An Interview with J. Dudley Herron - Journal of Chemical Education

Crystal Growth & Design, Energy Fuels, Environ. .... life from the time that he was a college student until he arrived at Purdue. ... Chemical Educati...
0 downloads 0 Views 118KB Size
Chemistry for Everyone

An Interview with J. Dudley Herron Liberato Cardellini Dipartimento di Scienze dei Materiali e della Terra, Università degli Studi di Ancona, Via Brecce Bianche 60131 Ancona, Italy; [email protected]

The Interview What were the main factors that made you become a teacher? I had no idea what I wanted to be when I left home for college. My high school adviser suggested that I become a chemist or chemical engineer, but I had no idea what either occupation entailed. My initial college advisor was a chemist. He suggested that I should major in chemistry because, according to him, it would be easier to switch from chemistry to chemical engineering than the other way about. Teaching had never entered my mind, nor would it for at least two more years. The $800 I had saved for college ran out when I paid tuition for my second semester. I tried, unsuccessfully, to find a job and I finally got down on my knees and turned the matter over to God. The next day a friend suggested that I try Eastern State (psychiatric) Hospital, where he had recently found employment. The half-time position that I secured as an occupational therapy aide paid $60/month plus room, board, and laundry. It got me through school and taught me a great deal about people. (I often say that I got my degree at the University of Kentucky, but I got my education at Eastern State Hospital.) By the beginning of my junior year I knew that I enjoyed working with people and I had difficulty seeing myself in a chemistry laboratory—the only work I could imagine getting with a bachelor’s degree in chemistry. I considered medical school, but by this time I had married a student nurse who swore that she would never be married to a doctor! Dr. Wagner, my chemistry advisor, asked if I had thought about teaching. I definitely had not! I couldn’t think of a single teacher whom I wished to emulate, and I said so. “Well, you don’t have to be like them,” he replied, and he suggested that

I enroll in an education course for students transferring into education late. Reluctantly, I agreed, and that did it. It was the best course, by far, that I had taken in college. I was challenged to think, to use the library, and to explore new ideas. Sputnik had just been launched, and there was great excitement in science education. I was excited about being a part of the impending change. How did you end up as Professor of Chemical Education at Purdue? Once I decided to teach, I had to modify my program of study to include additional courses in physics and biology. At that time Kentucky’s certification requirements called for a broad area of concentration leading to certification in all sciences. Instead of taking calculus and physical chemistry as I would have done as a chemistry major, I took introductory courses in physics, geology, and the life sciences. I graduated from University of Kentucky certified to teach every science and qualified to teach none! The first year I taught, students asked all sorts of good questions that I couldn’t answer, and it bothered me. By that time the National Science Foundation had begun supporting summer and academic year institutes for high school teachers. I applied for an academic year institute at the University of North Carolina and was accepted on condition that I take calculus before I arrived. With the help of my wife’s high school chemistry and physics teacher, I successfully completed the first semester of calculus by correspondence and took a second semester course my first semester at UNC–Chapel Hill. I was still poorly prepared for the partial differential equations encountered in my thermodynamics course. I could do the problems, but it took an inordinate amount of time that I felt could be put to better use. I went to the professor and asked for permission to do independent study on some of

About J. Dudley Herron Professor Herron’s career in chemical education spanned more than 30 years. He taught high school chemistry for four years before completing his Ph.D. in Science Education and going to Purdue University, where he held a joint appointment in the Departments of Chemistry and Education from 1965 to 1989. In 1989 he gave up his joint appointment to chair the Department of Curriculum and Instruction in Purdue’s newly formed School of Education. He left Purdue in January 1994 to chair the Department of Physical Sciences at Morehead State University in his native Kentucky. Herron retired from active teaching and research in June 1996, but in the fall of 1999 he accepted a one-year appointment as Distinguished Visiting Professor of Chemistry at the University of North Carolina–Wilmington. He is currently representing Morehead State University in Chengdu, China, where MSU is helping to establish an English language school. Herron edited High School Forum, a column for secondary school chemistry teachers that appeared in this Journal from 1975 to 1980. His most recent publications are The Chemistry Classroom. Formulas for Successful Teaching (American Chemical Society: Washington, DC, 1996) and Heath Chemistry, 3rd ed. (with D. V. Frank, J. L. Sarquis, M. Sarquis, C. L. Schrader, and D. A. Kukla; D. C. Heath: Lexington, MA, 1996).

