motivated. The combination of lack of strategies to deal with power motivated people, most of whom live in a win-lose world, my need to have an outlet for my creativity, and my need for achievement drove me to developing new initiatives to express my creativity and test my professional prowess each time a colleague tried to block my initiative or take over “my turf” (e.g., took over a course I was teaching, tried to steal a grant, etc.). I preferred to use my time, energy, and connections to develop something new rather than to battle over something that already existed. Typically after I did something once, researched it, refined it from the research findings, and tested it again, I was ready to teach someone else to take it over anyway. Later in my career, I described my propensity to avoid conflict in the academy and release attachments to initiatives I developed that others wanted to take over in a chapter with Dr. Patricia Simpson and Dr. Cyndy Leard titled, Making Lemonade from Lemons: A Road to Leadership for Women in Science Education (2009). We provided strategies consistent with women’s ways of knowing to enable the science education enterprise to gain maximum benefits from women in the profession, many of whom tend to be more affiliatively and achievement motivated than power motivated. Affiliation leads me to a highly socially interactive learning style that is most productive during experiential learning opportunities. My combined need for affiliation and achievement led me to strive to create win-win situations in whatever setting I was functioning. This is also the basis for my classes being run as communities of co-learners (now labeled communities of practice) and the collaborative relationships on an equal playing field I established with my graduate students. I enjoy writing papers collaboratively more so than alone. Talking through interpretations of data with people who have different perspectives intrigues me. Actually sitting together at a computer in person or via conferencing technology is great fun. Further, my socially interactive learning style explains my propensity for networking. I am inclined to telephone the person working on a topic to ask a question before I look for their work in print matter on the web or in hard copy. Networking beyond Syracuse when I lived there was kick started when I traveled around the country for Sea Grant. Those three years were pivotal to me developing national perspectives and establishing contacts with state education agencies, other government agencies, and professional associations. I learned the potential for large scale change inherent in professional science education associations when I presented at each of their board meetings requesting they follow the lead of the Council of Chief State School Officers and pass a policy statement to forward marine education.

Shifting Paradigms in Science Education    289

In the university, when multiple grants were awarded to me in the same year, each having salary funds for me, it was suggested that I use the excess salary for travel. This enabled me to participate in the meetings of the eight state commissions in Florida resulting from our governor leading the Education Commission of the States’ science and mathematics improvement initiatives. It also facilitated my travel to national professional association conferences. Throughout my career I have continued to contribute to the professional science education associations by serving on boards of directors and chairing committees and task forces to forward reform. Serving on advisory boards of a variety of NSF and USDOE-funded initiatives fed my desire for affiliation and achievement and provided opportunity to exchange ideas with many fantastic people. For example, my relationships with Dr. Paul DeHart Hurd, Dr. Joseph Piel, and Dr. Thomas Liao emerged from these board meetings. Paul became my role model. I adopted his perspective on the slow pace of change and it has kept me from being discouraged by continuously emerging barriers in change over time (Spector, 2006). I learned the cutting edge role of technology and engineering thinking needed in the science education enterprise from Joe and Tom. Grant funds also enabled me to support graduate students’ travel. My graduate students and new faculty I mentor always comment about the importance of me taking them where ever I went and introducing them to everyone they had read about. I have networked within science education and outside science education. To this day I am continually looking to connect with new networks of people in different disciplines and throughout the stakeholder community. Networking often stimulates collaborations as well as use of technology for distance learning. Collaboration and Distance Learning I deliberately explored strategies for collaboration (Spector, Strong, &King, 1996). I initiated and supported collaboration among science educators in the state; among professors in a variety of departments and colleges on my campus; among formal and informal science educators; among scientists and educators at all levels; among educators and stakeholders in numerous sectors of a community; among educators and policy makers; within my classes to create communities of learners; with my colleagues nationwide; and with students to do research and write papers. Early in my higher education career I collaborated for several years with Marianne Betkouski Barnes. Even though she was in Jacksonville and I was in Miami and then Tampa, we got together at each other’s homes for several days at a the time to explore ideas and write a book series, Science

290    B. SPECTOR

Teaching in a Changing Society: Grades 6–12 (1988–1990), and many conference presentations together. One day while discussing the difficulty science education has had over so many years of trying to get teachers to teach through inquiry, Marianne said something like, “I don’t see how they can teach through inquiry if they’ve never learned through inquiry.” That comment began my quest to design interventions that would give in-service and potential teachers opportunity to learn through inquiry. Eventually, I developed a model to structure an entire course as an inquiry into a focal question learners would answer by gathering data for the entire semester. For example, the Science/Technology/Society Course focused on this question: What is STS, and how is it related to science teaching? In the early years of the course, I provided numerous resources from the STS Center I directed. When the university made Web CT available, I developed a virtual resource center using the web to provide students with choices of what to explore, and the order in which to explore, from among a multitude of resources. Of necessity, students developed their own mini inquires nested within the focal course inquiry. Assignments were used as learning opportunities, not as tests to check on students’ work. They were open-ended, posted on the web for input from all class participants, and revised. The course final assignment was to convince the instructional team of their learning using any mode with which they were comfortable. Learners were encouraged to incorporate their outside activities and hobbies. Learners can be amazingly inventive! They have produced art work, poetry, plays, dance, music, woodwork, multi-media and more illustrating their meanings of STS. Inquiry requires learners to reflect on what and how they are making sense of experiences to determine what else they need to know. Using openended inquiry requires autonomy. Autonomy requires self-assessment on a regular basis to determine what more has to be done to reach a learning goal. Thus self-assessment is inherent in learning through inquiry. This need for students to exercise autonomy in their learning was a challenge. Research on this 1979 model revealed a variety of deficits students had acquired from their lives as dependent learners in schools. Among them were metacogniton, self-assessment and evaluation, time management, scientific discourse, and systematic inquiry (Spector, Burkett, & Steffen, 2002). In order to mitigate these deficits, I made some modifications to the open-ended inquiry and developed a pre-requisite course that was less open-ended and introduced extensive explicit self-assessment procedures. These courses combined to form the program described in NSTA’s Exemplary Science: Best Practices in Professional Development (Yager, 2005) in a chapter titled, Hey! What’re Ya Thinkin’? Developing Teachers as Reflective Practitioners (Spector, Burkett, & Leard, 2005). The STS course ultimately went fully online with help of techies Dr. Barbara Lewis, Dr. Ruth Burkett, and Mr. Dave Bethany. Feedback from presentations at technology and National Association of Science Technology Society

Shifting Paradigms in Science Education    291

(now IASTS) conferences suggested this was the first time an online distance learning course was structured as an inquiry and this was the first course to illustrate the way constructivism could be implemented online (Lewis, Spector, & Burkett, 2001). I applied the model to other courses I taught, including the various methods courses for elementary, middle, and high school pre-service teachers. I continued to elaborate on self-assessment procedures for use in all my courses as reported in Mitigating resistance to teaching science through inquiry: Studying self (Spector, Burkett, & Leard, 2007). My students’ enthusiasm fuels me to keep going into new ventures. I tend to get involved with students’ projects, processes, and the entire enterprise in which they are embedded. For example, I accepted an appointment to the National Advisory Board of Earth Force, a not-for-profit service learning organization, in order to support the work of one of my ISI graduate students employed by Earth Force. Fallen Trees on the Path My own propensity to constantly push the envelope of what is known is as much an obstacle to my career as it is a facilitator. Much of my research has been reported in papers that were presented at conferences but were never published. By the time I understood what I was investigating enough to write a coherent paper to give as a presentation at a conference and processed the feedback from my peers at the conferences, I was already on to another inquiry. I more often than not lost interest in the presented paper and did not go back to revise and format it for publication. My “to do” list, if I ever retire, will include revisiting those research endeavors and submitting them for publication. As noted earlier, the mathematics educator Dr. E. Ray Phillips was my collaborator in the GpEMSt program at USF. He served as a mentor and buffer for me because of his close friendship with the powerful associate dean of programs. We introduced processes and ideas consistent with the reform movement in both disciplines that were very different from what was going on in the college at the time. Once the program got a lot of good publicity, our colleagues in mathematics and science education were feeling their programs were threatened and started giving us a hard time. I suggested to the dean that we get an external evaluation of the program to find out whether it really was cutting edge and worthwhile. The dean agreed to hire a science educator who was a program officer for research at the National Science Foundation, Dr. Vince Lunetta. Vince came and did a three day site visit. He interviewed students in the program, professors who had participated in the program as guest speakers, and administrators in the university. One of the skeptical College of Education faculty suggested it was necessary to have

292    B. SPECTOR

an internal evaluator also, since the program needed to fit into the college appropriately. Dr. Doug Stone was the leading person in evaluation in the Research and Measurement Department in our college. I asked him to participate in the evaluation process. He wrote a separate report from Vince’s report. Both reports lauded the program as being cutting-edge, incorporating and modeling all the latest research about teaching and learning. Lessons for a Change Agent In higher education other people copying my work and taking credit for it and/or taking over my successful courses could be labeled a barrier to me continuing to do what I want to do. On the other hand, as a change agent, I use these events as facilitating factors to move onto other change initiatives. It all depends on one’s point of view. Over the years serving as an agent for change, I learned two key personal perspectives on the role of change agent: Remember imitation is the sincerest form of flattery, and an effective change agent has to leave his/her ego at the door. For example, a variety of topics and perspectives from my STS course are now being included in other professors’ courses. One explicit example was where a colleague and I agreed laboratory safety belonged in my methods course. Students were excited to have the opportunity to learn about laboratory safety. The next thing I knew, my colleague put laboratory safety into his course, also a required course in the program. Instead of being upset because my colleagues were stealing my thunder, I reminded myself that imitation is the sincerest form of flattery. I used the opportunity to remove laboratory safety and other items from my courses and filled those spaces with something new. Leaving my ego at the door is sometimes a downer. When I entered a group as a consultant to facilitate them solving a problem, the problem was solved, and the participants took ownership of the solution, I left the site. Months later while they were successfully executing the solution to their problem, if they said “Barbara who?” I knew I had done my job well. I often thought of my behind-the-scenes role as consistent with the traditional role of a wife in a household in bygone days: The best way to get what she wanted was to make her husband think it was his idea. This change agent role description is contradictory to the needs in higher education for tenure and promotion. A barrier in higher education to being a successful change agent is the need to get public accolades and credit for your ideas and initiatives. This is not only true when you are seeking tenure and promotion but continues as a requirement for merit pay or other rewards in academia. Being behind-the-scenes facilitating initiatives and others’ accomplishments quietly is not consistent with the reward and

Shifting Paradigms in Science Education    293

recognition processes in higher education. The ultimate achievement for a change agent, however, is when the client takes total ownership of a project and brings it to successful fruition. In essence the client has learned how to solve a problem and owns the solution. Teachers guiding students to learn to solve a problem is comparable to a change agent facilitating a client learning to solve his/her problem and owning the solution. An episode illustrating this in my book Science and Technology as Human Enterprises (Spector & Lederman, 1990) comes to mind. Staff in WOW were taught a Socratic approach to questioning. They answered a question with a question as a means to guide students’ thinking and illustrate the way a line of questioning can be developed to solve a problem. It is similar to guiding clients to invent a change strategy with the same ownership effect. Here are some students’ reflections from 1989. When you ask a question, often times a teacher would ask you questions about the problem. This may seem strange because they sound as though they are trying to avoid the question, but in actuality, by answering their questions you can develop an understanding of the problem and can answer your own question. This gives me a good feeling. It makes me realize that I know more than I give myself credit for. . . .  I learned to use that knowledge.

In Closing While shifting to the holistic paradigm for science education as a change agent for reform in multiple environments I pioneered STS, qualitative research, middle school philosophy, community resource-based teaching, online constructivist instruction, and cross-cultural communication among our stakeholders. I have contributed to the growth of new generations of change agents who are continuing to work in variety of environments from public schools, to universities, to government agencies, to informal science education organizations of many types. Students tell me I am a role model for those who want to foster change in our enterprise. Typical comments from interviewees for Lois Ball’s dissertation were explicit about “applauding Barbara for being vocal . . . where it doesn’t come across as being overly pushy. That’s hard to do. Women, whether they are teachers or researchers themselves, get overshadowed by men . . . (and) need to be heard.” “She is a very effective mentor and supporter of career advancement.” Our history shows developing new curricula is the default act to change what goes on in schools for science. I would like my legacy to be our enterprise focusing more on the human side of change, the relationship building required to support changing this complex enterprise, and rewards for

294    B. SPECTOR

those who want to step outside the box to make science relevant and meaningful to learners. References Barker, J. (1990). The business of paradigms [video recording]. Burnsville, MN: Charthouse International Learning Corporation. Bogdan, R., & Taylor, S. J. (1975). Introduction to qualitative research methods: The search for meanings. Hoboken, NJ: John Wiley & Sons. Druger, M., & Spector, B. S. (1979). Implementation of a college biology course in high school. The American Biology Teacher, 41(7), 413–415. Gallagher, J. (Ed.) (1991). NARST Monograph, #4 Interpretive Research and Science Education (pp. 19–42). Manhattan, KS: NARST Kansas State University. Glaser, B. G., & Strauss, A. L. (1967). The discovery of grounded theory. Chicago, IL: Aldine. Kuhn, T. S. (1962). Structure of scientific revolutions. Chicago, IL: University of Chicago Press. Lewis, B., Spector, B. S., & Burkett, R. (2001). A constructivist model of an online course. Tech Trends for Leaders in Education and Training. McClelland, D. C. (1958). Methods of measuring human motivation. In J. W. Atkinson (Ed.), Motives in fantasy, action and society. Princeton, NJ: D. Van Nos-trand. Mundry, S., Spector, B. S., Stiles, K., & Loucks-Horsley, S. (1999). Research Monograph No.17: Working toward a continuum of professional learning experiences for teachers of science and mathematics. Madison, WI: National Institute for Science Education, University of Wisconsin, Madison. National Research Council. (1996). National science education standards. Washington, DC: National Academy Press. No Child Left Behind (NCLB) Act of 2001, Pub. L. No. 107-110, § 115, Stat. 1425 (2002). Novak, J. (1998). Learning, creating, and using knowledge: Concept maps as facilitative tools in schools and corporations. Mahwah, NJ: Erlbaum. Spector, B. S. (1985). Generating a desired state for master’s degree programs in science education through grounded theory research. Journal of Research in Science Teaching, 22(4), 327–345. Spector, B. S. (1986). Science education reform: A state policy issue. Science Education, 70(2), 129–137. Spector, B. S. (1988). Community resources for meaningful learning. Dubuque, IA: Kendall Hunt Publishing Company. Spector, B. S. (1989). Empowering teachers: Survival and development. Dubuque, IA: Kendall Hunt Publishing Company. Spector, B. S. (1993). Order out of chaos: Restructuring schooling to reflect society’s paradigm shift. School Science and Mathematics, 93(1), 9–19. Spector, B. S. (2006). Serendipity: A paradigm shifter’s friend in academia. In S. Totten & J. Pederson (Eds.), Researching and teaching social issues: The personal and pedagogical efforts of professors of education (pp.  181–206). Lanham, MD: Lexington Books.

Shifting Paradigms in Science Education    295 Spector, B. S., & Barnes, M. B. (Eds.) (1988–90). Science teaching in a changing society: Grades 6–12. Dubuque, IA: Kendall Hunt Publishing Company. Spector, B. S., Burkett, R., & Leard, C. (2005). Hey! What’re ya thinking? Developing teacher as reflective practitioners. In R. E. Yager (Ed.), Exemplary science: Best practices in professional development (pp. 189–201). Arlington, VA: NSTA Press. Spector, B. S., Burkett, R. S., & Leard, C. (2007). Mitigating resistance to teaching science through inquiry: Studying self. Journal of Science Teacher Education, 18(2), 185–208. Spector, B. S., Burkett, R. S., & Steffen, C. O. (2002). Factors contributing to preservice teachers’ discomfort in a Web-based course structured as an inquiry. Journal of Educational Technology Systems, 30(3), 293–310. Spector, B. S., & Gibson, C. E. (1991). A qualitative study of middle school students’ perceptions of factors facilitating the learning of science: Grounded theory and existing theory. Journal of Research in Science Teaching, 28(6), 467–484. Spector, B. S., & Glass, M. (1991). What’s in a label: The vocabulary of interpretive research. in Gallagher, J. (Ed.), Monograph #4, Interpretive research in science education (pp. 19–42). Cincinnati, OH: National Association of Research in Science Teaching. Spector, B. S., & Leard, C. (2012). Stakeholder participation in course development: A model for community engagement and engaged scholarship. Unpublished manuscript. Spector, B. S., & Lederman, N. G. (1990). Science and technology as human enterprises. Dubuque, IA: Kendall Hunt Publishing Company. Spector, B. S., & Phillips, E. R. (1989). Excellence in graduate education for mathematics and science teachers: A sciematics approach. School Science and Mathematics, 89(1), 40–48. Spector, B., Simpson, P., & Leard, C. (2009). Making lemonade from lemons: A road to leadership for women in science education. In K. C. Wieseman & M. H. Weinberg (Eds.), Women’s experiences in leadership in K–16 science education communities: Becoming and being (pp. 47–62). Dordrecht, The Netherlands: Springer Publishing. Spector, B., & Strong, P. (2001). The culture of traditional preservice elementary science methods students compared to the culture of science: A dilemma for teacher educators. Journal of Elementary Science Education, 13(1), 1–20. Spector, B. S., Strong, P. N, & King, J. R. (1996). Collaboration: What does it mean? In J. Rhoton & P. Bowers (Eds.), Issues in science education (pp. 177–184). Washington, DC: National Science Education Leadership Association & National Science Teachers Association. Spector, B. S., & Yager, R. (2011). The many faces of STS: Social Issues in science education. In S. Totten & J. Pedersen (Eds.), The study of social issues: Major innovations, models, approaches, programs. Charlotte, NC: Information Age Publishing Inc. Van Sickle, M., & Spector, B. (1996). Caring relationships in science classrooms: A symbolic interaction study. Journal of Research in Science Teaching, 33(4) 433–453. Yager, R. E. (Ed.) (2005). Exemplary science: Best practices in professional development. Arlington, VA: NSTA Press.

296    B. SPECTOR

Chapter 14

E. Joseph Piel Pioneering Technology in Science Education Barbara S. Spector University of South Florida Rene Goytia University of South Florida

It was September 1941, my cross-country flight for my license as a flight instructor. Oh no, I flew over the White House! (FAA unforgivable pilot error). Poof! There goes my commercial pilot’s license! December, 1941: The same FAA recruited me to be an air traffic controller! I had to escape that seat job fast. I volunteered to be a marine combat pilot.

Introduction Marine Colonel Emil Joseph Piel, known to science educators as Joe, is a New Jersey native. He was born in Fairview, New Jersey, April 17, 1918, lived there

Going Back for Our Future, pages 297–307 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

297

298    B. S. SPECTOR and R. GOYTIA

for the first five years of his life, moved to Edgewater, to Fort Lee, and then went into the service during World War II (WWII). When this chapter was being written, Joe was ninety-four years of age residing in West Caldwell, New Jersey in a cozy retirement cottage with his second wife. His first wife, who died in 1995, was in the Navy when they met and married in 1945. He commented during our telephone interview that email reception was difficult in his current environment, which meant this amazing technology innovator was back to using “snail mail” (personal communication, July 7, 2012). Joe Piel chose his life’s career in science education when he was in high school. He had a choice of three different tracks to follow in school. He chose science. This was considered one step above the general academic program that was offered in the school and students in the science program considered themselves to be brighter than other schoolmates. Indeed, he certainly was brighter than most and was actively involved in whatever activities were available. “I was in Boy Scouts, when I lived in Fort Lee, yes, and I was a Life Scout and I was the assistant Scoutmaster. . . . Then, church groups, the YPF, the Young People’s Fellowship, I was in that group . . . .I was in the band, student director of the orchestra and drum major . . .” (Spector, 2005, personal conversation). His willingness to be a risk taker, even back then, enabled him to grow in leaps and bounds. Love Affair with Planes and Cars It may be that Joe inherited his outstanding abilities in mathematics and science from his father, who was an engineer at Alcoa Aluminum Company in Edgewater, New Jersey. Additionally, his father repaired cars in the days of the Ford Model T, a fascination Joe also acquired. He learned to drive his uncle’s Model T at nine years of age. As a youngster, Joe used to build model airplanes and stand on the porch of his home launching them into the air by twisting a rubber band around the propellers. He had the opportunity to learn to actually fly and get his private pilot’s license as a result of a government program, Civil Pilots Training Program, during his junior year in college. He was hooked for life! His love of flying and the automobile influenced many of the science/technology approaches he introduced to science education. Education Even though his father never lost his job during the Great Depression, the family did not have very much money. Joe was able to acquire a job on the production lines rolling out sheets of aluminum, because his father was

E. Joseph Piel    299

then in management at Alcoa. Joe was offered a scholarship to Stevens College (NJ) when he graduated in 1935, but was not able to attend because the scholarship was insufficient to cover the high cost of living expenses. In 1936, he accepted a band scholarship to attend college at Montclair State University (NJ). Joe kept the Alcoa job working night shifts and earning extra money playing with a small dance band. He graduated in 1940. For three years during WWII (1942–1945), Marine Corps pilot Emil J. Piel had an illustrious career. He studied a great deal about engines during flight school. His love of flying and his nerves of steel carried him through service in the Solomon Islands where this Second Lieutenant flew as a dare devil deliberately drawing anti-aircraft fire to identify the location of the Japanese guns buried in the jungle so U.S. bombers could destroy them. He has, on occasion, described this period as his government-sponsored vacation in the South Pacific. An extensive interview with Joe about his WWII experiences is in the Rutgers oral history library (Illingsworth & Whiteman, 2005). That interview makes it clear why the barriers in the education enterprise did not faze him in the least. Joe also used his teaching talents while he was in the military. He said this about it in the 2005 interview: They called me “Uncle Joe.” [laughter] . . . That was because I was older and because I’d been a teacher. That was another thing that I didn’t mention about pre-flight school. We also had people who were having difficulty, particularly with physics and with celestial navigation, and we had a busy day at pre-flight school, but those who were having difficulty and it looked like they were good specimens, they were good people, but they might not pass the ground school, then, I taught them at night. They said, “You’re a teacher,” you know. So, we had an upstairs in the barracks [where] we had lessons, in mathematics and physics, mostly. (Illingsworth & Whiteman, 2005, n.p.)

Joe made use of the GI Bill in the 1950s. He went back to Montclair State University after World War II to earn his master’s degree in physical science. He then earned his PhD in supervision and curriculum specializing in physics from Rutgers University in 1960. He attended Rutgers part time while he was teaching high school and took courses full time in the summers. His doctoral degree focused on curriculum reform. In fact, his first attempt at a dissertation was turned down, because it was not perceived as acceptable research. He told this story in an interview in 2005: When I got . . . just ready for the oral . . . my adviser said, “You know, I’ve passed your thesis around to a lot of people,” which they do, and he said, “They all came to the conclusion that you didn’t have a thesis here. What you had was a career,” and what I was planning to do would have been a career, not just a thesis, and so, my name was on the program, you know, the whole thing, but I didn’t get the degree, and so, it took me a year, and then, I did this study to

300    B. S. SPECTOR and R. GOYTIA get . . . background, which I already knew. . . . It was one of those things, you know, chapter three, you’ve got to really prove why it’s necessary for you to do all the rest of the chapters. [laughter] I knew why, but that wasn’t good [enough]. I had to go through the literature and find out why [laughter] and the literature was written by a lot of people I knew, and so, I got that squared away, and then, wrote this system for re-educating physics teachers, and then, my doctoral exam was very interesting, because it was three people from the Physics Department and three people from the Education Department and Education and Physics didn’t get along too well at Rutgers at the time and it got to be interesting. They asked me a question and I answered it, and then, somebody from Education was concerned about the kind of paper I used to write the thesis on, you know, [laughter] and somebody from Physics said, “What the hell has this got to do with what he wrote?” [laughter] and so, then . . . there was a little difference of opinion, discussions of physics education and education, and I was there mostly as a moderator [laughter] and it was really funny. I got finished. . . . You go out of the room. Now, you sit in the hall on the bench and I sat on the bench and I was there about two minutes and they called me back in. They said, “Congratulations,” and that was the end of my oral exam. (Illingsworth & Whiteman, 2005, n.p)

The fundamental problem Joe perceived that stimulated his interest in changing science education was his experience with the way physics courses were being taught. They were downright boring and did not relate to the real world in any way obvious to the learner. He was frustrated by learners memorizing and solving equations with no understanding and no concept of why they were doing it, or how that information could be applied to their daily lives. His understanding of physics contributed so much to success in his young adult life and in the Marine Corps that he wanted to ensure future generations would benefit similarly from learning physics (and other sciences). In order to ensure learners would benefit from physics knowledge, he pioneered the teaching of technology used to solve human problems in traditional basic science courses. Prior to his work, it was not considered legitimate for a high school science teacher in any science to address “applied science,” and certainly not technology. Professional Science Teacher Joe was way ahead of his time. What is currently referred to in higher education as the “flipped classroom” (Gerstein, 2012) was essentially part of Joe’s early teaching lexicon. The notion that classroom time should be used for active learning in contrast to passive lecture and that students could receive straight delivery of information outside of the classroom was a no-brainer for him. An experiential approach was the name of the game for Joe.

E. Joseph Piel    301

Joe described the influence of his war time experiences on his teaching this way: . . . teaching physics and having been a pilot, I used engines as the basis of my physics teaching. As a matter-of-fact, I taught a whole course, a whole physics course, around the automobile, because [of] the engine training that I had had in the Marine Corps, in the flight school, and so, I knew a lot about engines. As a matter-of-fact, right after the war, you could get an engine for fifty dollars. You could get an airplane for one hundred dollars, plus, one hundred dollars more for each engine that it had, and you had to fly it out of [the field]. Yes, they were getting rid of them that way, and so, I got an engine from a helicopter. . . . I was teaching in Fort Lee at the time and I had it shipped to Fort Lee and I had it right there in the classroom, and so, I could teach the whole physics course [around it]. Everything you needed to know in physics, you could teach . . . around the engine, especially . . . around the automobile. The whole thing with light and all the business of focusing light and . . . that whole business, that’s all the . . . automobile headlights, just as Archimedes’ Principle is the basis of the carburetor . . . the Archimedes’ Principle, and so, I taught the whole course. . . . At a convention, I would talk about this, you know, science teachers’ convention, and people would say, “You know, that’s very interesting, but can you ever get a kid in college [by] teaching that kind of a physics course?” and I’d say, “Well, I’ll tell you about one. I had a kid named Paul Gray in my physics course at Caldwell High School here,” and I said, “In spite of my teaching . . . about the automobile, he was able to go to MIT,” I said, “but he had a lot of trouble, because he couldn’t get out of MIT. He finally ended up as president.” [laughter] So, that stopped all arguments about, “Can you teach physics around the automobile?” (Illingsworth & Whiteman, 2005, n.p)

From 1948 to 1952, Joe taught physics in Caldwell High School (Caldwell, NJ), where he was qualified to teach all the science and mathematics courses offered in that school. From 1952 to 1960, Joe was the chair of the science department at East Orange High School (East Orange, NJ). A key event while working in East Orange High School was the opportunity for him to join the Physical Science Study Committee (PSSC). This was a combination of university and high school teachers established by the Massachusetts Institute of Technology to create a new high school physics course (Physical Science Study Committee, 1960). PSSC was the first of the “alphabet curricula” that were developed in the 1960s and 1970s to reform science education. While Joe was delighted with the laboratory experiences students were provided in PSSC, he was disappointed that the curriculum did not relate the physics content to societal problems. Joe also worked on another curriculum development project, Harvard Project Physics, through which he came to know F. James Rutherford (the father of Project 2061).

302    B. S. SPECTOR and R. GOYTIA

West Essex High School (North Caldwell, NJ). had the benefit of Joe’s leadership as its principal from 1961 to 1966 when he began working in higher education. In the summer of 1965, Joe worked full-time in the Bell Laboratories, which convinced him that teaching basic physics without simultaneously focusing directly on its application in the world beyond the classroom was a bad idea. Joe’s involvement with Bell Laboratories was a key influence on his work in science teacher education, especially curriculum design and implementation. In fact, when he chaired the East Orange High School science department, he made a point of placing his teachers in positions in Bell Labs, where they could be involved with actual research either during the summers or during the school year. The genesis of the idea for the curriculum titled, The Man-Made World (TMMW), which was to be a radically different curriculum published in 1971, was at Bell laboratories (Engineering Concepts Curriculum Project, 1971). There, Ed David, Jr., director of computers at Bell Laboratories, and John Truxal, chair of the electrical engineering department at Brooklyn Polytechnical Institute, agreed with Joe that, somehow, curriculum needed to be developed to show how technology and society affect each other. Joe focused on curriculum writing throughout his career after earning his PhD. He did whatever research was necessary to inform his curriculum writing and used the research to support proposals for grant funding to create materials making science relevant and usable by science educators at various levels in academia. He was very active in the New Jersey Science Teachers Association, often conducting workshops for teachers. From 1966 to 1972, the Brooklyn Polytechnical Institute, now known as the Polytechnical Institute of New York in Brooklyn, was Joe’s academic home. There he served as the executive director of the Engineering Concepts Curriculum Project (ECCP). This was a National Science Foundation (NSF) grant that supported the actual development of The Man-Made World (TMMW) Joe, John Truxal and Ed David, Jr. co-authored both the course and the book. (This was before any sensitivity to sexism in writing.) This science/technology course emphasized systems analyses, decision making, optimization, and related skills for high schools. Another NSF grant and the Sloan Foundation supported educating teachers from schools across the country in teacher centers in Houston, TX; San Diego, CA; and Lakewood, CO to learn to implement the course. Ultimately TMMW was translated by the Japanese and used to improve their science education: The Man-Made World targeted high school students somewhere between their sophomore science courses and junior physics courses. Physics was not a prerequisite for students to take TMMW course. When the “back to basics movement” started in the 1970s, Piel was vocal that TMMW was as basic as the 3Rs. TMMW enjoyed about a decade of publication and distribution by the book publisher, McGraw-Hill, and also served as a point of departure for

E. Joseph Piel    303 spinoffs, such as a college liberal arts curriculum titled, Technology: Handle With Care (Piel & Truxal, 1975); People, Technology, Society, for middle school students; and a videotape series, You and Me and Technology by M, Galey (1981) of the University of Denver for which Piel was a content specialist. (Spector, 2007, p. 300)

Joe was also instrumental in forming and sustaining the National Association of Science Technology and Society (NASTS) in the mid 1980s, now known as the International Association of Science Technology and Society (IASTS). Rustom Roy, professor of Solid State Geochemistry and Science, Technology, and Society at Pennsylvania State University, and John Truxal initiated the idea for the association and invited Joe to play a major role in its development. It was at meetings of NASTS that I (Barbara Spector) got to enjoy learning from Dr. Piel. He always took the time to share his thoughts. He was sorely missed when his wife became ill for a couple of years and he could only attend NASTS meetings briefly. Barriers Joe faced two major obstacles in his work. One is still common today as an obstacle for large-scale change in science classrooms: the need for ongoing in-service teacher education. The second obstacle was rather unusual and an issue of timing in a discipline other than science education. First, in order for teachers to use Joe’s curricula materials, they needed to embark on a paradigm shift from transmission teaching to being a facilitator of students’ learning by doing. This required significant in-service learning because it is was so new to the teachers. Further, most of the 400 teachers he educated in about ten years (1970–1980) were highly successful and moved on to other positions, leaving uninitiated teachers to teach the courses. Thus the courses were co-opted and eventually died, even in the schools where they had been most successful. The second obstacle emerged when a social studies curriculum titled Man: A Course Of Study (MACOS; Education Development Center, 1970) was implemented widely. In it, the tradition of Eskimos leaving their elderly to die peacefully by freezing to death on ice flows was reported. It created a huge furor in schools, and the curriculum was often removed from school offerings. The public frequently confused MACOS with The Man-Made World, thus giving the latter a negative reputation. Faciliators Joe’s harrowing and daring experiences in World War II left him totally unintimidated by anything in civilian life. His position at the State University

304    B. S. SPECTOR and R. GOYTIA

of New York at Stony Brook was a major factor in facilitating his work. He went to Stony Brook with six like-minded colleagues. One colleague, Jon Truxal, became the dean of the College of Engineering. Joe became a tenured faculty member with the charge to develop the technology and society department in the College of Engineering, along with Tom Liao, who was a faculty appointee. They all shared the same vision for advancing science education as an interaction of science, technology, and society (STS). Dr. Marburger, the president of Stony Brook, was an enthusiastic supporter of STS. Thus, Joe was under none of the usual pressures to publish research to get enough items on his vitae to attain tenure and maintain his university stature. He was able to focus on what he believed was the most important thing to do—create curricula and educate teachers to use those materials. As time passed and computers became readily available, Joe became an early innovator with the use of computers. He designed computer animations to elucidate difficult concepts in learning science, such as teaching the optics of vision with his bird in a cage. He integrated new ways to use computers into all his teaching. He continued doing this until he formally retired from Stony Brook and became professor emeritus in 1987. He consulted with the department of technology and society that he had created through 2002. After that, he directed his talents toward working with his church, enjoying his family, his second marriage, and traveling. Lasting Impacts Joe’s perspective on the importance of including technology and engineering in science teaching K–12 has had long lasting impact on the science education enterprise. It began with his work in Project Synthesis the project funded by NSF in the late 1970s to create a vision for what science education should be in the United States (Harms, 1980). The director of Project Synthesis, Norris Harms, invited Joe to be the chair of one of the six committees creating databases for the project. The committee was titled the Science, Technology, and Society (STS) committee. The Project Synthesis Final Report to NSF (Harms, 1980) in Chapter 11 described the desired state for STS in school science. Joe popularized the concepts of soft technology and hard technology for science educators. Hard technologies were described as the things humans create and use to solve problems in their lives, and soft technologies were described as the processes used for solving human problems. The classic example Joe liked to use was the automobile: Highways and street lights are hard technologies, and the rules of the road by which we live and drive are the soft technologies (Piel, 1980).

E. Joseph Piel    305

In addition to introducing the concepts of hard and soft technologies to science educators, the following list of things taken for granted in science education today are among contributions of pioneer Joe Piel: Using social issues to teach science, problem and project based learning, experiential approaches to constructivist teaching, using computers to facilitate learning complex concepts, legitimatizing teaching applied science and technology in basic science classes to make science relevant, promoting teacher placements in research positions in science laboratories on a temporary basis to give them real world science and technology experience, and the movement to recruit minorities into engineering, Project Synthesis’ desired state for STS was actualized into curriculum Joe developed through his funded projects and the products of others. For example, Joe worked with Roger Bybee when Roger was the Director of BSCS Innovative Science Education (initially Biological Science Curriculum Study) in Colorado, which is the longest-lived curriculum development organization in the United States. Joe’s influence is visible in BSCS products such as BSCS Middle School Science and Technology (1994). The influence of Joe’s vision continues to be visible today in the draft of the new Next Generation Science Standards (Achieve Inc., 2013) which includes the largest emphasis on engineering of any of the documents guiding reform to date. The recent use of the acronym STEM (science, technology, engineering, and mathematics) for reform initiatives is testimony to the growth in perceived importance of engineering thinking in science education. The continued expansion of the Department of Technology and Society in the College of Engineering and Applied Sciences at Stony Brook University (formerly State University ofNew York at Stony Brook) is ongoing testimony to Joe’s impact in academia. This department he began has grown to include an undergraduate major in science and technology, a master’s degree and a PhD program. The programs apply concepts and tools drawn from natural sciences, engineering, and social sciences to examine and enhance the relationship between technology and our society, both regionally and globally. These concepts include systems theory, methods and tools for decision making, and science-technology-society (STS) frameworks. (Stony Brook University, 2013)

So it is that at 95 years of age, Colonel Emil Joseph Piel, PhD has lived to see his vision becoming a reality. References Achieve, Inc. (2013). Next generation science standards. Retrieved February 10, 2013 from http://www.nextgenscience.org/

306    B. S. SPECTOR and R. GOYTIA Education Development Center. (1970). Man: A course of study. Washington, DC: Curriculum Development Associates. Engineering Concepts Curriculum Project. (1971). The man–made world. New York, NY: McGraw-Hill. Galey, M. (1981). You, me, technology. [Film series] Boulder, CO: University of Colorado. Gerstein, J. (2012) The flipped classroom. Retrieved July 9, 2012 from http://usergeneratededucation.wordpress.com/2012/05/15/flipped-classroom-the-fullpicture-for-higher-education/ Harms, N. (1980). Synthesis: an interpretive consolidation of research identifying needs in natural science education. [Final report]. NSF–SED–80–003. Illingsworth, S., & Whiteman, D. (2005). Oral history of Joseph Piel. Retrieved July 8, 2012 from http://oralhistory.rutgers.edu/donors/30-interviewees/interview -html-text/324-piel-emil-j Physical Science Study Committee (PSSC). (1960). PSSC physics. Boston, MA: D. C. Heath. Piel, E. J. (1980). Interaction of science, technology, and society in secondary schools. In N. C. Harms & R. E. Yager (Eds.), What research says to the science teacher (Vol. 3). Washington, DC: National Science Teachers Association. Piel, E. J., & Truxal, J. G. (1975). Technology, handle with care. New York, NY: McGrawHill. Spector, B. S. (2005). E. Joseph Piel: The making of a risk taker in academia. In S. Totten & J. E. Pedersen (Eds.), Addressing social issues in the classroom and beyond: The pedagogical effort of pioneers in the field (pp. 291–305). Charlotte, NC: Information Age Publishing. Stonybrook University. (2013). Department of technology & society. Retrieved January 12, 2013 from http://www.stonybrook.edu/est

E. Joseph Piel    307

This page intentionally left blank.

Chapter 15

Pioneers in Science Education Marvin Druger Marvin Druger Syracuse University, Emeritus

At age 79, it’s not easy to summarize a career in science education that dates back to student teaching at Midwood High School in Brooklyn NY in 1954. It seems that I’ve done almost everything in my science education career, including teaching, research, administration, program and curriculum development, and countless service activities at the local and national levels. Any attempt for me to write an inclusive chapter about my professional activities would be necessarily incomplete. So, in this chapter, I have tried to include a glimpse of some of the experiences that helped shape my career in science education. A core guide to my teaching career and life is that we learn from everything that we do and everything that we do becomes part of who we are. We tend to forget information, but we remember experiences. Unique, motivational, experiences are keys to teaching and learning. In this chapter, I’ll try to highlight some of the special experiences that helped shape my life and career in science education. Going Back for Our Future, pages 309–325 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

309

310    M. DRUGER

In my long career in science education, I’ve had the privilege of serving in many leadership roles, including being president of the National Science Teachers Association (NSTA), president of the Association for the Education of Teachers in Science (AETS, now the Association for Science Teacher Education or ASTE), chair of the AETS Editorial Board of the journal Science Education, twice president of the Society for College Science Teachers (SCST), two separate terms as chair of the Education Section (Q) of the American Association for the Advancement of Science (AAAS), three separate terms as secretary of the Education Section of AAAS, College Division director for NSTA, Executive Board member of the National Association for Research in Science Teaching (NARST), program director for the Science and Mathematics Education Networks program at the National Science Foundation (NSF), president of the Syracuse chapter of Sigma Xi, president of the Kappa chapter of Phi Beta Kappa (three separate terms), director of numerous funded programs for teachers and students, and many other contributions to science education. In these leadership roles, I helped move these organizations forward in many ways and established a stronger foundation for them. I have been amply recognized with awards for my contributions, but mainly have been rewarded with the satisfaction of having made many diverse contributions to science education. The awards included the James Howard McGregor Prize for Outstanding Promise as a Teacher of Zoology (Columbia University), two Gustav-Ohaus Awards for Science Teaching (NSTA), a Distinguished Service Award (NSTA), the Chancellor’s Citation for Exceptional Academic Achievement (Syracuse University), the first Alumnus Award for Excellence in Teaching (Syracuse University), a Distinguished Alumnus Award (Brooklyn College), the L. Douglas Meredith Award for Teaching Excellence (Syracuse University), the Chancellor’s Citation for Exceptional Academic Achievement (Syracuse University), a Lifetime Achievement Award (Technology Alliance of Central NY), the Philip Martin Award as an Educator of Excellence (Central NY Education Consortium), the Honorary Emeritus Award (AETS, now ASTE), and the Robert H. Carleton Award for National Leadership in Science Education (NSTA). I cherish these awards, but I realize that there are many other individuals who are equally or more deserving of the awards but don’t ever get them, for a variety of reasons. The awards do give me a feeling of satisfaction, and it was good to know that my peers showed their appreciation for my efforts and accomplishments in science education. I was born on February 21, 1934. My immediate family consisted of my mother, my father, an older sister and two younger brothers. We were crowded into a sparsely furnished apartment, near a noisy elevated train. My father was a hard-working truck driver with a sixth grade education, and my mother was an uneducated “housewife.” My future as a teacher

Pioneers in Science Education    311

was influenced by my social acquaintances. I learned how to get along with people by interacting with my childhood social/athletic club members. BZ and Junior lived around the corner, and they were my best friends. Irwin, Larry, Allen, Marty, Carl, Joel and a few others rounded out the club. We played basketball, punchball, softball, and stickball. We didn’t smoke, drink alcohol or use drugs, and we were not very interested in girls. Teachers have to be able to interact with many different people, and being a member of the Wildcats was where my youthful social skills developed. I was good at academics, and I earned good grades in school. My parents were indifferent to my school performance. This was not their world. My older sister was bright, but, upon graduation from high school, she had to get a job to help support the family. She never had a college education. My sister, two brothers and I all went our separate ways. I was salutatorian at graduation from Pershing Junior High School, and was valedictorian at graduation from New Utrecht High School. Upon graduation from high school, I was awarded highest achievement medals in biology, social studies, French and some other subjects. In high school, I learned some valuable lessons about teaching. Mr. Goodman was my geometry teacher. “Is there anyone who was able to do that difficult homework problem?” I eagerly raised my hand. “Marvin, can you put that problem on the blackboard ?” I wrote the solution to the problem on the blackboard. Mr. Goodman, who was impeccably groomed and had a deep, pearly voice, said, “Good job!” Wow! After that remark, I became highly motivated to work intensively on the homework problems. The power of praise became fixed in my mind as a future teaching tool. I never thought much about going to college, but, because my older sister went to work, I didn’t have to get a job upon graduation from high school. BZ and Junior told me they were going to college. At that time, attending Brooklyn College was free. I thought, “For free, I’ll attend college too.” At Brooklyn College, I took the standard curriculum. Even though I played basketball on the Brooklyn College Freshman Basketball team, I managed to achieve good grades. At the end of my freshman year, I realized that I was not going to end up in the NBA, so I quit the basketball team. Thereafter, my grades got even better. I ended up graduating magna cum laude. Since I was good at all subjects, choosing a major and minor was difficult. I had no more interest in science than I did in other subjects. I earned an “A”  in the second semester freshman biology course, and I received a postcard from the professor at the end of the semester. “Nice job!” he wrote. That comment motivated me to take another biology course and, before long, I became a biology major and minored in science education. During summers, I worked as a counselor at Camp Brydon Lake, in Andes, NY. This was a formative step toward my becoming a science teacher. A Columbia University friend and I were the nature counselors at the

312    M. DRUGER

camp. Little did the campers know that we would hide behind a screen door at night, terrified by moths battering on the screen. At camp, I was also the riflery counselor, and the overnight hike counselor. I wrote and directed skits and wrote the Boy’s Sing, a major musical production. These were exercises in creativity that helped set the stage for my future as a teacher. My experiences as a camp counselor influenced me to minor in science education at Brooklyn College. As such, I worked toward certification as a high school biology teacher. I had excellent professors in science education who taught me the fundamentals of how to teach effectively. My student teaching experience at Midwood High School in Brooklyn was especially influential. On one occasion, I was trying to have students respond to a question in my words. My supervising teacher commented, “Why don’t you let the students answer in their own words?” An important lesson was learned from a brief remark. As an undergraduate biology major, I saw a notice that Theodosius Dobzhansky, a genetics professor at Columbia University, was going to give a lecture on Drosophila species (fruit flies) in South America. His unusual name and the topic intrigued me, so I attended the lecture. It was fascinating. I still vividly remember Dobzhansky using a wooden pointer to point to different fly populations on a map of South America on a screen. I was intrigued that Dobzhansky would be so interested in studying populations of flies in South America, and not the people there. Dobzhansky was one of the world’s leading evolutionary and population geneticists, but that didn’t impress me at the time. It would later on when I attended graduate school at Columbia University. My life’s pattern has always been to do as many different things as possible. Life is short, and I wanted to experience everything that I could. I was interested in almost everything. As a student, I had all sorts of part-time jobs, including being an usher at the Roxy theater in NY, a clothing store salesman, a Western Union messenger, a painter of ceilings at a grocery store, a leather craftsman making numerals for golf club bags, and a sprinkler of walnuts on fruit cakes at a bakery. I discovered that non-science experiences were important for development of any science teacher. This theme guided all the unusual activities I tried to provide for my students. During some summers, I worked on my uncle’s chicken farm in Carmel, NJ. My imagination was stimulated and I did foolish things, like trying to eliminate rats in the chicken coop by tying a nail to the end of a broom handle and attempting to spear the rats. (It didn’t work). Or, I’d lie on the roof of the chicken coop, pretending to be dead, in the hope that the buzzards flying above would approach me so that I could get a good look at them. One summer, I brought my genetics class project with me to the chicken farm. The project was to make crosses with Drosophila to determine the genetic identity of an unknown mutant. I accidentally put the bottles

Pioneers in Science Education    313

of flies near a heating duct, and almost all of the flies died. Nevertheless I wrote a report. When I returned home, I saw my genetics professor walking toward me on campus. When he came nearer, he signaled an “A” with his fingers to indicate my grade on the report. I was thrilled and proud to have been able to synthesize a valid report from such a scant amount of data. I was encouraged to pursue my interest in genetics. I started graduate school at Columbia University as Lecturer in Zoology. I was in charge of teaching adult students introductory college zoology in the evening. The adult students were highly motivated and I related well to them. They were often sacrificing to go to school and earn a degree. They were in school to learn as much as they could, and they demanded good teaching. Teaching adult students was like teaching your friends and peers. I continued to teach adult students for many years thereafter, and this teaching was one of my most important formative experiences. When I started my graduate program at Columbia University, I remembered the lecture by Theodosius Dobzhansky at Brooklyn College. I approached him and became one of his PhD students. Only then did I realize Dobzhansky’s prominence in the field of population and evolutionary genetics. Dobzhansky was a great scientist and a humanist. Practically every noteworthy geneticist in the world visited our laboratory at Columbia. A valuable lesson learned from Dobzhansky was that everyone is unique, has special talents, and deserves to be treated as being someone special. Dobzhansky treated his graduate students as peers and friends. I was thrilled to have him bring world-renowned scientists into my small room and say with his inimitable accent, “Marvin, Tell Dr. . . . your work.” I even had the opportunity to accompany Dobzhansky on a field trip to collect flies in the southwestern US. Dobzhansky proved to be a friend, mentor and surrogate father to me. By that time, I was married to Pat, who was an undergraduate at Brooklyn College, and Dobzhansky treated Pat as a friend and as an important, special person. Dobzhansky motivated me to treat my students in the same way that he treated me. At that time, the Zoology Department at Columbia had a faculty of many leaders in different scientific fields. I benefited greatly from interacting with Francis Ryan, L. C. Dunn, Teru Hayashi, Priscilla and Arthur Pollister, John Moore, Herbert Taylor, Ned Hodgson, Ruth Sager, Lester Barth, and others. They helped shape my philosophy of teaching and my knowledge about scientific research. While I was earning my master’s and PhD degrees in zoology (genetics) at Columbia, I did substitute teaching in high schools. So, I maintained my relationship with high school teaching, while pursuing a research degree in genetics. I was awarded the McGregor Prize for Outstanding Promise as a Teacher of Zoology. This was a great honor, and further encouraged me toward a science teaching career.