JChemEd.chem.wisc.edu • Vol. 79 No. 1 January 2002 • Journal of Chemical Education

53

Chemistry for Everyone

the questions raised by my students. He agreed, and I became intimately familiar with the Journal of Chemical Education and a few other journals while finishing my master’s degree in science education. When I returned to teaching, I brought with me a bright physics teacher, Lee Allison. He, Barbara Tea, and I set out to build the best science program possible. The Chemical Bond Approach Project was beginning its second year of school trials, and I signed on. We all attended conferences on various new science curricula, and we agreed to participate in a national study on the use of special science and math teachers in grades 5 and 6. We arranged for special math and science courses for our elementary school teachers, and we sat in on the courses. We participated in the AAAS Traveling Science Library, and I read the books along with my students. Although I had not completed a degree in chemistry, the chemistry, physics, and mathematics that I took at University of Kentucky and UNC provided enough background for me to read most science books with understanding. During my Ph.D. program at Florida State University, I took additional courses in chemistry. Charlie Holcomb, a physical chemist and fellow graduate student in science education, and I took Michael Kasha’s quantum mechanics course together and ended up with the two highest grades in the class. Kasha agreed to be the chemist on my graduate committee, and it was undoubtedly his recommendation that tipped the scale when I was being considered for the position at Purdue. When asked by members of Purdue’s chemistry department if I was a chemist, Kasha replied, “No, but he could be if he wanted to be!” When I went to Purdue in 1965, all secondary education faculty held joint appointments in the Department of Education and an appropriate academic department. Until that time, Ralph Lefler, who held a joint appointment in physics and education, had taught a methods course for physics and chemistry teachers, and he had supervised student teachers in both fields. But Ralph’s background was in physics, and he had encouraged the Department of Chemistry to join with the Department of Education to hire someone with a chemistry background to oversee the preparation of high school chemistry teachers. I was confident that Purdue would not hire me. None of my degrees was in chemistry, and chemists were directing the search. Although I felt good about my education and experience in science education, I knew that chemists would be skeptical. My Ph.D. research was typical of educational research at the time. It was an evaluation of CHEM Study, CBA, and traditional high school chemistry instruction, and I was not particularly surprised when the first question following my presentation at the Purdue interview was, “You call that research?” “Call it a study, investigation, or whatever you like. I know that it isn’t anything like the research that chemists do. But it represents the kind of question that interests me, and it represents the kind of scholarship that I expect to do. If you wouldn’t be comfortable having someone in this department doing that kind of work, you had better not hire me.” This exchange and Michael Kasha’s recommendation evidently got me the job. After I left the room and faculty started discussing my candidacy, a senior faculty member moved that I be offered the job. His rationale: “Anybody who 54