314    M. DRUGER

A formative experience as a graduate student was when I did substitute teaching at Westinghouse Vocational High School in Brooklyn. This school had tough kids. We were told that the important thing was to take accurate attendance, because the attendance book provided a legal alibi for any student who might commit a crime during school hours. The English teacher impressed tough kids by displaying a revolver on his desk. To gain their attention and stimulate interest, I tried to teach them about sex. They knew more than I did. One day, a student was asleep with his head on the desk. I approached him. “What’s the matter? Are you tired?” He didn’t budge. The student next to him said, menacingly, “Leave him alone. He’s tired.” I left him alone. Subsequently, I asked that sleeping student to stand in front of the room and ring a bell at certain intervals. If the class was unruly, I’d say to the bell-ringer, “I wouldn’t stand for such noise.” “Shut up!” was his command to the other students. They shut up. My “teaching” at Westinghouse High School posed discipline problems and helped me learn how to maintain confidence and handle discipline problems effectively. A colleague said it well, “The lions know when the lion tamer is afraid.” I learned that it’s important to have a ”command presence” in the classroom. Another teaching experience as a graduate student was a job as a substitute teacher at Brooklyn Technical High School. I taught a course in quantitative analysis to a very bright group of male students. I had never taken a course in quantitative analysis, but the class and I learned the subject together. That was a wonderful experience in self–motivated learning. I became convinced that students could learn anything, if they really wanted to learn it. I recognized the importance of teaching students to want to learn. My PhD thesis dealt with the effectiveness of selection of body size at different temperatures in Drosophila pseudoobscura. I also did research on chromosomal polymorphism in Drosophila, which was Dobzhansky’s main area of interest. After completing my PhD, I spent a year doing a National Institutes of Health (NIH) postdoctoral research fellowship at the Commonwealth Scientific and Industrial Research Organization (CSIRO) in Sydney, Australia. I did research on the genetics of pattern formation of scutellar bristles in Drosophila, under the mentorship of James Rendel. One day, I made an interesting discovery and I sought out Dr. Rendel who was in the tea room. He was lying on a sofa, completely still. I thought he was ill, “Are you all right?” I asked. His reply was, “I’m pondering the definition of fitness.” As an obsessive, hustling post doc from the U.S., I was stunned at his comment. That taught me another lifelong lesson. Sometimes, it’s good to just think, and not always be “doing” something. My post doc in Australia enabled my wife and me to travel around the world. Since then we have been to many places, including Switzerland, France, England, Greece, the Netherlands, Spain, Portugal, Italy, Israel, Egypt, Japan, China, Tahiti, Hawaii, Fiji, the Galapagos, Costa Rica, and

Pioneers in Science Education    315

other countries. Our desire to travel was stimulated many years ago when we visited Japan for an international genetics meeting. We saw an elderly couple checking out of a hotel because they couldn’t sleep on the futons on the floor. We vowed to take advantage of every opportunity to travel, before we got too old to enjoy it. As participants in an exercise class for older adults at a health club, we now realize that the problem wasn’t that the elderly couple couldn’t sleep on the futons on the floor; it was that they couldn’t get up from the futons. Worldwide travel shaped my strong belief that all prospective teachers  should experience another culture during their college education. Traveling abroad enabled me to learn about different cultures and relate to a wide variety of people. My experiences in Dobzhansky’s lab and my worldwide travels helped provide me with one of my major themes in life, namely, that each person is unique, has special traits, and should be treated with respect. Someone can be bigger, stronger, richer, smarter than someone else, but nobody is better than anyone else. As teachers, we should help students identify their unique traits and help nurture them. Obtaining a PhD under the prestigious mentorship of Dobzhansky proved to be extremely important. Upon graduation and prior to my post doc research experience in Australia, I never thought much about looking for a position when I returned from Australia. Without seeking them, I received two offers, one at the University of Michigan and the other at Syracuse University. The University of California at Berkeley also asked me to stop there to give a seminar en route to Australia, but I declined that invitation. After my year as a post doc in Australia, I decided to accept an assistant professor position at Syracuse University for several reasons. First, my family was close by in Brooklyn, NY and, second, I had taught a summer high school program at Syracuse the previous summer and I liked the school environment. I was born and bred in Brooklyn, New York and moving to Syracuse was like moving to the country, with grass, trees, lakes and farms. Most important was that I was offered a dual assistant professorship in the Department of Science Teaching and the Zoology Department. The Department of Science Teaching was officially in the College of Arts and Sciences and received all of its financial support from Arts and Sciences, even though all of its students were enrolled in preparation for a career in education. So, I was also considered a faculty member in the School of Education, even though all of my salary came from Arts and Sciences. This dual appointment satisfied my teaching and research interests. It gave me the opportunity to interact with scientists and with School of Education personnel. For several years, I did genetics research with Drosophila and taught a non-major zoology course, a course in evolutionary genetics, a variety of biology courses for adults in an Independent Study Degree Pro-

316    M. DRUGER

gram, and a methods course for prospective biology teachers. For many summers, I also organized, directed, and taught in a special program for talented high school students. This program was funded through the Summer Students Science Training Program (SSTP) at NSF. This summer program launched many careers in science, and I’ve had very gratifying emails from former participants who are now established in science careers. This program reminds me of the powerful effects the experiences we provide in our teaching can have on a student’s life. In addition, I directed a variety of funded programs for science teachers. In 1970, the Zoology Department and the Botany & Bacteriology Department merged to form a single Biology Department. The zoology and botany courses were combined into a single biology course. I taught the genetics segment of the introductory biology course in collaboration with three other biology professors. Each of us promised to attend the other professors’ lectures, but this never happened. After each professor taught his specialty, he disappeared into his research lab. Because of my growing interest in science education, I volunteered to assume the responsibility of teaching the entire freshman biology course. At first, a separate course was being offered for prospective biology majors, but these students were merged into my single course. In order to do an effective job, I decided to give up my laboratory research and devote myself to the challenging task of teaching the introductory biology course to about 1000 students each year. I adopted an audio-tutorial approach and developed interactive tape recordings and guide books that students would use in the laboratory. I also gave lectures, and graduate teaching assistants taught weekly recitation sections. This approach offered several advantages over traditional methods: labs were open day and night so that students could spend as much time as needed to learn the content, self-quizzes and answers were included in the guide books, and students learned how to take responsibility for their own learning. I tried to make my teaching “adventures in life.” Some of my major teaching objectives were to have students: • increase subject matter competency (I want students to know the basic concepts in biology science and how to use basic science vocabulary. Critical thinking skills are important, but you can’t critically think about nothing. Knowledge of basic concepts and vocabulary are important). • view science as a human activity (Science is basically the attempt of humans to make logical sense out of nature. As such, there are many unknowns, mistakes, revisions, politics, and even fraud. Understanding the nature of science is an important goal).

Pioneers in Science Education    317

• recognize the relevance of science to everyday life and to the intellect (Not all knowledge is relevant to everyday life and society; sometimes it’s simply nice to know things.) • think about problems in a scientific manner • improve speaking, reading, writing and technology skills in science • develop a positive attitude toward science • become a motivated self-learner I developed a vast repertoire of special activities to enrich the core of the course. Each course enrichment had a special objective, consistent with my strong belief in the value of “motivational experiences” in impacting students’ future lives. We forget information, but we don’t forget experiences, and we’ll never know which experience will have a substantial future impact. An underlying goal was to teach students to want to learn. Some of the special activities included: a creative project about life for extra credit, a team research project on plants, a supplementary Frontiers of Science lecture series by local scientists, a Pathways to Knowledge lecture series given by graduate students to provide insights about graduate research, a weekly course newsletter, evening exams to allow ample time for students to complete the exams, special film nights, a bio-answer show on television to provide students with answers after exams, bio-lunches with small groups of students in resident dining halls, a bio-feast at the end of the fall semester in one of the dining halls, and open office hours 24 hours a day, seven days a week. Benefit-of-the-doubt credit was given to students who attended most of the optional events. Attendance was kept at these events, and if a student was on a grade borderline at the end of the semester, the final grade would be boosted for attendees. A supplementary special topics course was made available during the second semester of the course for the students who did well in the basic course first semester. The special topic courses were designed and taught by experienced PhD students in their area of specialization, with advice from faculty members. One popular activity was the “Druger Drop.” After each major evening exam, students would gather in front of the biology building and I would throw answer keys out of a second floor window to the eager crowd below. Why would anyone do that? Again, every component of my course had a rationale, consistent with the goals of the course. The Druger Drop was an efficient way of distributing answer keys to a large number of students immediately after an exam; it also provided excitement among the students and was fun; it was an unusual experience that would help students fondly remember the biology course many years later. Indeed, whenever I see former students, they never ask me, “Are you still teaching about protein synthesis?” They almost always ask, “Are you still throwing answer keys out the window?”

318    M. DRUGER

My introductory course was also offered to high school students through the Project Advance program at Syracuse University. In this program, introductory college courses were taught in the high school, by high school teachers for college credit. I prepared teachers to do the job, made site visits to each school twice a year, and gave presentations to the high school students. I organized a workshop for the Project Advance teachers twice a year. I gained many insights about teaching problems encountered by high school biology teachers. Unusual experiences and motivation are cornerstones of teaching and learning. All of the components in my biology course were designed to provide a motivational learning environment that would generate positive memories later in the students’ lives and a desire to learn more about science. One example of such an unusual experience was when I was a Senior Fulbright Lecturer at the University of Sydney in Australia. I organized a bio-marathon in a large auditorium. The bio-marathon started at 6:00 P.M. and ended at 6:00 A.M the next day. The auditorium was packed with excited students. Guest lecturers appeared in the middle of the night and we showed biological films. It was all about biology and created a motivational learning environment. I enjoyed developing television and radio programs, and this activity became an important component of my career. For more than ten years, I presented a radio program called Druger’s Zoo on WAER-FM 88.3 in Syracuse. The program involved interviews with a great variety of individuals to reveal their life stories and interests. When I was in Australia, I presented a similar program called Druger’s Australian Zoo, on 6NR in Perth, Western Australia. When I was in Washington, DC, as an NSF program officer, I presented a series of radio programs called Druger’s Washington Zoo, involving interviews with Washington personnel. These programs were recorded and sent to WAER in Syracuse for airing. Currently, I present a regular radio program on WAER called Science on the Radio that is intended to enhance the scientific literacy of the general public. I also developed and presented a series of television interview programs called Druger’s Working World on Newchannels Cable TV in Syracuse that were intended to inform students and others about the intricacies of different careers. In relation to my biology course, after exams, I presented The Bio-Answer Show on Syracuse University’s University Union Television Station (UUTV). Each program started with a humorous skit, followed by my going over answers to the exam, followed by drawing names from a fishbowl to award Dollar Store prizes. Currently, I am a regular columnist for 55-Plus magazine in the Central NY area. All of these media presentations were designed to stimulate and enrich the knowledge of the general public and provide them with “adventures in life.”

Pioneers in Science Education    319

I was always interested in the development of new science teachers. It was not uncommon to give pre-service teachers all sorts of courses followed by student teaching. Then, when new teachers actually got a job, they floundered and there was no support network to follow up and assist them in adjusting to the “real world.” In the 1960s, I organized a prototype program for helping new science teachers. The organization was called “First Year Science Teachers” (FYST). Many supportive activities were involved where new local science teachers could be helped by experienced teachers to deal with their teaching problems and get off to a good start in their teaching career. My published article about this program received an Ohaus Science Teaching Recognition Award from NSTA in 1965. In 1985, while I was president of the Association for the Education  of Teachers in Science (AETS), I also served as program officer for the Networks Program at the National Science Foundation (NSF). Being a program officer at NSF proved to be a very valuable experience. I learned the intricacies of funding projects, as opposed to applying for funding as a faculty member. My experience at NSF enhanced the scope of my knowledge about science education in the U.S. When I was president of NSTA, I persuaded Kendall-Hunt Publishers to sponsor a workshop breakfast for new science teachers at NSTA conventions. Experienced teachers would sit at tables with new teachers and discuss problems of new teachers. With modifications, that breakfast workshop has been continued at NSTA conventions. In more recent times, the Amgen Foundation co-founded with NSTA a New Science Teacher Academy to help the professional development of new science teachers. I was also interested in preparation of college science teachers. New faculty members were often dropped into the college classroom without adequate preparation in teaching methodologies. This interest developed from an early teaching experience as a graduate student at Columbia University. When I entered Columbia University as Lecturer in Zoology, Dr. Teru Hayashi was the professor for the evening course I would be teaching. I found him in his lab. He said, “Oh, you’re Druger. Good. Here’s the book.” That was my introduction to college science teaching. I wanted to make sure this cursory introduction to college science teaching would not happen to other new graduate students. Accordingly, in conjunction with my biology course, I developed a teaching assistant training program that involved regular meetings with TAs and a graduate-level course on “The Teaching of College Science.” In this course, tips on college teaching were discussed, and each student selected one important issue in college science teaching, prepared a comprehensive, literature-based paper, and led a discussion in class about the issue. I helped establish the university-wide TA training program that became a leading program nationally.

320    M. DRUGER

I served as chairman of the Department of Science Teaching for about 21 years, and I mentored many students who were enrolled in our program leading to a PhD degree in college science teaching. Teaching adult learners in the Independent Study Program at Syracuse University strengthened my involvement in science education. These students were making sacrifices to earn a degree and were mature, intelligent, and highly motivated, and they demanded good teaching. I taught them, and they taught me. Another influential experience in my career was teaching a biology course to inmates at Auburn Correctional Facility in Auburn, NY. The course was part of an undergraduate degree program sponsored by Syracuse University. I interacted well with the prisoners and came to respect the uniqueness and special talents of each inmate, despite the deviant acts that got them into prison. I wanted to have these students do a pig dissection and I promised the warden that I’d carefully distribute and count all the scalpels and collect them all back at the end of the dissection. The warden’s comment was, “Yes, you’ll collect them all in your back.” I decided not to do the dissection. Thus, in my years at Syracuse University, I was involved in preparing college and high school biology teachers, organizing and directing summer programs for high school students, teaching a variety of courses to adults in the ISDP program, supervising the teaching of my biology course in the Project Advance program, and being involved in many national roles in science education. All of these activities were consistent with my belief that life is short and I wanted to experience as much of it as possible in the service of others. Promotion and tenure were not issues that I concerned myself about, although, as a dual appointment, I had to satisfy the criteria for both the School of Education and the Biology Department. I simply pursued my interests and activities, but, thinking back, I realize now how difficult it must have been to do work in both education and biology to advance. I just did what it took to follow my passion. I just did what I did! Having a dual appointment had its benefits and its drawbacks. A dual appointment turned out to be a triple appointment, where I had roles in the School of Education, the Biology Department, and the Department of Science Teaching. Although this triple position enabled me to interact with research scientists and educators, it eventually proved difficult to cope with being in several places at the same time. Each area expected the same input as someone who was full time in that area. As my curriculum vita demonstrates, I tried to contribute to science education in countless ways, both on the local and the national levels. My national leadership experiences in science education had a profound influence on my life as a science educator. I tried to help these organizations

Pioneers in Science Education    321

function effectively in promoting science education. I focused on helping make these organizations motivational learning environments for teachers and scientists, thus extending the focus of my introductory biology course to the national level. Although I didn’t throw any papers out of a window when president of these organizations, I tried to do unusual things that would be remembered. One of the highlights was the NSTA convention in Philadelphia in 1994. I had a life-sized cutout of me made to stand in the convention center with a sign on it saying, “Welcome to the Convention.” The cutout proved to be a highlight of the convention. The cutout was incredibly realistic, and many people mistook it for me. I recall several teachers standing near the cutout and one teacher said, “OK, at 6:00 P.M. we’ll meet at Marvin.” Many photographs were taken. After the convention, my wife received a doorsized package. It was my cutout. Since then, the cutout has had an interesting history and, most recently, it stands at the entrance to the textbook department in the bookstore at SU. They hang signs on it, and the one I liked the best was, “Give to the Rescue Mission.” ”Druger dollars” and magnifiers were also hits at the convention. Druger dollars had a caricature of my face, and they were distributed widely at the convention. If you brought a Druger Dollar to the bookstore at the convention, you would receive a wallet-sized magnifier, and the hope was that you would purchase several books. The magnifier in a small plastic envelope with my caricature and contact information became my lifetime logo. I distribute them as widely as possible, and everyone I know now has one. When I approach someone in the street, they don’t bother to say, “Hello.” They simply say, “Thanks, but I already have one.” I wrote a series of articles for the Journal of Natural Resources and Life Sciences Education. This was a wonderful opportunity to express the views that I had about science education. Eventually, the essays were published as a book, Practical Perspectives on Science Education. A reviewer cited on the back cover of the book wrote: “Druger’s essays have influenced my personal teaching philosophy to include ‘providing meaningful experiences to enrich the lives of my students.’ This book is a must read for anyone with a passion for teaching.” I never thought that I would retire. However, several circumstances influenced me to retire after 47 years of teaching at Syracuse University. I had surgery for prostate cancer, and the Biology Department was moving to a new building. Also, I had reached age 75, and it seemed to be a good time for a change. I maintained an office with fax, phone, email, computer, and so on and periodically went into the office. I transferred my PhD advisees to other faculty, since I felt they needed an advisor who was “official.” I still teach a small orientation class and have given tours of the campus to

322    M. DRUGER

first-year students and to staff, and, at this writing, my life-size cutout still stands at the entrance to the Textbook Department at the Syracuse University Bookstore. Even though I continued activities at the University (including organizing and directing a Saturday science enrichment program for talented high school students in the local area), retirement proved to be a radical change in lifestyle. At this writing, more than two years after retirement, I still have not adjusted to getting up in the morning and not knowing (or caring) what day it is. I hope to continue to make contributions to society, but I realize that all things must come to an end. As we get older, and there is physical decline, we must develop a mental toughness. At a reunion of my childhood friends, one person said, “Let’s stop talking about the past. From now on, we should only talk about the future. What nursing home do you want to go to, and do you want to be buried or burned.” This joke is not as funny as it once was. My career would not have been possible if not for the support and encouragement of my wife, Pat. I met Pat when she was fifteen years old and I was twenty. She is an incredible woman. Pat supported me by working at a publishing company while she attended Brooklyn College at night. She worked with me for many years as the laboratory coordinator in my biology course and, later, as the course administrator. She then worked as an administrator in the writing program at Syracuse University. Meanwhile, she earned a master’s degree in mathematics from Syracuse University, raised three children, took care of a household, and managed all of our finances, and me. Now retired, she serves as a docent at the Erie Canal Museum in Syracuse, is on the Friends of Jowonio Board of Directors (a local inclusive preschool), does free tax returns for AARP, sews, knits, quilts, and takes care of the household and our finances. To date, we have been married 55 years. I could not have done what I did in my career without Pat at my side. I am the product of her love, companionship, and guidance. Syracuse University established an archive for me in Bird Library. It’s difficult to part with documents and mementos I’ve accumulated over 47 years at Syracuse University. At Pat’s urging, I finally started giving my books to students. Instead of throwing out old papers, I now put them in my archive. So, if I ever need the minutes from a meeting many years ago, they still exist, and I can retrieve them. In retirement, I present Science on the Radio on WAER-FM 88.3 in Syracuse, I organize and conduct a Frontiers of Science Saturday Enrichment Program for high school students. I write a regular column (Druger’s Zoo) for 55-Plus magazine. I have written two poetry books (Strange Creatures and Other Poems and Even Stranger Creatures and Other Poems), and I do poetry readings at elementary schools. I have also written a children’s book (Mr. Moocho and the Lucky Chicken). I have two other recently published books: Practical

Pioneers in Science Education    323

Perspectives on Science Education and The Misadventures of Marvin. Like the biocreativity project for my students, these projects gave me the opportunity to express my creativity. So, my activities continue in retirement. Retirement is a good time to think about what I have accomplished in my long career in science education. Aside from leadership roles, publishing articles and books, and organizing and directing many science education programs for teachers and students, I believe that my greatest achievement was to stimulate more than 40,000 students to think about life in a new, lasting manner. Many of my former students have gone on to professional careers in science, and the feedback from them has been positive and gratifying. I meet former students wherever I go, and they don’t run away. They express gratitude for what I’ve done for them through my teaching. I feel good to know that I’ve influenced so many lives in meaningful, lasting ways. That’s what being an educator is all about. I can’t resist the temptation to include an excerpt of an email that I recently received from a former student: Hello Dr. Druger, You will not remember me personally. I was one of several hundred students in your intro biology course at SU in 1980. Quite some time ago, now that I think about it. I just wanted to sincerely thank you for your efforts as my professor then and to let you know that you profoundly affected my life in ways that I will never be able to repay. It wasn’t so much your teaching (as excellent and fun as it was), it was more your inspiration. You were determined to inspire us to explore science, life science in particular and, indeed, truth itself. And it made all the difference. For me, your inspiration has held my interest my entire life. I became an engaged biology major at SU, went on to medical school, and have enjoyed a satisfying career as an anesthesiologist. Throughout, I have maintained an active interest in all branches of science, subscribing and actively enjoying Scientific American, and whatever other scientific readings I can get my hands on— mostly on the Internet these days. Thank you for inspiring me and countless other students over the course of your career. This funny little blue/green globe of ours is truly a better place because of you.

I’m sure that other teachers get this kind of feedback from former students. When I see a former student take the time in a busy life to write such a communication, it makes me feel appreciated. I once told a high school class that teaching was a great profession, but there was one bad thing about being a teacher: Doctors and lawyers get lots of respect, but teachers don’t. A student quickly responded, “But without teachers, there wouldn’t be any doctors or lawyers . . . or anybody.” Wise words to justify a career in teaching. One day, while driving in my car, I was thinking about what teaching should be all about. The answer struck me like an epiphany. The mission goes well beyond acquisition of subject matter competency and the goals

324    M. DRUGER

I stated above. Our mission as educators is to provide meaningful motivational experiences that enrich the lives of students and help them identify their unique traits and where they fit in life. I think that’s what I’ve done for many students and that’s what I’m most proud of. For Further Reading: Partial Bibliography of Marvin Druger Anderson, O. R., & Druger, M. (Eds.). (2009). The world of protozoa. Arlington, VA: National Science Teachers Association Press. Cook, S., Druger, M., & Ploutz-Snyder, L. (2011). Scientific literacy and attitudes towards American space exploration among college undergraduates. Space Policy, 27(1), 48–52. Druger, M. (1968). The concept of FYST: An association for first-year science teachers. In M. Druger, Practical Perspectives on Science Education (pp. 63–66). Madison, WI: The American Society of Agronomy, Inc. Druger, M. (1970). Using media to individualize biology teaching. Australian Science Teachers Journal, 16, 17–21. Druger, M. (1979). Humanizing the introductory college biology course. Journal of College Science Teaching, 8(3),164–167. Druger, M. (Ed.). (1988). Science for the fun of it: A guide to informal science education. Arlington, VA: National Science Teachers Association Press. Druger, M. (1994). Individualized biology series (Units 1–18). Dubuque, IA: KendallHunt Publishers. Druger, M. (1997). Preparing the next generation of college science teachers. Journal of College Science Teaching, 26(6), 424–427. Druger, M. (2002). It all depends: a perspective on science teaching at all levels. Journal of College Science Teaching, 31(7), 493–494. Druger, M. (2004). Strange creatures and other poems. Syracuse, NY: Druger-by-theLake Publishers. Druger, M. (2007). Our mission in education. Journal of Natural Resources and Life Sciences Education, 36, 159–160. Druger, M. (2010). The misadventures of Marvin. Syracuse, NY: Syracuse University Press. Druger, M. (2010). Practical perspectives on science education. Madison, WI: The American Society of Agronomy, Inc. Druger, M. (2012). Science teacher preparation revisited. Journal of Natural Resources and Life Sciences Education, 41, 65–67. Druger, M., & Allen, G. (1998). A Study of the role of research scientists in K–12 science education. The American Biology Teacher, 60(5), 344–349. Druger, M., & Borgstede, W. (1984). Long-term effects of a student science training program (SSTP). Journal of College Science Teaching, 14(2), 128–133. Druger, M., Siebert, E., & Crow, L. (Eds.). (2004). Teaching tips. Arlington, VA: National Science Teachers Association Press. Lederman, N., & Druger, M. (1985). Classroom factors related to changes in students’ conceptions of the nature of science. Journal of Research in Science Teaching, 22(7), 649–662.

Pioneers in Science Education    325

This page intentionally left blank.

Chapter 16

Luck—A Defining Element of Success Or, How a Few Borrowed Innovations, Time, Effort, and Money Combined with Opportunity and Support, Are Creating Success for Some Professors and Many Students John E. Penick North Carolina State University

I’ve often thought about how my views of teaching and teacher education, and my career, would have been different without the lucky break of having met and worked closely with Professor Dorothy Schlitt at Florida State University (FSU). From working with Dorothy, I came to understand at least six concepts about research and teaching that were significant to my career and, I firmly believe, the careers of many of my students. These concepts can be briefly described as:

Going Back for Our Future, pages 327–349 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

327

328    J. E. PENICK

1. observe, identify, and focus on the roles of teachers and students, emphasizing data collected personally from classrooms; 2. combine these personally collected data with knowledge gleaned from the research literature to identify and develop the implications of those data; 3. use student experience and discussion as a vehicle for teaching; 4. expand class time well beyond the minimum required; 5. expect teacher education students to fully describe their classroom environment and to articulate and explain their teaching goals and roles; and 6. have high expectations in regards to teaching and learning. But, to understand these ideas and how they eventually developed and influenced me and others, you need to know a few things. I Knew I Would Never Teach From an early age I knew I would never be a teacher, as both my parents were teachers. With a few rare exceptions, my teachers were not exciting, especially in science. Beginning in first grade, I rarely paid attention to teachers, as I was more interested in art, numbers, and reading. Immersed in reading, I was always being called on unexpectedly, forced to invent a quick answer to avoid embarrassment. By fourth grade, art had been replaced with science, and I did little except read, experiment, and explore the Everglades, which happened to be where we lived. If I had developed an operational definition of teaching from my experience as a child in school, it would have focused on restating what is known, testing to see if students could repeat it, and then moving on to something else. I could never see myself doing that as, while I was always fascinated by learning new things, I was far more interested in actively exploring the unknown, wondering what was around the next corner or under the water, seeking uncertainty and adventure. Thus, I set out to be a scientist—not just any scientist—I was going to be the next Jacques Cousteau. But, since he never left his marvelous position and the Prince of Monaco did not heed my call, relatively quickly I was forced to find a new career path that, quite unexpectedly, kept intersecting with teaching. Luck in the Form of Teaching Having finished a degree in zoology and chemistry, and not being particularly disposed toward the medical mycology lab where I worked as a se-

Luck—A Defining Element of Success    329

nior at the medical school or the rat urinary physiology I was studying in graduate school, I was looking for options. One unpleasant option looming over me was Vietnam, as I had a deferment only by staying in school. I felt trapped into slogging along, collecting rat urine and feces and analyzing them for diurnal cycles of sodium and potassium. Then, I learned that science teachers were exempt from the draft. And, as luck would have it, since I had completed a year of graduate work in a PhD program in biology, I learned I could become an instantly qualified teacher with a three-year license without any additional preparation, as I had completed many content courses and scored high on the GRE. Equally fortunate, I was offered a position in the same high school where my wife, Nell, taught English. As many people discover, I found that teaching was not as easy as I had assumed. I was fortunate to be placed in a school using what were then experimental programs from Biological Sciences Curriculum Study (BSCS) and with a most enlightened and supportive administration, especially the principal and department head. Although I immediately saw their elegance, the BSCS descriptions of classrooms and teaching were not what I had experienced as a student. BSCS gave me a vision of students and teachers actively learning together in a way that made me excited to think about teaching. I could do this! Discovering the BSCS definition of teaching that relied on inquiry was an early lucky break and an inspiration that gave my teaching a focus it would have otherwise lacked. Unfortunately, as with many good ideas, I had no specific instruction in inquiry (or teaching, for that matter) and distorted or missed many of those key ideas in the process of implementing them. Little did I know at the time, but, like many teachers of that era and later, even though I had a reasonably adequate vision and thought I did all the labs as they were designed, I was doing most of the active inquiry and students were, if I was lucky, merely going along with me. The administration considered me a “good” teacher, but, looking back, I later decided that, rather than because of my fine teaching, I was viewed as good because, in a tough, inner city school, I never sent a student to the office, turned my grades in on time, and always volunteered for whatever was needed. Luck Favors the Prepared Always energetic, after teaching three semesters of biology and a little chemistry while taking two to four university courses each term, I received a consolation prize from my failed PhD program: an MA in biology with an emphasis on teaching biology at the junior college level. I was not yet fully licensed to teach, as I needed three years experience to substitute for student teaching. However, with a graduate degree, more luck intervened

330    J. E. PENICK

as the department head left to be athletic director. Out of 10 teachers in the science department, I was the only one with a master’s, making me the department head in my third year. Not only did this provide an extra $1,000 a year, a large sum at the time, it gave me a reduced teaching load, a leadership position where I needed to have ideas and implement them, and opportunities for personal advancement. I was thrilled and eager for the challenge. Amazingly, almost immediately luck presented itself again. At the end of my second year of teaching, the county held a department head meeting (there were about 20 high schools in Miami at the time) to discuss a variety of issues. Since I had been named as the head for the coming year, I went in the stead of our current department head. Most of the meeting was as I predicted—dull, boring, unnecessary, and tiresome. Then, two professors from Florida State University, one of whom was Dorothy Schlitt, appeared on the stage. They described a grant-funded program for preparing science teachers for which they were seeking to attract top high school students. Their passion was evident as they asked us to help recruit candidates. Their program, they said, was demanding and innovative while providing an unusual amount of individualized attention to the highly qualified candidates. They were sure that, with their program, these students would become first-rate science teachers. But, they needed highly qualified students who could succeed in a rigorous set of science courses. Midway through their presentation, I recall whispering to my neighbor, “Why would I want to encourage an outstanding science student to be a teacher?” Clearly, I was not yet a good teacher. I continued listening, noting that both professors were eloquent, had bold and seemingly powerful ideas and seemed like intelligent, likeable people. Little did I know that what they were talking about was going to be a major force in changing my worldview, and not just about teaching. This transition began as they announced that they would like to invite several science department heads to visit them in Tallahassee for three days to see their teacher education program in action. A plane ride? (I had never been anywhere on an airplane.) Three days out of school? At their expense? What more could one want, even if I was not sold on the idea of teaching as a career? My hand was in the air before the last words left their mouths. Soon, I was in the air and on my way to Tallahassee. Again, luck was with me as I found that Dorothy and her faculty colleagues were far from average in their understanding of science and teaching and in their profound abilities as teachers. I had lucked into a treasure trove of priceless jewels in this faculty. During my three-day visit, they worked me tirelessly, visiting classes, interviewing students, visiting with faculty and even administrators (both the dean and president of Florida State were from science education). I could see that the Science Education department at FSU had some powerful faculty members. Most had

Luck—A Defining Element of Success    331

advanced degrees in science; almost all had worked as scientists in fields such as nuclear warhead research, medical research, and even engineering. They seemed a far cry from the part-timers who were teaching the foundations courses I had been slowly taking to obtain permanent licensure as a teacher. Then, I spent an afternoon with a group of student teachers. Fortunately, by that point I was already convinced the faculty were exceptional and they knew how to help someone become an excellent teacher. I was ready to listen and learn. But even with that preparation, I was not prepared for the level of discourse I heard from these student teachers. I said little, for as unenlightened about teaching as I was, I could tell that those student teachers, mere novices, knew more about many aspects of teaching than I did. I kept quiet and listened intently. Equally discomforting for me (and impressive), the student teachers clearly and easily described the classroom environment, the role of teachers and students, and how to teach, eloquently and in ways that had never occurred to me. Not a total dummy, I heard that one of them was moving to Miami with her husband, who would be in medical school. With no hesitation, I offered her a teaching position (how things have changed since then!) and, with the encouragement of Dorothy Schlitt, her mentor, she accepted. In the same moment, without knowing it, I accepted the role they were seeking, as a recruiter for their program at my high school. Six months later, without a single high school student showing any sign of interest in science teaching, I was feeling rather low. My own teaching was getting closer to true inquiry, but I knew I would not long stay a teacher as salaries, opportunity, and respect all seemed low compared to being in administration. The next step in what seemed a logical career path in education was to become an assistant principal, and I had already taken the required exam and scored high enough to be near the top of the list. It did not, however, look like a very desirable position, especially going to juvenile court weekly. I was not sure what to do. My wife, Nell, who taught English in the same school for three years, had noted that I had become a true evangelist for the program at Florida State. As usual, she had great insight where I had none. “If the program is so good,” she said, “why don’t you go yourself?” So I did, and with great support from her and the Science Education Department. And, the luck, opportunity, support, and work did not end there. Sometimes it Takes a 2 × 4 to the Head to Get My Attention I had no financial support when I left my teaching position in Miami and moved to Tallahassee with Nell and our one year old, Lucas; but, Paul

332    J. E. PENICK

Westmeyer, the department head at FSU, had told me he was confident something would happen, and it did. When I and nine other new, full-time PhD students arrived in late August, he offered me an assistantship that provided tuition and a small salary in return for working 20 hours a week with Dr. Ernest Burkman. By the end of the first semester, Dr. Westmeyer announced that two of our group, Jim Shymansky and I, were among only seven on the FSU campus to be awarded National Defense Education Act fellowships. Luck was still in my corner, and I wanted to take full advantage of the support it offered. Although a fellowship meant I did not have to work, I chose to volunteer my time to work with Dorothy. By then, I knew Dorothy as a serious, stern task mistress and critical intellectual who presided over the secondary science teacher education program that she called QUEST (Quality Undergraduate Education for Science Teachers). She was quick, insightful, verbal, and tolerated no compromise, in class or out, from her ideals and integrity. She was also direct and not hesitant to call me out, as she did quickly. It was exactly what a less-than-insightful person like me needed in order to change my behaviors. And, without a change in behavior, I would not have fully understood or changed my views, much less my behavior. Time on Task: Expand Class Time to Promote Desired Learning Working with Dorothy initially, she just treated me as an older, more experienced student in her methods class. Like many of Dorothy’s ideas, this one stuck with me in later years as I followed this same pattern with graduate students who chose to work with me. As a beginner and unfamiliar with Dorothy’s style, I said little in class initially. As these small seminar classes were not your usual university lecture or recitation sections, Dorothy quickly had me involved in discussions or providing anecdotes from my own teaching. The other students were similarly involved, as all had experience in classrooms. Although I did not initially notice, these teacher development seminars met for many more hours per week than any I had experienced as an undergraduate. Thus, long before I knew of any research related to time on task, I learned an early lesson in program innovation from Dorothy: expand the time with students. Dorothy had done this in a grand scale and in a most opportune manner. The summer before I arrived in Tallahassee, FSU had changed from a semester to a quarter system. This meant changing hours and credits earned for all classes. Under the semester system, Dorothy’s program had a typical schedule where students completed an introductory and almost generic secondary methods course one semester, a science methods class and a

Luck—A Defining Element of Success    333

practicum the next semester, and student taught the third. Dorothy calculated the number of hours involved in these and successfully made the case to the administration that the students needed more practicum time each term. Once that was approved, she noted that with the clock hours needed for science methods and practicum, there was little time for other courses that same quarter. She proposed that in the new quarter system students would take the generic methods with an associated practicum in the first quarter and in the second quarter could only take science methods and the practicum. There was no room for any other courses during the methods quarter. Thus, during the science methods and student teaching quarters, Dorothy had total control of her students’ schedules, leading to the first of many innovations that I might never had thought of on my own. Expanding the time in class and practicum of the teacher education classes I taught has been the single most important contributor to my feelings of success as a teacher educator. How lucky that I met and worked with Dorothy Schlitt! And, since I have noticed that many doctoral students with whom I have worked have followed the same pattern of class expansion in their careers, I have no doubt that others have taken this idea to heart as well. Reflecting on this, at an Association for Science Teacher Education (ASTE) meeting in about 1980, I made a presentation specifically about enhancing the power and impact of science methods courses. “You all say you need more time in methods,” I said, “Why not add an hour a week just by changing the time of your classes, say two days a week from 2:00–3:30 instead of 2:00–3:00?” While some grumbled that they did not control the schedule, most noted that they could, in fact, make such changes easily. All, however, noted that the extra time had to come out of their own unscheduled time. Months later, I got a letter (we used to write letters then!) from a senior colleague who noted something to the effect of, “That extra hour a week has changed my life! I love teaching science methods again! Thanks for suggesting it.” With the new quarter system at FSU, when students registered for the secondary science methods and associated practicum course, it showed up as two courses: a seminar meeting Monday through Friday from 9:00 AM to noon and a practicum from 1:00 to 4:00 PM the same days for 12 weeks. While we did not actually meet all those hours, the 10 to 12 students typically met us each day at 9:30 and usually stayed in a seminar setting until noon. With so much time scheduled for the seminar, there was no rush to complete discussions or ideas. There was plenty of time for group work and analysis (reflecting and modeling the type of teaching Dorothy ultimately desired from her own students) and room for digression without penalty, covering all the topics necessary. This, I soon noted, was the essence of the BSCS inquiry model I had failed to comprehend fully. Seeing the seemingly leisurely pace of the seminar and the huge impact the class discussions had

334    J. E. PENICK

on students (as compared to telling them what we wanted them to know and do) made an indelible impression on me—one that led me to expand the time of every class I ever taught after that point. Teach Via Experience and Discussion As the term progressed, if Dorothy felt a discussion or idea would be better served by a classroom experience, often we would go to a local school that same afternoon to observe something or to try out an activity. This was possible because of some local teachers’ bad habits and was the source of another lucky insight for me. Dorothy had identified several junior high and high school teachers who should not have been teachers, as they seemed to do as little teaching as possible. She could call one of them in the morning and ask, “Can our class come out and do some teaching in your fifth hour class today?” The answer was almost always “Yes, please do!” as long as we would take the entire class for the whole hour; these teachers were far more interested in getting out of a class period than they were in teaching. And we could generally rest assured that what the methods students would do in class, whether interviewing students, having small group discussions, or doing some activities, was probably more educational than the reading, films, and worksheets that were the norm in those classes. Dorothy had found a valuable benefit in her knowing some less than competent teachers. When combined with methods students being scheduled all day, we could almost immediately take advantage of opportunities to demonstrate, test, or practice ideas being raised in the methods class. Thus, in my first semester with Dorothy, I had learned about expanding the time available for my own teaching as well as some ways to place methods students in real classroom experiences as needed. While in those classrooms, Dorothy’s students always spent structured time observing both teacher and students using an instrument called the Schlitt-Abraham Test of Interactive Coefficients (SATIC) that she and Mike Abraham (University of Oklahoma) had created (Abraham & Schlitt, 1973). With SATIC, students learned to observe classroom behaviors and events systematically, consistently, and relatively objectively. In the process, they developed a vocabulary and language that allowed for facile communication about classrooms and teaching. Each time our students worked in a classroom, they carried audio recorders that allowed them to review their interactions with students. And, even in 1971, we videotaped our students many times, even though the three-quarter-inch, reel-to-reel equipment was heavy and difficult to use. Seeing the power of teacher education students developing both the observational skills and the capacity for classroom discourse analysis led me to make videotaping and observational analysis the

Luck—A Defining Element of Success    335

centerpiece of most of my later teaching, as I determined that without a common way of looking at teaching and the classroom, we could not possibly discuss what was happening, much less change it. As a result, each of my students, undergraduate and graduate, learned multiple ways to record and observe classroom roles and spent many hours using the resulting language for discussions and analysis. Dorothy also focused on analysis of teaching, both before and after the class. As a result, she spent a lot of time planning ways to develop her students’ knowledge and teaching and analytical abilities, even though that was not always obvious in her class. She had specific topics in mind for each day’s seminar and a series of messages that she kept returning to, always finding ways to insert these ideas into whatever was occurring. For example, among other concepts, she wanted her students to reflect on and apply what they knew about Piagetian levels of cognitive development, the role of the teacher, use of open-ended questions, and wait time. Although these were rarely the topic per se of the lesson, these concepts, along with many others, were threads that could always be found in Dorothy’s questions, actions, and responses. These same threads were also found in the extensive readings she assigned from research journals. Dorothy expected her students to use the research ideas from the readings and discussion to support their ideas in class. In doing this, she consistently asked students to defend their ideas with what they had read. Dorothy, of course, was a vivid and excellent model of using research in the class discussions, as she was well read, had a keen intellect, and knew how to focus on the dynamics of classroom events. These aspects of her character and work were not lost on me after just a quarter working with her. But I was far from understanding it all or being able to teach as she did. Seeing the Roles of Teachers and Students I went to class excited every day. As I listened to the discussion and observed what Dorothy was doing, I could reflect back on my three years as a high school biology and chemistry teacher, soon seeing multiple reasons for why my own discussions had failed. As a teacher, I had thought of discussions as the process of getting students to answer my questions, usually of the recall variety and to which I already knew the answer. I would then praise those who got what I wanted and, if the correct answer was not forthcoming, I would provide it, usually fairly quickly. Students soon learned to wait for me to provide that answer, saving them a lot of effort. My classroom discussions rarely lasted more than a few minutes. As I learned to observe classroom interactions, I saw that Dorothy, unlike me as a high school teacher, asked open-ended questions with multiple possibilities and then continued ask-

336    J. E. PENICK

ing more questions, expanding the original or using student ideas, working with the students to identify possible solutions and then seeking to home in on the most plausible or useful. Her discussions seemed to have lives of their own; I could not believe how easily she seemed to be able to go on asking new and pertinent questions indefinitely. At the same time, she was helping them hone their verbal and imaginative skills, as her students did more talking than she did. But, we all noted, when she spoke, everyone listened. By the end of the first quarter, I could see much of what she was doing, but I was far from cognizant of the nuances and even further from being able to do it myself. I did, however, note to myself that it was my three years of experience that allowed me to understand the overall concepts that Dorothy was advocating. How could we prepare pre-service teachers to understand the concepts if they did not have appropriate classroom experiences? One way, in addition to practicum experiences, it turned out, was to find ways to obtain their solid commitment to ideas and a need to explicate and expound on them. Have High Expectations I was not immune to Dorothy’s techniques either. Regularly, she put me through the paces, asking me to express ideas, explain them in detail, support, and defend them. And, with equal regularity, I found myself in the same position as the undergraduates, stumbling and contradicting myself. Sometimes I felt like a dog chasing its own tail; although from the pain it seemed as if I were getting in some real bites. Dorothy also confronted me on multiple occasions about my personal transgressions, such as talking too much, being arrogant or authoritative about an idea, failing to follow up on others’ ideas, not paying attention, or, worst of all, acting like a traditional teacher. Once, she wrote me a long, hand written letter, expressing her disappointment in some of my behaviors, describing them in excruciating detail. Then, she confronted me with each accusation in an evening session with six or eight of her top undergraduate students, each of whom had a copy of the letter. It was humiliating, especially since I foolishly tried defending myself rather than accepting that whether I was right or wrong was irrelevant, as she was obviously upset with me and I was the one who had to change. The lesson was a hard one, but it cemented in me firmly the idea of explicit confrontation as one way of changing and educating someone, especially when all else had failed. Once again, Dorothy was effectively teaching by modeling, but this time I was the subject. It hurt, but sometimes learning and gaining knowledge are painful. From my personal experiences with Dorothy and her methods, I learned about the power and role of an influential teacher.

Luck—A Defining Element of Success    337

After two and a half years of 40 and 50 hour weeks with Dorothy, including a full year with her teaching high school biology for one period a day at the lab school, I had gained many new insights, lots of knowledge, and a few skills. I did not feel at all confident of myself as a teacher compared to what I had ignorantly felt three years earlier, but I had certainly changed my ideas about teaching and learning. One clear piece of evidence for this arose in my third year at Florida State. I went to visit my old high school in Miami. There, in my former classroom, was a woman teacher, perhaps 10 years older than me, who was the only student teacher I had supervised in my first three years. Watching through the glass beside the door, I was amazed as I saw her doing many of the things I did while she student taught with me: the way she positioned herself in the long, narrow room; her mannerisms; the apparent flow of the instructional events. I saw lots of science talk, writing on the board, short and ineffective discussions, no seeking of student ideas beyond the factual, rejection of student ideas—they were all there. She had excellent control of the class, but I did not relish the image of a teacher that I saw. She was in too many ways the antithesis of what I had recently learned about effective teaching. After a few minutes, I walked in. She was absolutely delighted to see me, giving me a big hug and smiles and then turning to the class she said, “This is Mr. Penick. He taught me all I know about teaching!” I just smiled and accepted this painful acknowledgment. I had learned a lot and had been confronted one more time. Now, could I do something useful with my new understandings? During those same years at Florida State, I was lucky and supported yet again in that the FSU lab school was under fire from the legislature for not being actively involved in research. This led the elementary science faculty of the lab school to approach Charles Matthews, also on the FSU faculty, to help them. He proposed a large research project that would involve all of the elementary teachers. As luck would have it, he and Dorothy were close colleagues, so he knew of some of my interests and capabilities. Thus, he asked me to join him, Dorothy, Jim Shymansky, and a few others as part of a research team to design studies and collect data from grades one to five science classrooms. I did not realize at the time how special this opportunity was. My dissertation eventually came from this study and the project collected huge amounts of data. While the team collaborated on the analysis, Jim and I did much of the writing, preparing multiple papers for presentation at the National Association for Research in Science Teaching (NARST) and ASTE conferences. This opportunity, another lucky break combined with considerable effort, meant that by the time I went for my first university job interview, I had presented five papers at NARST, two at ASTE, and had several research papers in press with the Journal of Research in Science Teaching and Science Education.

338    J. E. PENICK

Students Articulate and Explain Their Teaching Goals and Roles The methods quarter at Florida State ended with students writing extensive papers that described and justified their goals for students, how they would achieve those goals in the classroom, and how they would evaluate themselves and their students at the end of a unit of instruction. Dorothy referred to this as a Research-Based Rationale for Teaching. I had never seen or thought of such a thing and, even after my first three years of teaching, could not have produced what she expected of her students before student teaching. I was amazed at the depth of scholarship she was able to coax out of undergraduates. I knew this was going to be part of my future teaching. But the best and most impressive was yet to come. Rather than a final exam at the end of the methods quarter, each student met with Dorothy for an hour or more to discuss and defend his or her rationale paper. Dorothy directed me to sit in the back of her office and listen, admonishing me not to say a word and to remain out of the student’s line of sight. In each instance, I had read the paper, marking up each with ideas, questions, and suggestions, much as I might have done with any term paper. What Dorothy did with them was more time consuming but much more powerful. As I listened to those sessions, I came to see how Dorothy led the student to explain the ideas in the paper and then justify them, preferably with ideas from the literature. As the defense unfolded, even though students might write clearly about significant concepts, I was often shocked by the shallowness of their apparent understanding when asked to expand and deviate from the written script. Even seemingly simple concepts like the use of open-ended questions could, during the ensuing discussion, be seen to suffer from a lack of understanding. From listening to students in class or reading their papers, I would have rated them much higher than what I saw when they were explaining their papers while one on one with Dorothy. Later, Dorothy and I discussed how in such a high-stress environment it was easier to determine what students truly understood and conceptualized as compared to what they could say when they had time to prepare each response, such as in a paper. I saw that investing 20 or more hours in listening to students one-on-one was an excellent use of faculty time, as each student had a chance to be heard and Dorothy could not only evaluate, she could instruct individually at a moment ripe for learning. After watching a number of these sessions, I began to see more and more of Dorothy’s strategy as she used students’ written ideas to help them identify internal contradictions in their own papers, evidence of not seeing or understanding the big picture. Typical of beginners, many of these students were somewhat concrete operational. They often focused on only one variable at a time.

Luck—A Defining Element of Success    339

So, for instance, a student might emphasize asking open-ended questions as a way to get more students to respond, but not mention using adequate wait time. Or they might suggest using overt praise or evaluation, neglecting the research they had cited earlier about how evaluation inhibits divergent responses. While some of these conflicts were purposefully created as “gotcha” moments (especially with difficult-to-influence students), Dorothy usually used them as a way to get a student’s attention, to focus on a new understanding. I also saw how writing down their ideas led to a form of ownership and commitment to an idea that was much stronger than when they were merely expressing an idea in class. Once they owned the idea, they felt compelled to defend it, even though they may not have really thought it through. I definitely had lucked into a new technique for assessing and improving student learning simultaneously, albeit a time-consuming, difficult, and sometimes stressful technique. An Aside: My First Professional Presentation Dorothy also gave me another opportunity that had long-term consequences. In the winter of 1971, she was scheduled to present a short paper at the Southeast Regional meeting of ASTE. It was a short paper, perhaps seven or eight pages at most, describing QUEST, the undergraduate science teacher education program she had developed. Dorothy had come to dislike attending professional meetings, as she felt they mostly focused on classical research and traditional ways of teaching. She wanted to discuss innovations in science teacher education and supervision, while many of the papers at meetings focused more on traditional, quantitative research models. She did not want to go or present the paper. But, she had agreed and was on the program. Like most of the department, I was planning on attending the meeting. Then, with no prior notice, she informed me that I would read her paper at the meeting in Jacksonville, Florida. I knew the material well, as I had been working closely with Dorothy in the program. It seemed an easy task. I was a good reader and I was comfortable with her words and phrases. I was nervous but fully prepared. The venue was the usual large hotel, but, as it was a small meeting, all participants were in one room and all listened as each paper was read in a published sequence. Perhaps the lowest ranking speaker among the 10 or so in the morning session, I was scheduled at the end. We each had 15 minutes though, plenty of time. Then, of course, people ahead of me took more than their allotted time, as, in addition to their reading, they added impromptu embellishments. About midway through, the moderator whispered in my ear that time was short; “Can you do it in 10 minutes?” “Of course,” I said with certainty, figuring I could read a bit faster and skip a

340    J. E. PENICK

small introductory section. Then, everyone started pointing out the window and we all rushed over to see a few snowflakes falling and gathering on the windowsill. Snow in Florida! Another five minutes evaporated while we marveled at this rare sight. Meanwhile, I rehearsed my 10-minute presentation, cutting here and there. Shortly before my turn, the moderator informed me that I only had five minutes, as we had to get to lunch on time. With that, I did a fortunate thing; I tossed the paper aside and, rather than reading the paper as was the norm, I looked at the audience and told them the story and message Dorothy wanted to convey. I did not stumble, and it was well received. We not only got to lunch on time, but I learned a most valuable lesson: present information via story telling—don’t read to the audience what they could read for themselves. Luck was indeed with me that day. As the result of that experience, I went from being afraid of public speaking to relishing it, as I found it easy and fun to look at the individuals in the audience and tell them the story I had in mind. Now, all I needed was more experience and more stories. Applying What I had Learned Accepting a position as the only science educator at Loyola University of Chicago in 1973, I thought it would be just like Florida State, and I expected to imitate the Florida State science teacher education program there. This was my first experience with the need for professional support and opportunity, and I needed it when I found out they did not even have science education, just a generic secondary teacher education program. While I found that I could do almost anything I wished within the confines of my own classes, my modest innovations did not provide a coherent program that would educate students as Dorothy was doing. Changing a program meant involving other professors, gaining their commitment, and asking them to make substantial alterations to business as usual. At the beginning of my second year at Loyola, I had added a little time to each of my classes and even with 30 students in a generic, secondary methods class, all students wrote and defended a rationale paper each semester. Then, Bob Yager invited me to participate in a small science teacher education conference at the University of Iowa in the fall of 1974. There, I met Vince Lunetta and, equally important, I saw that the Iowa pre-service science teacher education program had not only a large science education faculty, it had many of the same program elements as Florida State. Equally important, Bob had a significant grant for developing a model teacher education program and Vince had the experience, energy, vision, and initiative to create a program, rather than providing a mere series of courses.

Luck—A Defining Element of Success    341

With the grant and with Bob’s leadership creating opportunities, Vince had maneuvered so that Iowa science education students had their own sections for foundations courses, an early practicum course, and two science methods courses, one for life science and one for the physical sciences. Students could take either one prior to student teaching for a full semester. The Iowa conference focused on these aspects, and I again saw the benefits of both a large faculty (like FSU, Iowa had nine full time science educators) and money. Near the end of the conference, Bob announced that he was looking for a new assistant professor for the fall of 1975. I returned to Loyola knowing that Iowa was an excellent fit for me and my ideas. Optimistically applying, I made the finals, got interviewed, and in April learned that I had finished as number two. Accepting my fate, I prepared additional plans for Loyola, bought a new TV for our home, and kept writing, something made much easier by my work with Jim Shymansky at FSU that had continued as he went to Iowa and I went to Loyola. Less than two years after I completed my PhD, together we had already published two papers each in JRST and Science Education, as well as in several other journals, had more in press, and had made 13 presentations at NARST. My career was going well. Then, in May 1975, Bob Yager called and told me that their finalist was not accepting the position, “Would you be interested?” he asked. With no hesitation Nell, Lucas, Megan, and I were on our way. Arriving in Iowa City, I found that Vince, who had much more high school teaching experience than I did (and was more successful, I suspect) was only two years ahead of me as a professor. But, his maturity, leadership, and continual thinking about teaching and programming were immediately apparent. We spent many hours discussing how to use the opportunities that funding provided to build a successful experience for our students. And, like at Florida State, we were blessed with a number of full-time graduate students who were themselves vital in our efforts to create a dynamic model of science teacher education. While a critical factor, having talented graduate students like Ron Bonnstetter (retired from Nebraska and now a vice president at TTI Performance Systems Ltd.), Bill Kyle (University of Missouri-St. Louis), and Gerry Foster (DePaul, deceased) was in itself a challenge, as they were bright, eager to try things, and I had little idea how to use them as a staff. My only experience was with Dorothy, so I approached it from that perspective. This led to an apprentice model with the graduate students participating in the methods classes as students (although with teaching experience), having some supervisory experiences with student teachers, and freely offering ideas for innovation. Since our graduate students often had more teaching experience than I had, I took advantage of that (like Dorothy) to have them discuss their situations, ideas, and to lead small group discussions related to teaching and learning.

342    J. E. PENICK

Vince and I spent a lot of time talking and thinking through our goals and opportunities. One observation we made quickly was that most of the pre-service students would be licensed in the life sciences but would have a minor area in the physical sciences. Thus, those students really needed both science methods courses, although only one was required for licensure. Similar to the Dorothy Schlitt model, we proposed that all students begin the sequence with the introductory practicum that was already in place and that we add a seminar for introducing classroom observations and planning small group lessons. This seminar and practicum would give students opportunities to collect, analyze, and review observational data related to teaching and learning. The second semester, students would enroll in life science methods, the third semester would be physical science methods, and the final semester would be student teaching. This was a relatively easy change to make, creating in 1976 what we then informally called Science Methods I, II, and III (they still retained their old names along the lines of “Methods and Materials for Life Science” etc). Initially, Methods II and III each met for one day a week for two and a half hours, satisfying the minimal time requirement for three semester-hours credit. Again, considering Dorothy’s admonition to “make big changes when you can,” I proposed that we could expand the time for each class. Vince, always an innovator, ready to do whatever was necessary to educate our students, agreed quickly and we changed Methods I to a 90-minute seminar with a weekly practicum experience. Methods II would meet Tuesday and Thursday for two hours each day, and we scheduled Methods III for Monday and Wednesday for two hours each day. As Dorothy told me several years earlier and as I had observed, all it took to make these changes was to change the time on the schedule of classes that we were handed each year for editing and proof-reading. No one ever questioned our changes. Having more time allowed us to be more relaxed in class and still deal with all the issues we wished to cover. I also added six days of intense practicum to Methods III, where students taught a three-day lesson in a junior high school while being video taped. We would then analyze the tapes in class, and students would revise their lessons, teach the same lesson to another group for three days, and again review the tapes. Wisely, Vince wanted still more practicum time, leading to creating an extensive practicum for Methods II, this one in either a junior or senior high school. Soon, however, being in a small town in Iowa, we were having trouble placing all of our students in suitable classrooms. Again, thinking a little differently, we moved the Methods I practicum to elementary classrooms, as there were many available. We also felt it was appropriate for our students since they were assigned to interview students and use Piagetian tasks to analyze student thinking. At the same time, elementary science activities were easier and less threatening to our students

Luck—A Defining Element of Success    343

than preparing lessons for highschoolers. We also found that our students were welcomed and had great status with elementary teachers, many of whom had little formal science coursework. The Methods II practicum met in a school for one class period per day for the entire semester, giving our students a more extensive clinical experience. In just a little over two years, we had gone from the innovative program I saw in 1974 to an exceptional one in 1978, where teacher education students had more than 120 hours in classes to develop their expertise as teachers, 100 hours experience in elementary and middle level classes, and a full semester of student teaching. While Vince and I might have come up with all of these ideas alone, I knew Dorothy was always whispering ideas in my ear. While most of our students were traditional undergraduates, we always had a few post-graduate students asking about our programs. About 1979, we were authorized to create a Master of Arts in Teaching (MAT) program that would lead to licensure. While many in our college felt they would have to create new methods and foundations courses for these MAT graduate students, we saw no need. Graduate versions of educational psychology and other foundations courses were available, and we already had senior level numbers on many of our courses. Since beginning master’s students were allowed to take senior level courses, we could generally use our existing courses. Following the same pattern as with undergraduates, we developed a four-semester curriculum that would lead to an MAT at the end. Equally useful, those students who transferred to us as seniors could graduate at the end of Methods II and then complete Methods III and student teaching as graduate students. While not leading immediately to a degree, the year of graduate work gave them an advantage over many students when applying for jobs and, in Iowa and some surrounding states, often led to a higher starting salary. Equally interesting, having older, beginning graduate students in the classes raised the bar for the undergraduates with whom they were in class. Nor did graduate students receive a diluted graduate experience (as some have predicted). It was a win-win situation for all, including the professors, as the entry of graduate students in our methods classes escalated the level of intellectual dialog, enhanced visible student commitment, and provided new role models for the undergraduates. And we had more fun. At the same time, Bob Yager’s continual vision and exceptional leadership led to another NSF grant that allowed us to create a series of applied science courses for our teacher education students. These courses, created as we developed the teacher education program, were to be capstones, a place where our students would apply what they had learned in the science departments by examining science concepts and related societal issues. As a faculty, collaborating with advanced graduate students who were experienced science teachers, we developed science applications courses in

344    J. E. PENICK

environmental science, life science, chemistry, physics, and earth science. Since science education was officially the Department of General Science in the College of Liberal Arts, we could assign science course numbers to the applications courses. With these capstones, science teacher education students now registered for nine separate required courses with science education faculty as instructors. With as much exposure to the views of our faculty as to faculty in the sciences, our students truly had a double major in science and science education and we, as a faculty, had multiple opportunities to educate and influence them. While our transformed program was a boon to the education of our undergraduates, it was equally powerful for our graduate students, as they could see a working model of an innovative and comprehensive program in action, they had more opportunities to interact with us and students, and we had specific roles for them. Many of these roles were in the form of apprenticeships, where graduate students worked through our program as students, assistants, and, sometimes, instructors. The most committed and available doctoral students (not all were on teaching assistantships) would spend several semesters in the methods courses as assistants and then teach Methods I as the primary instructor for a semester or two, under the supervision of a faculty member. These apprenticeships provided a powerful mechanism for educating graduate students in aspects of teacher education that are often lacking in education doctoral programs. While the Iowa science teacher education program quickly became more extensive than what Dorothy created and involved far more faculty, the intensive apprenticeship experience closely mirrored what I had experienced in Tallahassee. And, with more courses to teach, the Iowa program was capable of impacting far more students at all levels. Another lesson learned from Dorothy was expanded and applied. A Few Other Lessons Applied I had not forgotten the other lessons learned from Dorothy Schlitt and her colleagues. I typically taught Methods I and III and Vince Lunetta taught Methods II. I required all my students to write a research based rationale paper in Methods I and then revise it during Methods III, creating a threesemester opportunity to develop and organize their thoughts, almost like a journal where they could explore and refine their ideas. And, of course, each semester they sat down with me, one on one (plus a graduate student or two) to defend their papers, usually for 75 minutes. While this is time consuming, the students clearly made progress in learning to express their ideas clearly and to justify them with support from the literature. Several evaluations of our program indicated the power and effectiveness of our

Luck—A Defining Element of Success    345

program, including those by Penick (1979), Ron Bonnstetter (1984), and Joe Krajcik (Michigan State) (Krajcik & Penick, 1989). After each evaluation we made changes, some small and some large. By 1981 Methods III, a three semester-hour course, had expanded again and was meeting in seminar Monday and Wednesday for two and a half hours each day, twice as many hours as a typical three-semester hour course. By 1986, we added Friday for another two and a half hours, giving us seven and a half hours each week in this seminar. When students asked why this course met three times more than was required, we had interesting discussions about the nature of excellence, the extra attention each individual received, and how time on task was related to learning. More astute students noted that science education classes were not only small (no more than 15 in a methods class), but each was taught by a senior faculty member, and each class usually had several graduate students in attendance, adding to the instructional power. About this time, we learned that an audit had determined that science education, rather than being an inexpensive program instructionally, like most in the College of Education, was among the most expensive on campus, comparable to engineering. Rather than being chagrined, we were delighted, as we knew science education deserved excellence and needed intensity and that cost and effort were related to the quality of the teachers leaving our program. The science teacher education program at Iowa functioned well, with changes here and there, until that day about 1991 when Al Colburn (now at Long Beach State University) pointed out that our sequence would be more effective for students if Methods II and III were reversed in the sequence. Typical of Al, his argument was flawless. Clearly, changing the sequence would provide more coherence, put more practicum experience where it was needed (right before student teaching), and strengthen our students and our program. But, our students were in a four-semester, lockstep program, where each cohort passed together to the next phase. It couldn’t be done without severely impacting many of our students. “Great idea,” I said, “but we can’t do it.” Typical of Al, he kept at it, eventually figuring out a way to make the switch by having one semester where we had some overlap. His plan was solid, and he was willing to do a lot of the work. With no hesitation, we made the big switch and in the process officially renamed the courses as Science Methods I, II, and III (looking back, we wished we had named them something more meaningful and less mechanical, such as “Science Teacher Development I, II, and III”). By 1992, our students were spending 195 clock hours in science teacher development seminars during their four semester programs (for seven semester hours credit where only 87 class hours were required). These extra hours added immeasurably to our abilities to educate more fully. At the same time, practicum experiences had been increased to 132 clock hours prior to student teaching, including

346    J. E. PENICK

45 hours in an upper elementary classroom, 12 hours in a middle level class as teacher in charge, and, with Methods III, one hour a day for an entire semester in either a middle or high school classroom. Each of our students was videotaped while teaching and then reviewed more than 12 times prior to student teaching. Thanks, Al, for a marvelous idea! As another interesting example of innovation not related to Dorothy (but one she would have recognized and applauded), Bob Yager pulled off a most interesting feat. As part of his statewide teacher development effort called Chautauqua, Bob made a formal proposal to the State of Iowa. In this proposal, he suggested we pull teachers from classrooms during the school day so we could provide them with local and regular professional development. As teacher in-service was difficult for many teachers in the many small, rural schools of Iowa, it was best done during the school day. Bob’s proposal was for us to place a student teacher in the classroom for the semester who would then take over while the teacher was participating in a professional development activity about once a week. While the state was sympathetic, Iowa law required that a licensed teacher be in each classroom, and student teachers were not yet licensed. Bob, in his resolute manner, solved this by noting that many of our pre-service students were post-baccalaureate and thus eligible for a substitute’s license. Bob also noted that the state had provision for “experimental licensure” to test innovations. With these ideas, Bob gained permission for our students to be the teachers of record while student teaching, if the regular teacher was not present. The school districts liked this so well our student teachers got paid as substitutes on the professional development days. Bob, as usual, went further than perhaps necessary, but, as I had seen with Dorothy, this is often how real innovation occurs. At that time in Iowa, all licensed teachers had to have a “student teaching experience” at the grade level for which they were licensed as well as appropriate subject matter expertise. Bob made the case that our students spent a semester teaching at each level and were, thus, eligible for elementary, middle, and high school science licensure. While the state approved this, seemingly with little trouble, our colleagues in elementary education were not happy. Several said, “If you are going to license your students as elementary teachers, we will license ours as secondary teachers!” I, for one, encouraged them, but it seemed no elementary teachers wanted to complete the 48 semester hours of science courses necessary for this. Instead, some elementary teacher education students with strong science backgrounds (usually former nursing students) transferred to our program. Not a way to win friends, perhaps, but a great addition to our program. To me, the takeaway was to always seek value-added components, as the K–12 licensure became a central part of our recruitment efforts. Even students who had no interest in anything except teaching high school science could see that having

Luck—A Defining Element of Success    347

additional licensure would not hurt when applying for jobs. Many noted that the additional licensure areas could have positive implications for career advancement in the future. Some Final Thoughts By the time I left Iowa in the fall of 1997, more than half our pre-service students came from states other than Iowa, more than half took jobs outside Iowa, and quite gratifying, they usually had multiple job offers. Each fall, our placement office provided a list of all Iowa graduates who had taken teaching positions that year, including at the university level. Amazingly, many of our MAT students had the highest starting salaries among the group, placing ahead of university positions, as many went to suburban Chicago schools where salaries have historically been high. Many of our graduates won awards from state and national organizations, and a number have completed PhDs, including Mike Clough (Iowa State), who replaced me for a few years when I left Iowa. Our teacher education program provided an excellent education for many students, and I give Dorothy Schlitt considerable credit for that. At the same time in 1997, we had a waiting list three semesters long to enter the four-semester program, as we were still holding to a 15-student limit per entering class. Since most of our students by then were post-baccalaureate, this effectively meant that students were willing to wait seven semesters to obtain licensure even though they could go to many other institutions and finish in as little as 16 weeks. Equally significant, our entering pre-service graduate students had higher mean GRE scores than did our doctoral students. And, new doctoral students working with me usually knew far less than the second- or third-semester pre-service students about current research related to teaching. My take on this has not changed: high ability teacher education students do not want the easy way out; they want quality and the opportunity to excel. The other side of that same coin is that a high-quality, demanding program will attract high-ability students who are willing and eager to be part of a highly intellectual teacher education program. Why is teacher education seen as the easy, less intellectual path? Perhaps, unfortunately, because in many institutions it is. My experiences at Florida State and Iowa made it quite clear that the more we demand of our students, the more effort they put forth and the more they succeed. As department head at North Carolina State for more than 10 years, I interviewed dozens of prospective faculty members. Invariably I asked them, “What are some characteristics of your ideal teacher education program?” Equally invariable were their responses, as all were unprepared for the question and

348    J. E. PENICK

usually described their own preparation and failed to mention anything particularly innovative. Shades of the rationale defense: none ever mentioned more time on task, even though they probably would have waxed eloquent about “time-on-task” as an important variable for K–12 learning. Numerous teacher education and doctoral graduates from Iowa have told me over the years how having an explicit, written rationale for teaching science provided them with an exceptional power to impress during interviews. Receiving calls from school principals, asking, “Can you send me another one?” made my day and confirmed the power and elegance of what we were doing. Unfortunately, too few teacher education students (some of whom ultimately become professors of science education) experience the same intellectual standards and challenges in their education courses as they routinely find in their science preparation. Without extensive knowledge and challenging experiences, our current students, undergraduate and graduate, are not receiving a full education. Think of the sciences—most in-depth graduate learning is in the lab, not the classroom, and students in innovative labs are more likely to succeed in their careers than those from more traditional settings. Over the years I have always asked graduates of our programs about their interview experiences. Almost without fail, both undergraduates and graduate students have reported to me that in their interviews they were asked about educational innovations. Most were actually concerned that they had too much to say, but hoped they would be sensitive and aware enough to be low key and not overwhelm (and frighten) the interviewer. I am proud of where our graduates went, what they have done, how they have taught, and how they have succeeded. I treasure their successes that have allowed me so much success at the same time. Dorothy Schlitt can take pride in her continuing legacy. References Abraham, M. R., & Schlitt, D. M. (1973). Verbal interaction: A means for self-evaluation. School Science and Mathematics, 73(8), 678–687. Bonnstetter R. J. (1984) Characteristics of teachers associated with an exemplary program compared with science teachers in general. Unpublished dissertation, University of Iowa, Iowa City, Iowa. Krajcik, J. S., & Penick, J. E. (1989). Evaluation of a model science teacher education program. Journal of Research in Science Teaching, 26(9), 795–810. Penick, J. E. (1979). Formative, descriptive evaluation of the Iowa-UPSTEP model. Technical Report 17, Science Education Center, The University of Iowa, Iowa City, Iowa.

Luck—A Defining Element of Success    349

This page intentionally left blank.

Chapter 17

Why I Became a Teacher J Myron Atkin Stanford University, Emeritus

In 1932, like almost all other children in Brooklyn about five years old, I began school at the kindergarten in the local elementary school. Mine was Public School 210, across the street from the two-bedroom apartment in which I lived with my parents, my younger brother, and my Uncle Julius. Schoolwork seemed interesting to me, and I liked my teachers. Neighborhoods in New York City tended to be homogeneous communities of about 20 square blocks. With rare exceptions, everyone was of the same race, religion, and income level—in my case, White, Jewish, and poor. The nearest high schools, which covered a wider geographic area, were only marginally different. There was more diversity, but the informal student social groups tended to mirror those of the student’s immediate neighborhood. But I didn’t go to the nearest high school. The New York City public education system has long had a handful of specialized high schools that today would be called magnet schools. They include schools for the performing arts, the fine arts, design, and technology. Students were and still are selected by competitive examination, and they can live (and commute) from any of New York’s five boroughs. My high school, Stuyvesant, specialized in science and mathematics. It required daily trips of about 40 minutes in each Going Back for Our Future, pages 351–365 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

351

352    J M. ATKIN

direction, including the subway ride and the walking required at each end. In my case, it was from the Crown Heights neighborhood in Brooklyn to East 15th Street between First and Second Avenues in Manhattan. Stuyvesant is the oldest specialized high school in the City, having opened its doors in 1904. I entered those doors in 1940 after completing ninth grade at 210. The school was an eye-opener in many ways to a 13-yearold. The school had students of all races and religions, who had interests and abilities like mine. There were no girls, though. It wasn’t until 1969 that girls were accepted for admission to Stuyvesant. I thrived at this large school of about 2500 students. Again, I studied with teachers who knew their subjects well and took pride in the ways they taught them. I learned later that they were among the best in the City and highly selected. Furthermore, and crucially in terms of my ultimate choice of a career, they had secure jobs right through the Great Depression (unlike my father who was unemployed for 14 months in the late 1930s). I was elected president of my senior class and served as associate editor of the Stuyvesant yearbook. About 500 boys graduated in June 1943, and there were two similarly sized graduations a year during that period. I was barely 16 years old when I graduated from Stuyvesant in June 1943. Elementary school students with high marks skipped grades during the 1930s, possibly as a way to save public funds during the Depression, though I don’t know that for a certainty. It was the height of World War II, and almost all of the 18-year-olds in my graduating class were drafted immediately. A close friend was killed in the Battle of the Bulge. I was admitted to the City College of New York and started my studies there a week or two after my Stuyvesant graduation. CCNY, established in 1847, also was tuitionfree and funded by New York City itself. Again I was impressed by most of my teachers. For the most part, they were well educated, dedicated, intelligent, and seemed to enjoy their work. And, for the most part, I liked and respected my fellow students. There was the same amalgam of students as at Stuyvesant, except that it was co-ed. When I returned to college in 1946 after 13 months in the Navy, I completed my degree in chemistry and graduated in 1947. I had attended CCNY during summer sessions and also had received nine semester-hours of credit in electrical engineering because I had completed the Navy’s excellent program for training electronics technicians. To please my father, I then applied for admission to medical schools in New York City. My chemistry grades were OK, but not good enough for Jews—especially from City College, for whom there was only a small quota at all the City’s medical schools—and I couldn’t afford to live away from home. So I entered the graduate program in science education at privately funded New York University that prepared me for a license as a science teacher.

Why I Became a Teacher    353

Actually, I couldn’t have afforded the tuition at a private university, even a local one; but I had the benefits of the GI Bill, plus a small scholarship from the State of New York that was solely for veterans. All this enabled me to earn a PhD on a part-time basis over a seven-year period while working full-time as a teacher, and taking advantage of a large number of course offerings during the summer. None of my formal education, kindergarten through the doctorate—which I consider to have been excellent—would have been possible without funds from one governmental agency or another. Only on reflection many years afterward did I realize how profoundly my decision to enter the teaching profession at elementary and secondary levels and my subsequent career at two universities were a result of (a) the Great Depression: teachers had secure jobs, and economic security was a priority; and (b) anti-Semitism: Jews from New York faced high barriers to entering medical school. Absent these factors, I almost certainly would have become a physician. Entering the Teaching Profession My first full-time employment after my brief Navy service was as a science teacher. I took a comprehensive examination to teach science in New York City’s junior high schools and passed. In retrospect, the examination is worth describing because I don’t believe there is anything comparable today. First, of course, I had to have a degree in one of the sciences. Then I took a three-hour written exam in science and principles of education. A three-person panel then interviewed each person who passed. The panelists sat at the rear of an empty classroom while the candidate stood at the front. A strong accent was a disqualifier, but I don’t know how pronounced one’s New York dialect had to be for him/her to be rejected. Each of the surviving candidates was then required to appear at a specified junior high school a week or two later to teach a lesson to an actual class. In my case, I was asked to show up at 9:00 AM at a school in Long Island City in the borough of Queens. On arrival, I was told that at 10:30 I was to teach a demonstration lesson about convection currents as part of unit on weather. To assist in my planning, a laboratory assistant (!) was assigned to set up the equipment I thought I would need. He was also helpful in telling me about what the students already had studied. There were three examiners at the back of the room taking notes. The written examination was passed by 125 people. Of this group, 26 passed this performance test. We each received a license to teach science in New York City public junior high schools at a starting salary of $2,700 for those with a master’s degree, $2,500 for those without. I had the master’s.

354    J M. ATKIN

It might be noteworthy that teacher certification by the New York State Department of Education was not necessary. Decades earlier, the City had obtained a waiver from state requirements. I believe other large cities (Chicago, Buffalo, Philadelphia) were judged to have more rigorous standards than those at state level and had similar exemptions. I did not accept the appointment, however. While considering it, someone at NYU told me that Ramaz School, a Jewish day school on the Upper East Side of Manhattan, was extending its well-established K–8 focus to the high school level. The school was looking for a teacher to launch the science department by teaching general science in ninth grade and biology in tenth grade. A new science subject would be added each year as the first group of students moved to succeeding grades: chemistry in eleventh grade and physics in twelfth. The salary scale was identical to the City’s public schools. I interviewed for the position and was offered it on the strength of my college record and recommendations, plus the fact that I had passed the test for teaching science in the City’s public junior high schools. As a teacher in a school that was just opening, I had the task of teaching general science and, unusual for a new teacher, of deciding on the biology curriculum for the tenth graders. Developing the Curriculum I started by choosing a biology text published in 1946 titled Biology for Daily Living. The authors were Ernest Bayles and R. Will Burnett. (When I later went to the University of Illinois as an assistant professor in 1955, I met Burnett. He became my mentor and good friend.) Biology for Daily Living covered the topics and principles expected in a basic high school textbook in the subject at the time, including cell structure, classification, heredity, reproduction, taxonomy, ecology, and evolution. However, ecology, and especially conservation, was at the heart of the book. My students learned about the scorching heat and years-long droughts throughout the plains and upper Midwest during the mid-1930s. They learned about the resulting dust storms that destroyed crops, impoverished farmers, and forced hundreds of thousands of destitute families to leave the land and head west. The drought extended to the southern mountain states. The human side of the Dust Bowl tragedies was portrayed vividly in popular novels like Steinbeck’s Grapes of Wrath, which was made into a popular and award-winning film staring Henry Fonda. What caused these social upheavals and human suffering? How might they have been mitigated and perhaps prevented?

Why I Became a Teacher    355

And so my students learned that farmers had turned the land upside down. Native grasses were buried because prices for wheat were high. It was terrain that was a desert even under normal weather patterns. All these hardships unfolded against the background of the Great Depression that they themselves had lived through during their own childhood. For more than a few, their parents had been unemployed for months, even years. What needed to be done? My Manhattan students learned about crop rotation and contour farming. A key concept was that it was a national priority to reinvigorate the land and keep it productive. Conservation was a core theme of the textbook. A national Civilian Conservation Corps (CCC) was created that employed more than two million jobless young men to reshape and replant the terrain. The CCC lasted until the early 1940s when the services of adult males were conscripted to meet national needs elsewhere in the world. And the book highlighted the role of the federal government led by a committed and popular president. Elementary School Science While teaching full time at Ramaz, I pursued a doctorate Saturdays and summers at New York University. There I worked primarily with two professors: J. Darrell Barnard and Rose Lammel. Barnard headed the science education department at NYU and chaired my doctoral committee. He focused much of his writing on children’s problem solving in science. That topic became the heart of my dissertation. I served as a teaching assistant to Lammell for a course on elementary-school science education, and my interests gravitated toward children from kindergarten through grade six. As a result, I resigned from Ramaz and applied for and was appointed a K–6 science “consultant” in the Great Neck Public Schools, a suburban community just outside the New York City limits on the north shore of Long Island. At the elementary-school level, the district was heavily committed to the kinds of curriculum for children associated with the progressive education movement and centered on the writings of John Dewey. Children investigated the world around them. They worked in groups to solve problems. Subjects were “integrated,” which meant that the problems they worked on drew for their elucidation on community issues in the primary grades and on the different subject fields (reading, writing, social studies, art, and science). Since teachers were not necessarily strong in all these subjects, consultants were available for subjects like science, home economics, industrial arts, music, and visual arts. Our goals were to help the teachers become sufficiently competent in these subjects to handle it on their own. In fact, about 70 percent of my time was spent in demonstration teaching, by ap-

356    J M. ATKIN

pointment. The remaining time was in conference with individual teachers, leading workshops in science, and collecting instructional materials. Inasmuch as the goal was to make the classroom teacher more self-reliant, I didn’t demonstrate the same lesson twice. Rather I might teach it jointly with the classroom teacher the next year. Essential to the success of this approach to what was basically an in-service teacher education program was the fact that I and all the other consultants were on the same salary schedule as the classroom teachers. In some cases, all I had to do was provide the books and equipment. The books were seldom texts but rather trade books for children on the topics we were studying, usually written by professional writers of such books. I learned a great deal about teaching from these extraordinary colleagues during the five years I taught in Great Neck—and about the limitations of textbooks, too. Both of these experiences stand out for me 70 years later as being particularly influential as I tried to find my own professional paths over the decades to follow. Entering the Professoriate Soon after entering the doctoral program at NYU, I was asked by Darrell Barnard to join him and Celia Stendler, a professor of child development at the University of Illinois, in developing a series of textbooks in science and health for elementary-school children. They had enlisted Dr. Benjamin Spock, the best-known pediatrician and advisor to new parents of the late 20th century, to serve as an advisor. They were also seeking writers for each grade level and asked me to write the books for fifth and sixth graders. By 1955, I was almost finished with my work on this enterprise, and also with my dissertation. Celia Stendler asked if I might be interested in being considered for an assistant professorship at the University of Illinois. I was, I visited, and I was offered the position. The senior professors in science education at Illinois were R. Will Burnett (the author of the biology text I had chosen for students at Ramaz) and Herb Zim, a prolific author of science books for children, including the Golden Nature Guides that were (and are) essentially identification books for all ages about the biological and physical worlds. No one could ask for more prominent and knowledgeable mentors. At Illinois, I taught courses about elementary-school science teaching that were required of all students preparing to become elementary-school teachers. I also taught a seminar on research in science education for PhD candidates. Universities were expanding rapidly in those days, due partly to the influx of veterans with GI Bill benefits, partly to the anticipation of an enrollment surge resulting from a sharply increased birthrate in the

Why I Became a Teacher    357

post-war years, and additionally because higher education was attracting an increasing number of women. A person with a PhD and a good pulse rate could advance rapidly. I had both, and I did. In 1957 I was advanced to an associate professorship, a position with tenure. In 1960, I was promoted to full professor. In1970, I was appointed dean of the School of Education. In the early 1950s, the Carnegie Corporation of New York, under the leadership of John Gardner, had supported a curriculum- development project at the university called the University of Illinois Committee on School Mathematics (UICSM). It turned out to be a modest label for what was to become a national initiative that snowballed and shaped the teaching of mathematics and science in secondary and elementary schools for decades to come. The force behind UICSM was Max Beberman, a math teacher at the University of Illinois High School and an assistant professor in the College of Education. Max, working with Professor Herbert Vaughn in the university’s Department of Mathematics, took the view that the prevailing curriculum did not teach important mathematics, by which he meant mathematics as understood by the world’s leading mathematicians. Thus UICSM introduced topics like set theory and number bases other than 10. UICSM was the forerunner of dozens of projects that eventually covered all subjects in elementary and secondary school mathematics and science. In 1956, the Physical Sciences Study Committee (PSSC) was launched at MIT with support from the National Science Foundation. The project was under the leadership of physics professor Jerrold Zacharias. Then, after the launch of the first artificial space satellite by the Soviet Union in 1957, the NSF expanded its education portfolio dramatically, initially with projects in chemistry and biology and eventually in all science and mathematics subjects taught in school. UICSM was essentially the model: Get leading university-level scientists and mathematicians involved in curriculum construction and teacher education below the college level. Working with Stan Wyatt and Bob Karplus below High School Level: “Discovery or Invention” Elementary school science was to prove problematic for the NSF. Providing leadership and funding for the development of new curricula for science taught in high schools seemed a logical extension of the foundation’s education portfolio. NSF’s founding legislation when created by the U.S. Congress in 1950 specified that the agency’s function was twofold: support both scientific research and science education. In the early years, the latter meant support of graduate studies in science. After Sputnik, it seemed sensible and feasible to expand the NSF mission to include improvement of secondary

358    J M. ATKIN

school science. Furthermore, the chairman of the foundation’s board at the time, James Conant, was the leading advocate of the expansion. Conant, a chemist and former president of Harvard University, had devoted several years to investigating the state of American high schools and was committed to their improvement. Nevertheless, he opposed the NSF’s moving into elementary-school science and mathematics. He reasoned that there were about 50,000 teachers of science in American high schools at that time, and NSF-supported summer and year-round institutes at hundreds of campuses every year could reach them. Additionally these people were usually full-time science teachers. Elementary school science, on the other hand, was only one of many subjects in the curriculum—if it was taught at all. Furthermore, while high school science teachers usually spent their working lifetimes in teaching, the half-life of elementary-school teachers at the time was about three years. That is, half the group left teaching after a short time. Teachers at that level were usually female, and they frequently left the profession early to start a family. One element of the counter-argument by NSF staff was that it was almost the same 50 per cent involved in the turnover, and the remainder had no more turnover than high school teachers. Furthermore, many who left on maternity leave returned to the profession when their own children started school. NSF decided to support a study to be conducted by the American Association for the Advancement of Science to make recommendations about the desirability and feasibility of expanding its work to include elementary and junior high school science teachers. Major expansion of NSF’s activities at these levels was the result. At about the same time, NSF made a cautious and exploratory entry into the new territory. A grant was awarded to the University of Illinois to support a summer workshop in 1959 for about 35 people whose responsibility was to improve science teaching in elementary schools. Some were science supervisors for both secondary and elementary schools. Some had responsibility for all elementary school subjects. In all cases, they were the key persons with responsibility for science teaching in their schools or school districts. The eight-week summer program featured work in astronomy and ecology in the morning and seminars on science education curriculum and teaching in the afternoon. Astronomy was chosen because I had developed a collegial relationship with Stanley Wyatt, a professor of astronomy at the university with long-standing interests in the teaching of astronomy. Funds were included in the grant to cover my travel during the school year 1959– 1960 to visit the summer participants and report on the effects that the summer programs seemed to have had.

Why I Became a Teacher    359

A second grant was awarded to the University of California, Berkeley, to support the work of Robert Karplus, a young and brilliant physicist, who was beginning to turn serious attention—and the remainder of his career—to the teaching of physics in the primary grades. As NSF grantees, Karplus and I were invited to join the frequent meetings in Washington of NSF curriculum development project leaders to exchange ideas and help the foundation plan new initiatives. In 1961, I took a sabbatical leave to work with Karplus in Berkeley. One result was to coalesce the deep reservations that both of us had developed to a growing and popular belief that science in the elementary school should be taught by providing young children with the opportunity to “discover” basic science principles. Karplus took the view that it was fanciful to believe that children could “discover” major concepts like Newton’s laws of motion or Copernicus’ heliocentric solar system. However if such concepts were presented to them, they could then collect data that would help them see how the basis for such concepts could reasonably be questioned, defended, and confirmed. Tycho Brahe had collected more complete data than Copernicus about planetary motion, yet he could not accept the fact that planets revolve around the sun. Newton’s Laws seemed to fly in the face of everyday observations that a ball will stop rolling after someone gives it a start. So we found ourselves introducing concepts like parallax, energy, and friction—then helping students “discover” some of the manifestations. Students addressed questions like, “Is the book on the seat of a car moving?” and “How can the retrograde motion of Mars be explained if we assume a heliocentric solar system?” Evaluating Science Curriculum Innovations By the early 1960s, additional projects below high school level were funded by NSF. A grant was made to some of the people closely associated with the PSSC group to start a new and independent effort, the Elementary Science Study, under a new and independent entity called the Education Development Center. A second grant went to the American Association for the Advancement of Science (AAAS) to launch Science, A Process Approach (SAPA). Not unreasonably, there was a call to justify the rapidly expanding efforts to change the school curriculum. The federal government was spending increasing millions of dollars to make significant changes in mathematics and science teaching. What was the evidence that new curricula were fulfilling their promise? Were students learning to view science more as scientists do? Capitalizing on this development sooner than most, AAAS asked Robert Gagné to develop an evaluation of SAPA. Gagné, a distinguished psychologist who had helped to designed evaluations of technical programs for the

360    J M. ATKIN

Air Force during World War II, approached the task by asking what students should be able to do, just as he approached his Air Force assignment. What processes do scientists themselves employ? He and his colleagues asked such questions of eminent scientists. Answers? We observe, we measure, we record, we draw inferences, we test our ideas, we experiment, and we subject our ideas for criticism. So SAPA built its curriculum around such behaviors. Let’s observe corn popping. What do we hear? What do we smell? What do the kernels feel like before and after? A hierarchy of observations was developed that emphasized these and other observations. I was concerned. Where’s the science? After all, its methods are directed toward helping to understand larger ideas about how the physical and biological worlds work. Yes, scientists observe, measure, record, and infer. But these methods are utilized to discern and articulate broad and testable principles. Process and content are intertwined and not neatly separable. This concern led to my writing a long letter to Science magazine, the AAAS journal. Called “‘Process’ and ‘Content’ in Grade Schools,” it appeared as a full page in a 1966 issue of the magazine. This statement was a further iteration of my growing concern about the application of social and behavioral science methods and principles in comprehending and making judgments about the quality of educational programs. In the early 1960s, evaluators of new curricula were exhorted to define the aims of their program in behavioral terms. What should a student be able to do as a result of learning that light can be seen as both a wave and a particle? What can a student be able to do in demonstrating the role of variation and adaptation in the survival of a species? Identifying the behavioral objectives of a course, at the outset, became the watchwords of the education evaluator. This approach seemed problematic. Does it provide adequate knowledge on at least two counts? For one, all courses have outcomes that are not predictable. Some are long term, others short. Some are desirable, others not. (Does the course encourage the student to continue further study? Does it lead the student to dislike the subject?) A program evaluator has an obligation to examine all noteworthy results. Michael Scriven, a philosopher of science then at Indiana University, coined the term goal-free evaluation to emphasize the point that educational programs should be judged by all their outcomes, not only those stated as objectives. Another reason this approach seemed problematic was that teaching to predefined objectives, especially those that will appear on statewide tests, leaves little room for taking advantage of unpredictable learning opportunities. It would be pedagogically impoverished and educationally constrained if a ninth grade science teacher ignored an unexpected seismic or meteorological event, or if a biology teacher did not devote time to the once-ina-generation drought that has beset the farming community in the county.

Why I Became a Teacher    361

My essay on the subject was published in 1963 in the second issue of the Journal for Research in Science Teaching. It was titled “Some Evaluation Problems in a Course Content Improvement Project.” As I write these words, I wonder whether there is room for essays any more in the Journal—or in social science publications generally. The editors often seem to model their publications on their conceptions of research in physics. These ruminations foreshadowed what was to become a parallel career in education evaluation. In that work, my inclination has been to try to fathom and prioritize the perspectives of the classroom teacher and project director. What do these people value? How can I highlight practical reasoning, reasoning directed toward action? This extension of my own intellectual development led me to scholars like Stephen Toulmin and Martha Nussbaum and to the intellectual and practical importance of action research in better understanding and trying to improve educational practice. This shift, which developed gradually and tentatively, was fortified by my extended associations with education evaluators like Bob Stake, Barry MacDonald, and David Jenkins—as well as with related work on action research, especially that of John Elliott. Undergirding much of this orientation is scholarship in fields like philosophy, the history of ideas, and even fiction. These humanistic perspectives supplement those that a social scientist brings to attempts to understand and improve educational practice. This expansion of perspectives opens doors for greater focus on, and appreciation of, the efforts of those closest to the educational scene. The OECD Projects: The Relevance of Developments Abroad In the 1970s, I began to get involved in the education initiatives that had begun to surface in the Organization for Economic Cooperation and Development (OECD). The OECD was created as an extension of the Marshall Plan for European recovery after World War II. As the benefits of collaboration became evident, the organization expanded to include not only Western but also other industrialized democracies that could benefit from cooperative policy-oriented analyses of issues of common concern. The United States, Australia, and Canada were early members. Gradually Japan, several nations in South America, and other industrialized democracies were added. By the late 1960s, the OECD created a new unit: the Center for Education Research and Innovation (CERI). It was launched with a substantial initial grant from the Ford Foundation. I was asked to assist in a study of teacher-

362    J M. ATKIN

education programs in some of the member countries. A colleague at the University of Illinois, Jim Raths, agreed to join me in this modest effort. In 1986, after having served by then as education dean at Stanford for seven years, I was asked by Bassam Shakhashiri, a chemist at the University of Wisconsin, to come to the National Science Foundation, where he was now head of the education unit, as his advisor. The opportunity could not have come at a better time. My major responsibilities for 20 years had been in administrative work at Illinois and Stanford. I agreed to come for a year, while on leave from the university. Accepting the NSF offer proved to be an extraordinary opportunity to reorient myself to developments in science education. At about the time I was taking up full-time professorial responsibilities at Stanford after leaving the deanship, David Thomas at CERI asked me to help him plan and direct a new project that would examine innovations in science, mathematics, and technology education in member countries that wished to participate. Thirteen countries joined the project. Each member of the Steering Committee took special responsibility for liaison with the coordinators at the country level. My graduate assistants and I had the opportunity of becoming particularly knowledgeable about the projects in the United States, Spain, Germany, and Japan. By then, we had attracted Mary Budd Rowe to Stanford, and she was of particular help with some of the American projects. Each of the 13 countries chose the innovation it wanted to highlight. Case studies were the method of choice. The country-level project directors constituted the Steering Committee, which met regularly. Very soon after the project was launched, David suggested that Paul Black of King’s College London, a physicist who became deeply involved in science education policy and practice, be asked to co-chair the committee. Ultimately, 21 cases were written and published locally. Paul and I prepared an overview of the entire project at its conclusion (Atkin & Black, 1996). Senta Raizen took responsibility for assembling the American cases (Raizen & Britton, 1996). Looking Backward—and Forward— with Paul Black While I had been engaged deeply in program evaluation for many years, I knew very little about the assessment of individual students when I first met Paul. He is, as many readers of this book know, deeply involved in helping the scholarly community and key policy makers understand the distinction between assessment for purposes of making judgments about the quality of

Why I Became a Teacher    363

a student’s work and assessments that also help students to enhance their learning—and, importantly, productive relationships between the two. In 1999 during a joint residency at the Rockefeller Foundation’s Bellagio Center, where we were completing a book about our respective careers in science education (Atkin & Black, 2003), we received word that the National Science Foundation would approve our proposal to support research on how teachers can improve their classroom assessments of students so as to improve science learning. The grant to Stanford provided for a subcontract to King’s College London, where much of the foundational work had been done (Black & Wiliam, 1998). That project resulted in additional publications, including a report of the work my students and I had been doing in several San Francisco Bay Area middle schools (Atkin, Coffey, Moorthy, Sato, & Thibault, 2005). Where Do I Go From Here? The actuaries say that a person 85 years old has a better than even chance of living to 90. If I’m lucky enough to do so, and my health holds up, the project I see for myself seems clear. And it may be as much political as professional. All public expenditures are under scrutiny these days. They should be, of course. But a special focus seems to be on the public schools, and it has turned venal. In today’s polarized political climate, that attention seems to translate into the demonization of teachers. I am keenly aware that public-school teachers provided the only education available to me. But even before entering kindergarten, I benefited from New York City’s free baby clinics (that had been established in the 1890s) for the check-ups and immunizations I required as an infant. Then taxpayers supported my schooling from kindergarten through the PhD. Taxes paid by the citizens of New York City paid for me to go to school until graduation from high school. Then I attended the City College of New York, for which the City had been paying the full cost of qualified residents’ education since 1847. My master’s degree and PhD were almost entirely funded by the GI Bill. I use a litmus test to gauge the merits of any education policy: Is it more likely or less likely to attract the country’s most able people to the profession? Fortunately it often does, despite the relatively low salaries and usually lessthan-optimal working conditions. But when elected officials loudly accuse teachers as a group as misfeasors, the choice of a career in the classroom becomes problematic enough to fear for the future of public education. One way to increase the attractiveness of the profession—and thus improve the quality of education itself—is to involve teachers more directly

364    J M. ATKIN

in efforts to improve their own practice. How? Increase opportunities for collegial interactions. For example, provide more time for teachers to discuss their work. In many European countries and in Japan, teachers work directly with students for about 20 hours per week. In the United States, the figure is closer to 30 hours. The time when teachers are not in the classrooms is devoted to joint planning, discussions of curriculum, and—especially in Japan—observing their colleagues teach. The observations are discussed later with the teacher. Often, the teachers decide to modify the lessons and try a different approach. They begin to experiment more in their own classrooms. “Lesson Study” is the label used in Japan. Elsewhere, particularly in Europe and in an increasing number of American schools, the teachers also test their new ideas and discuss the results. These forms of professional development are sometimes called action research, and the results are occasionally published in professional journals. Teachers also might, and sometimes do, participate in the evaluation of other teachers. Peer evaluation is part of the rigorous process of the National Board for Professional Teaching Standards that leads to certification (along with interviews and the preparation of a portfolio of professional accomplishments). It also can be one element in the evaluation of classroom teachers for tenure. These reflections lead me to the possibility of trying to play a more public role in policy discussions. I haven’t been shy about speaking to and writing for professional groups trying to highlight these issues. I also have testified before Congress. Some of my thoughts appear on the Internet. Occasionally I have written a letter to the editor of some journal or newspaper. Maybe I can satisfy my sense of collegial and social responsibility by continuing to do just what I’ve always done—but now within a rising sense of a deepening unfairness in American society and a deep determination to try to do something about it. References Atkin, J M., & Black, P. (1996). Changing the subject: Innovations in science, mathematics, and technology education. New York, NY: Routledge (with the Organization for Economic Cooperation and Development, Paris). Atkin, J M., & Black, P. (2003). Inside science education reform: A history of curricular and policy change. New York, NY: Teachers College Press. Atkin, J M., Coffey, J. E., Moorthy, S., Sato, M., & Thibault, M. (2005). Designing everyday assessment in the science classroom. New York, NY: Teachers College Press. Black, P., & Wiliam, D. (1998). Inside the black box: Raising standards through classroom assessment. Phi Delta Kappan, 80(2), 139–148. Raizen, S. A., & Britton, E. A. (Eds.). (1996). Bold ventures. Dordecht, The Netherlands: Kluwer Academic Publishers.

Why I Became a Teacher    365

This page intentionally left blank.

Chapter 18

Don McCurdy The “Overachiever” Peggy Tilgner University of Nebraska-Lincoln

Introduction When I contacted Don about writing this chapter, he protested, “I’m not a pioneer. If anything, I’m an overachiever” (D. McCurdy, personal communication, March 22, 2012). When one reviews the list of Don’s accomplishments, perhaps he was not a pioneer in the sense of revolutionizing science education, but rather a leader in introducing literally hundreds, if not thousands, of students to inquiry methods of teaching science. In addition, he was a leader in helping science teachers create integrated science curriculum materials that bridged the discipline islands. His 45 years of service to NSTA is a record of which to be proud. Of equal importance is his impact as a mentor to many science educators. This is illustrated by Don’s recent visit to Lincoln when two former students, an associate dean and the state science consultant, rearranged their busy schedules just to be able to spend a few minutes visiting with Don.

Going Back for Our Future, pages 367–378 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

367

368   P. TILGNER

His journey toward becoming a leader in science education circles is like many others, long and filled with numerous side trips. Much of the journey was undertaken with thoughts on how this would affect his family. Being a good provider was paramount, something that was evident in the way he mentored his graduate students. We begin this narrative with a brief history of the events leading up to Don McCurdy taking a position as a faculty member in Teachers College at the University of Nebraska in Lincoln, NE. Don’s father was a highly respected teacher. In fact, he was Don’s only teacher through the 8th grade in a small country school in eastern Missouri. Don attended a small high school in the St. Louis area and entered the University of Missouri to study chemical engineering along with many GIs returning from WWII. He soon realized that it was the chemistry part of chemical engineering that intrigued him more than the engineering part, so he switched his goal to becoming a science teacher. He graduated in 3½ years and then completed his master’s degree in another year while waiting to be drafted for the Korean War. Upon completion of his master’s, he took a position in north St. Louis teaching a split class of 5th and 6th graders. This was his introduction to teaching diverse urban populations from low socioeconomic backgrounds. He finished the school year, something two previous teachers had failed to do. That summer he reported to Lackland Air Force Base in San Antonio where he taught classes in Officer Candidate School—not science but Air Force Administration and then Effective Expression. Upon completion of his Air Force service, Don returned to public school teaching at his alma mater in St. Louis County, MO. Here he had the good fortune to be paired with Charles Lackinger, a more experienced science teacher, who believed children should be taught science with the scientific method. In one of his units, students designed experiments to test advertising claims with appropriate controls and variables. In another unit, students tested superstitions to determine if there was any scientific validity behind them (McCurdy, 2011). With a growing family, it seemed wise to work toward a degree in educational administration in order to become a principal. However, Don soon discovered that science teachers in California made more money than principals in Missouri, so the family headed west. It was at this time that the launch of Sputnik spurred the National Science Foundation to begin funding teacher training grants to bring teachers’ science content up to date. The institutes paid a summer stipend so teachers didn’t have to find other summer employment. Don took the opportunity over the next three summers to learn more about bacteriology, astronomy, and geology. From the geology course in New York City, Don became acquainted with what is now termed place-based education while making geologic maps of Manhattan.

Don McCurdy    369

The next educational adventure was a year long NSF Academic Year Institute at New Mexico Highlands University in Las Vegas, NM. The courses taught at the institute demonstrated how sciences are interrelated rather than isolated silos. One particular course stood out, Radiation Biochemistry, as it connected physics, chemistry, and biology. It appears this idea of integration of disciplines had an impact on Don, because many of his funded projects throughout his career dealt with integration of science disciplines and integration of science with the social sciences. With another master’s degree in hand, Don began teaching introductory chemistry and physics at Mary Hardin-Baylor College in Texas. It was an enjoyable experience teaching undergraduate students, but it soon became apparent that to survive at the college level, one needed a doctoral degree. After three years, the McCurdy family again pulled up stakes, and Don entered the doctoral program in science education at the University of Missouri (McCurdy, 2011). In the summer of 1967, with his EdD in hand, Don took a position at the University of Nebraska in Lincoln. His first assignment was as a generalist in the Curriculum and Instruction Department in Teachers College. In 1969, Don transitioned into a faculty position as a science educator when his colleague, Dr. James Rutledge, moved to a position as an associate dean in the Graduate College. It was Dr. Rutledge who encouraged Don to become involved in NSTA, an involvement that has continued to his present service on the Nominating Committee. Integrated Science Curriculum Don’s experiences teaching elementary students, high school science, and college chemistry, as well as his participation in NSF-funded programs, played a major role in one of the problems in science education that Don worked on throughout his career, the integration of the science disciplines. Sometimes we artificially draw lines to separate the science content and process. I felt this was especially significant to the teaching of chemistry and physics. Thus after a few years at the University of Nebraska, I submitted an application for a grant from the National Science Foundation to develop an individualized and integrated two-year course in chemistry and physics. The grant was funded and we were able to bring together a group of outstanding teachers of chemistry and physics to work on the development of modules that could be pursued individually by high school students. This program was identified as the Nebraska Physical Science Project which was used for a number of years in many Nebraska schools and a few out of state. (personal communication, May 2, 2012)

370   P. TILGNER

The materials were intended to be self-paced units used by high school students, although some were used for whole-class instruction (McCurdy & Fisher, 1969, 1971). Much of the research from this project focused on the ability of students to work as self-directed learners. In related work with the Intermediate Science Curriculum Study (ISCS) students found that older students were more self-directed than younger students. Not surprising, younger students who had had some experience with individualized learning fared better than those engaging in this type of learning for the first time (McCurdy, 1975; Wood & McCurdy, 1974). Individualized competency-based curriculum was also the basis of the Nebraska University Secondary Education Project (NUSTEP) that Don helped develop and directed for several years. The project consisted of four spirals made up of multiple learning tasks. Each learning task included instructions for working toward the objective, performance tests, and learner supports. The intent was to include practical experience along with methods instruction, principles of educational psychology, and fundamentals of American education in one course. Follow-up assessment showed that most areas of secondary education showed an improvement of student attitudes about teaching. Students completing the project also viewed inquiry methods as more motivational for their students (Witters & Sybouts, 1972). As a student in the project, I found the instruction to be hard to internalize as a student teacher, but found it to be extraordinarily useful in an actual classroom setting teaching science. I was able to use the inquiry methods with my students and can verify the increased levels of motivation as I have employed inquiry teaching methods for 40 years. Don’s belief in individualized activity-based curricula led him in the early 1980s to serve as a designated trainer for the Individualized Science Instructional System (ISIS) as part of the National Diffusion Network. ISIS was intended to provide an alternative to the traditional textbook approach and has morphed into twenty-seven activity-centered modules oriented to the problems of modern society. The modules, which can stand alone or be assembled into an integrated science curriculum, included core concepts in social studies, language arts, and mathematics (Florida State University Office of Research, 2012). Don felt that some students who needed to be pushed to reach their potential might be better served with a instructional system that allowed them to progress at a faster rate than their peers. Don’s philosophy about integration of disciplines extended to mathematics. He states, I was always concerned about the application of mathematics to science. Too often mathematics is taught as a theoretical subject without application to other disciplines. A good example is the teaching of calculus without atten-

Don McCurdy    371 tion to its numerous applications to physics. Calculus is much more relevant if it uses examples from physics. Another area that needs attention is the applications of physics and chemistry to biology. In my opinion one needs to have a good background in physics and chemistry to thoroughly understand biology. Yet, the normal sequence in high school is biology followed by chemistry and then finally physics. (McCurdy, personal communication, May 5, 2012)

While integration of mathematics and science remains an important focus at some major universities, the transfer to elementary and secondary schools is spotty at best. And the sequence of science taught in most secondary schools remains the same as set forth by the Committee of Ten in 1893—biology, chemistry, and physics. Science for the real world Don was also concerned about the lack of real-world connections to the science being taught in the schools. There seemed to be too much reliance on textbooks and memorization of terms rather than “doing science.” This led to submission of several grants to work with teachers in developing teaching materials that not only integrated science disciplines but also crossed over into the realm of social sciences. One such project funded jointly by the U.S. Office of Education and NSF was the Lincoln Area Environmental Education Project (LAEEP). Science and social studies teachers were paired to develop lesson plans and assessments for secondary classrooms on a variety of environmental issues such as solid waste disposal, noise pollution, and chemical runoff and groundwater. These units included print and audio-visual resources as well as lab activities teachers could do in their classrooms. Each unit had both social studies and science outcomes included. This was 25 years ahead of the standards movement! A teacher could take the unit plans and teach them as constructed or modify them by selecting elements that worked best in his/her particular situation. Another project was the University of Nebraska Energy Education project, which consisted of five summer institute workshops for secondary science and social studies teachers. Don was trying to help teachers deal with the 1970s energy crisis in a way that would help their students make informed decisions about energy use and alternative energy sources. The teachers in the institutes were introduced to a wide variety of materials that were also developed into a curriculum the teachers could use in their own classrooms. The energy crisis of the 1970s “ended,” and soon energy issues were no longer receiving high priority coverage in the news. Teachers found they were less able to connect the materials to student lives and stopped updating and using the materials. One of the strong features of the program was that there was a balanced approach including materials

372   P. TILGNER

from the energy companies as well as those groups concerned about the environmental impacts from extracting and using different energy sources. Lack of motivation among science students was a driving force behind many of the curriculum projects Don undertook. He felt that teachers would be better able to motivate students if teachers had the necessary tools such as questioning strategies, understanding how to develop cognitive dissonance in students through the use of discrepant events, and use of the learning cycle for planning. He also felt that introducing students to a variety of curriculum materials would also help teachers motivate their students. In order to be able to provide his students with as many resources as possible in making the curriculum relevant, Don served on a number of advisory boards. He was a member of the advisory board for Chemistry in the Community (CHEMCOM) for six years. He felt that a lot of teachers really didn’t know how to connect principles of chemistry in context. These curriculum materials were in sync with what he believed would help more students succeed in the physical sciences. Chemistry in the Community has just released its 6th edition textbook, suggesting that the project is an ongoing success at teaching chemistry using real-world problems. Don was a member of the Board of Directors for the Nebraska Outdoor Encounter Program from 1976 to 1983. This may not seem to be a significant item on a curriculum vita, but it points out another way Don was able to provide resources for teachers to use in making content more relevant to their students. Outgrowth of this program included introducing thousands of teachers and now pre-service teachers to acclaimed environmental curricula Project Learning Tree (PLT) and Project WILD at little or no cost to workshop participants. I think it is clear that environmental awareness was important to Don in the types of activities he engaged in to create an environmentally literate teaching population. Don was concerned that his students have the opportunity to experience the latest in science content as well as having access to new curriculum materials to teach the latest scientific information. In the mid-1980s he brought several Chautauqua Short Courses for Science Teachers to UNL to update Nebraska science teachers in the area of molecular biology. We learned about recombinant DNA technology, had the opportunity to learn how to do electrophoresis, and also were presented with learning materials that helped us teach our students about the ethical issues involved with this new technology. These are all common in today’s high school biology curriculum but were cutting edge 25 years ago. The strength of these programs was that the content was delivered using best teaching practices and in a manner that was easily transferable to the biology classroom. Don’s interest in providing science teaching resources for science teachers extended to elementary teachers as well. Don’s wife, Mary, was instru-

Don McCurdy    373

mental in Don becoming involved in teaching methods and providing resources for hundreds of elementary teachers. Mary McCurdy taught elementary grades, but she became so involved in science education that she was elected to the Board of NSTA representing elementary teachers and was very active in the Council for Elementary Science International (CESI). She was honored by having an endowed lecture named for her, which is delivered every year at the NSTA national conference, as well as a scholarship in her name by CESI. Mary played a large part in how Don structured his elementary science methods classes. Students learned science content through the use of discrepant events, questioning strategies, and lab activities. They put the science content learned into practice planning and teaching lessons to their peers and to elementary students in the local schools under the guidance of Don’s graduate students. Once students left Don’s elementary science methods classes, they had more appropriate science content knowledge and a set of inquiry tools for teaching that content at a developmentally appropriate level. His model has served as a template for the 40 plus elementary science methods classes I have taught over the past 25 years. Many of us who did our graduate work under Don McCurdy learned that a key part of the success of an elementary science program was engaging school principals in learning about teaching elementary science. I was not surprised when I saw that Don had been a disseminator for an NSTA project on Promoting Science Among Elementary Principals. To enhance the elementary science methods program, Don became director of a field test site for the newly developed Full-Option Science System (FOSS). These inquiry-based modules contained all the lab materials and software teachers needed to teach a wide variety of concepts commonly taught in elementary school science. The program continues to evolve, with much of the material now being found in a web-based format (Fossweb.com, 2012). Don didn’t stop with improving the quality of science teaching in elementary and secondary schools. He teamed with Dr. David Brooks, chemistry professor at UNL, for Training in Education for Assistants in Chemistry (Project TEACH). The two professors teamed up to develop a set of eight modules that helped graduate teaching assistants (TAs) improve questioning skills, tutoring skills, and interactions with students in small groups (Brooks, Lewis, Lewis, & McCurdy, 1976). TAs were videotaped approximately once a month and were provided feedback using Flander’s Interaction Analysis in order to improve their instruction (McCurdy & Brooks, 1979). This program won a Gustav Ohaus Award for Innovative College Teaching from NSTA, the first of two Ohaus Awards Don received during his career. Don’s talent as a teacher was recognized not only by his students, but also by his peers across the university. He was awarded the Distinguished Teach-

374   P. TILGNER

ing Award, an honor reserved for only the very best teachers across the University of Nebraska’s multiple colleges and departments. The UNL Parents Association also awarded him a citation for Outstanding Contributions to the Lives of Students. He holds the Wesley C. Meierhenry Professorship in Science Education in recognition of his efforts in science education not only as a college teacher but also as a science educator impacting teachers across the state and the nation. Work in NSTA Don is proud of the teaching awards he has won, all of the programs he worked on to help science teachers keep current in the field, and the many graduate students he has mentored. One of the most satisfying accomplishments might have been his contribution to NSTA. Don first got involved in NSTA at the urging of Dr. James Rutledge at the University of Nebraska. This was the beginning of 45 years of continuous service to NSTA that continued until 2012 as a member of the Nominating Committee. Included among the many NSTA committees on which Don has served are the Building Campaign Committee, the Finance Committee, the Committee on College Science Teaching, the Student Programs Committee, and the Retired Teachers Advisory Board. Don also served as the NSTA representative to Section Q of AAAS. Don was elected president of NSTA in 1980. He was president-elect when Bill Aldridge was selected to be the Executive Director of NSTA. It was clear to me that Bill had studied our organization well and knew our problems and how to solve them. The organization was having severe financial problems and he had insight into what needed to be done to solve them. We worked together extremely well. He did most of the preparation for a testimony that I was to give before a U.S. Senatorial Committee on Appropriations in 1980. This involved asking Congress to continue the support of science education at the time that President Reagan was attempting to eliminate the Science Education Division of the National Science Foundation. Our efforts and that of many others resulted in the continuation of these programs. (personal communication, May 5, 2012)

Bill Aldridge wrote, “Don was a calm voice of reason on issues who listened to and respected the views of others. He guided NSTA through its most difficult years when he was the president of the organization and may well have saved the organization from disintegration” (personal communication, June 6, 2012). During Don’s year as president of NSTA, he and his wife Mary took leaves of absence from their positions and traveled approximately 20,000 miles

Don McCurdy    375

on behalf of NSTA. They visited 100 schools, 15 colleges and universities, 11 science resource centers, and five science museums. In addition, they attended 17 state conventions and two area conventions, speaking at 30 group meetings and making contacts with science teachers in 42 states (McCurdy, 1981a). Clear goals for the organization were developed at this time including appointing a new Horizons Committee (on which he later served) to chart the future of NSTA. Joint work with AETS was begun to develop a set of appropriate standards for preparation of elementary, middle, and junior high science teachers. The Society for College Science Teachers became an affiliate of NSTA during this time (McCurdy, 1981a). One of the more important accomplishments during Don’s tenure as NSTA president was his testimony before the Subcommittee on Science, Research, and Technology of the Committee on Science and Technology of the U.S. House of Representatives. At that time the Office of Management and Budget proposal was to lower the share of total NSF budget for science education to 6.4 percent and eventually eliminate funding for science education in NSF. The decision was not made with input from the director of NSF, the National Science Board, or the Science Education Advisory Committee. Don pointed out that the “documented failures and deficiencies in American science education can be attributed in part to the erosion of funding support for science education at NSF over the past 15 years” (McCurdy, 1981b, p. 50). Don enumerated the number of NSF-funded science education programs that had helped improve the skills and knowledge of many science teachers, such as NSF Chautauqua-Type Short Courses. He concluded his testimony by spelling out the areas where funding needed to be focused to improve the quality of science teaching in the United States. These included upgrading the scientific, mathematical, and technology competencies of teachers at the pre-college level, developing applicationsoriented and practical math and science instructional materials, helping colleges and universities upgrade lab apparatus including microcomputers, and continuing the process of research on how people learn science and mathematics (McCurdy, 1981b). The struggle for funding for science education continues to the present, but thanks to the efforts of individuals like Don McCurdy and his leadership at a critical time in the history of NSTA, many teachers have the opportunity to receive high quality training and curriculum materials to help their students learn key concepts in science. NSTA has evolved into one of the premier professional organizations in education thanks in part to the foresight and persistence of Don McCurdy at this critical time in the history of science education.

376   P. TILGNER

Conclusion Don McCurdy does not consider himself a pioneer in science education. Those who worked with him might agree because of his low-key gentle approach to leadership. Jim Rynearson, past president of the Nebraska Association of Teachers of Science, said in a recent conversation that Don worked hard to provide learning opportunities and support for the teacher in the schools where six to seven preps were common (Rynearson, interview, May 30, 2012). Gerry Saunders, one of Don’s graduate students in the late 1980s, adopted Don’s elementary methods teaching model emphasizing the dual nature of science, the content and how to teach it (personal correspondence, May 6, 2012). John Clementson and Mary Kalen Romjue, two other students Don mentored, valued Don’s help in introducing them to leaders in professional organizations. They also remarked on Don’s mentorship in learning how to teach science methods. Mary remarked, “Don’s program at UNL was not the run of the mill program that many universities had for their students. Instead, it was a leader in innovative ideas in science education” (personal correspondence, May 7, 2012). Gerald Skoog was not a graduate student of Don’s, but he worked with him first as a science teacher in Lincoln, and later, as president of NSTA. He writes, Don is a very conscientious person and has a strong record of involvement in the science education profession and community. Thus, I’m confident that his teaching was informed by experience, his scholarship and his continued preparation. Because my service as President of NSTA contributed much to my professional and personal growth during my three years as an officer and through the opportunities this service opened up for me then and later, I’m very appreciative of Don’s role in making these experiences possible. Overall, Don served as a mentor for me during my time in Lincoln, the years we shared a hotel room at NSTA conferences, and the time we’ve spent together the past four decades. I’m confident that other science educators that he worked with at UNL and through state and nationwide activities would make the same claim, i.e., Don is a mentor of note. I appreciate very much the contributions he has made to my life. (personal correspondence, June 4, 2012)

What Don taught so many of us is that it is important to take advantage of the opportunities for growth and leadership. He modeled that throughout his professional life and continues to do so by reconnecting with many of his former students and colleagues and keeping active in NSTA. I know that all of us who have worked with Don McCurdy consider him first and foremost a friend, and also a valued professional colleague.

Don McCurdy    377

References Brooks, D , Lewis, J., Lewis, K., & McCurdy, D. (1976). The teaching coach. A means of enhancing student-teaching assistant interactions during recitations. Journal of Chemical Education, 53(3), 186. Florida State University Office of Research. (2012). ISIS—Individualized Science Instruction System. Retrieved from http://www.research.fsu.edu/techtransfer/showcase/isis.html McCurdy, D. (1975). An analysis of qualities of self-directedness as related to selected characteristics of I.S.C.S. students. Science Education, 59(1), 5–12. McCurdy. D. (1981a). Report to the NSTA Board of Directors. Personal papers of Donald W. McCurdy. McCurdy, D. (1981b). Science education’s future: A case for government support. Science Teacher, 48, 48–50. McCurdy, D. (2011). Autobiography and family history. Unpublished memoir. McCurdy, D., & Brooks, D. (1979). Project TEACH: To teach teaching assistants. Journal of Research in Science Teaching, 8(4), 233–234. McCurdy, D., & Fisher, R. (1969). Physical science project: An individualized twoyear chemistry-physics course. The Science Teacher, 36, 60–63. McCurdy, D., & Fisher, R. (1971). A program to individualize instruction in chemistry and physics. School Science and Mathematics, 7(6), 508–512. Witter, L., & Sybouts, W. (1972). NUSTEP: A program for improving teacher education. Journal of Teacher Education, 23(3), 301–306. Wood, F. H., & McCurdy, D. W. (1974). An analysis of characteristics of self-directedness as related to success in an individualized continuous progress course in chemistry and physics. School Science and Mathematics, 74, 382–388.

378   P. TILGNER

Epilogue It was October of 1973 when the young man exited the subway at the 116th Street IRT station and walked north until he found and entered the ornate classical style building at Teachers College, Columbia University on 120th Street in Morningside Heights in Harlem. He climbed the stairs to the second floor large classroom in Thompson Hall. The young man was a first year graduate student in the Science Education doctoral program at New York University, and Jim Rutherford, his program director and boss on the NSF project for which the young man was a research assistant, had said they were to attend the yearly meeting of the Association for the Education of Teachers of Science (AETS), some gathering of science educators from the northeast United States. The young man had just finished his first four years as a science teacher in one of the toughest of New York City’s high schools and was more than a bit intimidated by this whole doctoral journey that he had undertaken, no less so by this looming attendance at some academic conference. Upon entering the room at AETS, the young man was glad to see not only another doctoral student from his program, but also Herb Schwartz, a science education professor from NYU. Herb suggested the young man get some pastries and tea and then began to introduce the young man to Stephen Winter from Tufts University, Harold Tannenbaum from Hunter College, Willard Jacobson, the ASTE host from Teacher’s College, Fletcher Watson from Harvard, and a group of other men who all seemed to know one another, yet greeted the young man genially. As best as he could remember it, the rest of the young man’s day was taken up by these professors and their graduate students, each presenting on some science education topic for about twenty minutes. They made presentations to

Going Back for Our Future, pages 379–382 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

379

380   Epilogue

the whole assemblage of about twenty-five educators, many of whom asked friendly questions about each presenter’s work. Needless to say, the young man sat there quietly, intimidated by this academic interchange. If he had known then about the history of the building in which he sat, where Kilpatrick had collaborated with the visiting John Dewey in the beginning of the century or with Ralph Powers as he trail-blazed the notion of science education at the university, the young man would have been even more intimidated. If he had known the prominence of the professors with whom he chatted at lunch, he might not have expressed any of the “teacher” opinions he interjected during their casual conversations. That young man turns out to be one of the editors of this book, Paul Jablon. As Paul reflects, “I find it humbling that over thirty-five years later after reading the chapters of this book how little I knew about the lifelong impact these people, some who became my colleagues, had on the lives of children in our country and the quality of their education in science.” In the three years following that AETS meeting, Paul not only worked for and learned from Jim Rutherford and his colleague Fred Geis, but as a project assistant he regularly conversed with Mary Budd Rowe, Audrey Champagne, Leo Klopfer, Fletcher Watson, Bob Stake, and Gary Bates as they came to consult for their project. Jim and Fred discreetly made Paul aware of their contributions to science education so that he had a context within which to speak with them about their ongoing work with inner city New York junior high schools. For example, Paul learned that Mary Budd Rowe had worked in Harlem schools before she went to Florida, where she was at that time traveling about in her motor home from school to school continuing on her “wait time” research that she had started back in 1964. It was important to them that Paul know what went on before, so that he could build upon the successes of the past, and not make the same mistakes that had occurred in assessment or curriculum endeavors or political maneuvers to get policy implemented in a city or state. Hence, when Paul finished at NYU, he was soon working with Joe Piel and Tom Liao, which ultimately led to his desire to become proficient in Science, Technology and Society (STS) based teaching. Which brings us back to the purpose of this book. Although it is a history book, history books typically are revisitations of events and dates of the past. This book, however, is not just a history book about the past. Although it memorializes contributors to our profession, its purpose is much greater and will hopefully have a greater impact upon our science education community of learners—now and in the future—than being just another book of history. It is why it is titled Going Back for our Future, Carrying Forward the Spirit of Pioneers in Science Education—since the past that is revealed in its pages certainly is a harbinger of what science education has become today and where it can go in the future.

Epilogue    381

As science teacher educators, we strive to help the science teachers we work with not to get lost in the minutia of science phenomena in the science classroom. We diligently work to help them see ways through which they can have their own students pull out the large, overarching concepts that help make sense of how the natural world works. Similarly, rather than just having a litany of what occurred in the lives of our colleagues who pioneered our profession, we are looking through the lens of history to unearth the large ideas, understandings, and trends that continue to define what we need to be successful in the future. Although this book is about preserving the history of the individuals who struggled and persevered, it is much more about what they and their contemporaries knew, how they knew it, how they accomplished what they did within the political constraints of their day, and how it informs our present quests. If we are to be successful in our ventures with science education, we need to build upon our predecessors’ legacies. We must, however, go beyond just reading about them here. We need to also go and read their work, speak about it at conferences, and learn what we can through communications with them directly (when possible) and with those who studied with them. To help us glean a better understanding and become the effective science educators we endeavor to be, we need to learn not only what worked for them, but also what didn’t work, and the source of their courage to persevere. Perhaps as important as learning the connections linking one pioneer to another is recognizing their legacy and building on it by applying the knowledge of the work done by each of them to ensure we do not reinvent the same ideas and publish about them using new names, over and over again. As we work and grow as a science education community to create new and effective methods of assessment, standards for science teaching, international relationships, curricular materials, legislative strategies, and better teacher preparation programs, we need to remember to stand on the shoulders of those who have gone before us, that we may benefit from what they have accomplished for us as a whole. As we edited this book and interchanged ideas with our authors, our realization was cemented that this book is not only for new people just entering our profession, but rather has much to teach and inspire each of us. We can now see how each of these pioneers has influenced our own work and values and how we interact within the science education community—in some cases directly, as we worked along side them, but mostly because they mentored someone who was of great influence in our professional lives. Understanding this legacy and the knowledge embedded within it makes us all better and more effective at what we do. The legacy aspect is an important one. Not all of us will have had the direct contact experiences that Paul Jablon related earlier in this epilogue. However, the reaches and influences of our science education predecessors

382   Epilogue

have touched most of us. Some of us have worked with mentors during our graduate (and undergraduate) programs and during early stages of our careers who were most capable in guiding and teaching us in concert with the ideas and philosophies set out by the pioneers. The names of those mentors may not make the international “pioneers” list, but their contributions to our educational and career development is nonetheless vitally important. Such mentors may have labored in one of the myriad smaller teacher education institutions and yet had significant impact on uncounted teachers and students, some of whom continued further with their educations and are now themselves conducting research and teaching at outstanding institutions. Equally likely is that some of these mentors who are not nationally or internationally “famous” may have briefly interacted with one of these pioneers and influenced the pioneer’s thinking radically. All this perpetuates the legacy. If one were to begin making lists of which pioneers connected to other pioneers, it would quickly become evident that there exists a definite network among them. We have attempted to provide some diagrammatic help through the connectivities diagrams included with each chapter and at the end of the book’s Introduction. This current volume includes eighteen pioneers, and the diagrams hopefully illustrate how each was somehow connected to the others. (And yes, we clearly recognize there are more pioneers to be included in subsequent volumes!) As noted in the Introduction, attempts to diagram these connections essentially ends up looking like spaghetti. Adding to the spaghetti, one should be able to expand backward from where he/ she is now, connecting to his/her mentor or colleagues, who then connects to his/her own mentor, and so on, logically connecting at some point with one or more of the pioneers included in the diagram. The network would become denser with nodes and connecting lines, eventually getting to the point where individual lines are almost indistinguishable from each other. None of us is really outside of it or disconnected from it. Our historical connections provide our anchor, functioning the way a stand of trees with connected root systems survive a hurricane while others that are free standing, even with deep roots, succumb to the perils of the storm. We are connected—and we are heirs of the legacy started nearly a century ago by those with perseverance, inner drive, and vision. As we move forward, building upon this past, we are building our own vision of what science education will become. And we are building it on the foundation laid by the pioneers.

About the Editors Jon E. Pedersen, PhD, is a professor of science education, associate dean for Research in the College of Education and Human Science, and director of Science Education for the Center for Mathematics, Science and Computer Education at the University of Nebraska-Lincoln. He received his BS in agriculture (Animal Science and Nutrition, 1982), MEd in administration, curriculum and instruction (Science Education, 1988), and PhD in administration, curriculum and instruction (Science Education, 1990) from the University of Nebraska-Lincoln. Pedersen began his teaching career as a secondary school science teacher and taught high school chemistry and physics. Pedersen is very active in several national organizations, including the National Association for Research in Science Teaching (NARST), Association for Science Teacher Education (past president, ASTE), and the National Science Teachers Association (NSTA). He is the author of over one hundred publications, most of which focus on science teaching and/ or the incorporation of social issues into the extant curriculum. He has also published nine books and two teacher manuals. During his tenure in higher education at the University of Arkansas in Fayetteville, Arkansas; East Carolina University in Greenville, North Carolina; The University of Oklahoma; and The University of Nebraska-Lincoln, Pedersen has been primary investigator and co-primary investigator of numerous grants and supported projects on science curricula development, science in-service education, middle level education, and international education totaling well over nine million dollars. Over the years, he has worked in more than a dozen different countries around the world.

Going Back for Our Future, pages 383–385 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

383

384   About the Editors

Kevin D. Finson, PhD, is a professor of science education and co-director of the Bradley University Center for STEM Education. He received his degrees from Kansas State University, including a BS in Science Education (1975), M. in Curriculum and Instruction with a focus on science (1980), and PhD in Science Education (1985). He began his career teaching earth and physical sciences at the high school level and then moved into teaching earth, life, and physical sciences at middle school. His college teaching has included earth science and geology, physical science, energy curriculum, teaching pedagogy and methods, and instructional theory. Among the professional organizations of which he is a member, Finson is very active in the Association for Science Teacher Education (ASTE). Other organizations in which he is active include the National Science Teachers Association (NSTA) and the National Earth Science Teachers Association (NESTA). He has authored over seventy journal articles, one book, several book chapters, and co-edited several books. A major focus of his work has been on students’ perceptions of scientists (such as Draw-A-Scientist-Test research) and more recently on models of conceptions of scientists in learners’ schema. He has also been active in publishing and work related to making science more accessible to students having special needs. Extramurally funded projects for which he has been a co-director include those on science and special needs and development of science potential in middle school students. Barbara S. Spector, PhD, a fellow of the American Association for the Advancement of Science and fellow of the National Institute for Science Education is a professor of Science Education and director of the Informal Science Institutions Environmental Education Graduate Certificate program at the University of South Florida. She received her degrees from Syracuse University: PhD in science education, a master’s degree in combined sciences, and a bachelor’s degree in biology. She holds a New York State Administration and Supervision certificate and taught high school biology for more than a decade. She has been awarded 67 grants totaling eight million dollars through which she initiated and maintained partnerships with professors in a variety of colleges and universities, with informal education agencies, and with business and industry, leading to the development of 31 courses. Her research focuses on how change occurs in education. She developed and directed three major centers at USF: the STS Center, Project 2061 Research and Development Center, and the Center for Ocean Science Education Excellence-Florida (COSEE-FL). She has been elected to national positions on the boards of directors of the key science education associations, has been active in policy making nationally and at the state level, has served on various advisory boards including for NASA and NSF, and been a consultant to a variety of state boards of education across the U.S. She has been honored for her pioneering work introducing the use of

About the Editors    385

social issues in teaching science through the interaction of science, technology, and society, and for her long-term leadership in the reform movement in science education. Paul Jablon, PhD, is an associate professor in the Graduate School of Education at Lesley University in Cambridge, Massachusetts. He received his BS in biology from Manhattan College in 1970 and a PhD in science education from New York University in 1989. Jablon taught as a science and interdisciplinary teacher for 19 years in the New York City public schools. He and his colleagues created critical small high schools that were so successful for students previously disengaged with science, and school in general, that they are still being replicated across the country through grant funded consortiums. During this time he was also president of the New York Biology Teachers Association and worked with the New York Junior Academy of Science. For over two decades, while directing science education and interdisciplinary programs at Brooklyn College CUNY, the University of Massachusetts at Lowell, and Lesley University, he has created innovative and effective pre-service and in-service programs that have positively affected the practice of thousands of teachers in urban areas. As primary investigator and co-primary investigator of numerous multi-million dollar grants, he has had a direct impact upon countless students in many states and urban areas throughout the northeast. He has authored numerous journal publications, book chapters, a complete curriculum on air quality monitoring (NESCAUM), and five other teacher guides in science and interdisciplinary project-based learning, and he has produced and directed three educational videos about urban science teacher training and project-based learning. He has taken an active role in national organizations such as the National Association for Science, Technology, and Society (NASTS) and the Association for Science Teacher Education (ASTE) and continues to run weeklong science and interdisciplinary staff development workshops in urban areas throughout the country for teachers in critical small high schools.

This page intentionally left blank.

About the Contributors Hanna J. Arzi started as a chemist and chemistry teacher, and her career has been an interplay between practice and research with ongoing interest in long-term learning and sustainable change. Her BSc and MSc degrees in chemistry and biochemistry are from the Hebrew University of Jerusalem, and her PhD in science education is from the Weizmann Institute of Science. Following post-doctoral research at Cornell University and Monash University, she was the establishing director of a prototype regional center for science education in Tel Aviv (Hebrew acronym—HEMDA). Currently she is an independent scholar, researching, consulting, and teaching. Dr. Abby B. Bergman currently serves as an independent consultant in strategic educational planning and programming. He was previously the Regional Science Coordinator for Putnam/Northern Westchester BOCES, a consortium of school districts in New York State, and prior to that he served as an elementary school principal for over 30 years and as a university instructor. He earned a BA degree from Hunter College of the City University of New York, and an MA, EdM, and EdD from Teachers College, Columbia University. Julie A. Bianchini is associate professor in science education at the University of California, Santa Barbara (UCSB). She received both her undergraduate degree in biological sciences and her PhD in curriculum and teacher education from Stanford University. Mary Budd Rowe was her dissertation adviser. Her research investigates pre-service, beginning, and experienced science teachers’ efforts to learn to teach science in equitable and effective ways. Bianchini serves as the faculty director for UCSB’s Science and MathGoing Back for Our Future, pages 387–390 Copyright © 2013 by Information Age Publishing All rights of reproduction in any form reserved.

387

388   About the Contributors

ematics Initiative, a UC-wide effort to recruit more science and mathematics undergraduates into teaching. She is also section co-editor of Science Teacher Education for Science Education. Dr. Rodney L. Doran has retired from his position as professor of science education at the University at Buffalo, where he taught for 33 years. Prior to that, he taught high school science in Minnesota. His BS degree was earned at the University of Minnesota in 1961, his MST from Cornell University in 1966, and his PhD from the University of Wisconsin in 1969. Rod worked closely with Willard Jacobson in the U.S. participation in the Second IEA Science Study. Robert G. Fuller (1935–2012), late emeritus professor at the University of Nebraska–Lincoln (UNL), was a physicist and physics education researcher. He worked tirelessly to make widely available any educational innovations and resources that could help improve student learning and teacher effectiveness in science, especially physics. Fuller’s interest in education was triggered when he sought to understand the difficulties pre-med students faced in their physics classes. He became a pioneer in multimedia educational materials, creating for example the first interactive videodisc for physics instruction (The Puzzle of the Tacoma Narrows Bridge Collapse), a Skylab film series, and Guilty or Innocent: You Be a Car Crash Expert, a prize-winning computer game for teaching mechanics. Fuller began collaborating with Karplus in 1973 to offer workshops on physics teaching and the development of reasoning. At UNL, he took the idea beyond physics and directed a Piagetian-based multidisciplinary program for freshmen called Accent on Developing Abstract Processes of Thought (ADAPT) from 1975 to 1997. To make similar strategies available to today’s college faculty, he led the preparation of College Teaching and the Development of Reasoning (2009). In 1999, Fuller embarked on a project to republish in one carefully organized volume all of the education-related papers of Robert Karplus, along with context-setting essays by several Karplus colleagues. In preparing Love of Discovery: Science Education—the Second Career of Robert Karplus (2002), Fuller became the world’s foremost expert on Karplus’ contributions to science education and the ideal lead author for this chapter. Among numerous honors, Fuller was awarded the Robert A. Millikan Medal of the American Association for Physics Teachers in 1992, for his outstanding contributions to physics education. He received his BS in physics from the Missouri School of Mines and Metallurgy and his master’s and PhD degrees in physics from the University of Illinois. Early in his career he taught high school physics in Burma and was a researcher at the U.S. Naval Research Laboratory.

About the Contributors    389

Dr. Beverly Karplus Hartline is vice chancellor for Research and Graduate Studies at Montana Tech, and she is the first of Robert and Elizabeth Karplus’ seven children. Previously she was Associate Provost for Research and Dean of Graduate Studies at the University of the District of Columbia. She has held teaching, research, and leadership positions in universities, in Department of Energy national laboratories, and at NASA, and she has worked in policy at the White House Office of Science and Technology Policy. She has been interested in science and science education since childhood, and she has been very active throughout her career in outreach to pre-college students and teachers—especially focused on engaging more girls and underrepresented minorities in science, engineering, and mathematics. Her bachelor’s degree is in chemistry and physics from Reed College, and her PhD is in geophysics from the University of Washington.  Prof. (Emeritus) Avi Hofstein was the head of the chemistry education group in the Department of Science Teaching at the Weizmann Institute of Science (Israel). He was also head of this department. His research focuses on all facets of the curricular process in chemistry—namely development, implementation, research, and evaluation. He has conducted research in learning environment, affective issues, learning difficulties and misconceptions in science learning, professional development, and laboratory work. In recent years, he has been involved in the development of leadership amongst chemistry teachers in Israel in order to promote reform in the way chemistry is taught in high schools. He has published more than 100 papers in refereed journals and more than 20 chapters in various books. Nicole I. Holthuis was honored to have Mary Budd Rowe on her dissertation committee for two years until Rowe’s death in 1996. Holthuis conducts classroom research and evaluation work in science education, primarily at Stanford University. Her work focuses on improving students’ educational experiences and outcomes by exploring the intersection of student and teacher talk and the nature of scientific knowledge. Holthuis has a PhD in curriculum and teacher education and an MA in design and evaluation of educational programs, both from Stanford University. She has a BA in biology from UC Davis and a California Secondary School Science Credential. Dr. Catherine Lange teaches a variety of graduate and undergraduate science education and science courses in the Department of Earth Sciences and Science Education at Buffalo State College. Her research interests include the study of twentieth century scientists, inventors and science educators. She is the curator of the Donald L. Birdd Historic Science Education Curriculum Library, which includes significant holdings of the alphabet curricula of the 1960s.

390   About the Contributors

Susan Mundry is a Deputy Director of Learning Innovations and Science, Technology, Engineering and Mathematics (STEM) at WestEd. She serves as a Senior Researcher for the Regional Education Laboratory-Northeast & Islands (REL-NEI) and lead facilitator for the Research Alliance on Educator Effectiveness. She also staffs the National Center for Cognition and Mathematics Instruction and designs the professional development program. Prior to this, she was co-Principal Investigator for two NSF projects on teacher development: Curriculum Topic Study and Building Systems for Quality Professional Development. Mundry is an author of several books, including Designing Effective Professional Development for Teachers of Science and Mathematics , The Leader’s Guide to Science Curriculum Topic Study , and Leading Every Day: 124 Action for Effective Leadership. Dr. Peggy Tilgner is a professor of practice at the University of NebraskaLincoln. She has been actively involved in science teacher education for more than 20 years. She was a post-baccalaureate student, doctoral student, and professional colleague of Don McCurdy. Anat Zohar is a professor of education at the Hebrew University, Jerusalem, Israel. Pinchas Tamir was her thesis advisor. Between 2006–2009 she worked for the Israeli Ministry of Education as Director of Pedagogy. In this role she led a national program implementing higher order thinking and teaching for understanding throughout the school system. Her academic research fields are the development of students’ higher order thinking, metacognition in science learning, inquiry learning, teachers’ professional development, gender and science learning, and bridging the gap between policy and practice in the area of teaching higher order thinking.