can stand up for himself like that should do just fine around here!” Sitting in on lectures given by other new faculty, I learned that having one’s Ph.D. in chemistry did not guarantee unimpeachable knowledge of the subject. And since my degrees were in science education, it was not difficult for me to confess my ignorance and ask for help in sorting out things that I didn’t understand fully: this earned me respect among my chemistry colleagues. Do you think that your lack of formal chemical training was an advantage in thinking clearly about chemical education? Did this place you nearer to the student situation? I have never seen much advantage in being ignorant, and I am confident that lack of chemical training in no way made me a better teacher or chemical educator. However, the training that I received instead of additional courses in chemistry was extremely valuable. After taking the introductory physics sequence for science majors, I needed additional physics credits and had difficulty finding a suitable course. I was enrolled in an introductory physics course for liberal arts majors— something that would never happen to a chemistry major. That course provided a far better qualitative understanding of physics than the course for science majors. Similarly, after struggling through the standard calculus sequence, I sat in on an introductory course for teachers. Once again, the less quantitative course led to a deeper understanding of the basic principles than I had gained from my previous exposure. It is possible that it was the second exposure to the ideas that led to the deeper understanding, but I think that the more important factor was that the qualitative course afforded more opportunity to put ideas in context and to see principles illustrated through concrete examples rather than working mathematical problems ad nauseam. An anecdote from my early experience teaching a onesemester remedial course at Purdue reinforces the conviction expressed here. Since placement of students in a remedial course is subject to error, I carefully monitored students’ early performance in the course and encouraged students who understood with minimal effort to drop my remedial course and enroll in the normal introductory course. I was surprised when a bright young woman, who clearly understood the material I was presenting, begged me not to force her to transfer. “I managed to get A’s in high school chemistry,” she explained, “but it never made sense to me. I just memorized. In this course it’s all coming together to make sense. I enjoy the course. I’m not struggling to keep up, but I’m not wasting time here. If you make me transfer to [the course for science majors] I’ll work harder and pass the course, but I won’t understand as well.” Reluctantly, I allowed her to remain in the course, fearing that she would lose interest after it was too late to drop-add. But she didn’t lose interest, and by the end of the course I was convinced that she demonstrated more wisdom than I. In summary, the paucity of my chemical training contributed no insights into chemical education nor did it place me nearer to the situation of beginning students. But several experiences that I would not have had as a chemistry major led to insights about my learning that I believe pertain to others as well.

Journal of Chemical Education • Vol. 79 No. 1 January 2002 • JChemEd.chem.wisc.edu

Chemistry for Everyone

Did this influence the way you wrote Understanding Chemistry: A Preparatory Course (1)? You seemed to introduce fewer ideas than normal. Does this mean that you subscribe to the idea that learning less may, in fact, lead to learning more? My experience in good, qualitative science courses was one of many factors that shaped Understanding Chemistry, but you are correct that I deliberately set out to reduce the number of ideas presented in the book and to seek a deeper understanding of those ideas. More important, I think, was an attempt to focus on “deep structure” rather than surface features. A major emphasis was placed on proportional reasoning, which many college freshmen find difficult.1 The major influence in writing Understanding Chemistry came from Piaget’s work. Several studies had shown that many college students do not use scientific reasoning—what Piaget called formal operations—to make sense of the world. Since a large number of science concepts and principles rely on those reasoning patterns for their understanding, science doesn’t make sense. Such students are in an untenable position. They are presented with ideas that they are expected to comprehend, but in order to comprehend students must apply reasoning patterns that they have never developed. Given enough time and the right experience, all students can develop those reasoning patterns and comprehend the ideas, but the pace of typical science courses precludes that happening. The only hope that students have is to memorize formulas and definitions in order to pass the impending examination. For hard-working students, that strategy usually works. They are able to pass the course, often with a respectable grade, but the material still doesn’t make sense, and the material learned by rote is quickly forgotten. The rationale behind Understanding Chemistry was to focus on one aspect of scientific reasoning (proportions), to present a limited number of proportional relationships in a manner that might reveal the nature of proportions, and to provide enough practice with the ideas presented so that students became comfortable with the concept. Information from other research also influenced Understanding Chemistry. For example, there is evidence that when two closely related ideas are presented in proximity, the ideas are easily confused. However, after one of the ideas is understood, the second idea can be learned quickly by focusing on similarities and differences between the new idea and the one already implanted in our mind. In Understanding Chemistry concentration is presented in terms of molarity, but molality and normality are not mentioned. It is assumed that other concentration terms can be taught later by contrasting them with molarity. Although the rationale behind Understanding Chemistry is fundamentally sound, it is only true that learning less may lead to learning more. It doesn’t necessarily happen. Free will is a mixed blessing. Students insist on doing things their own way, and the most carefully designed curriculum materials require student cooperation if they are to be successful. Does the traditional “logical order” by which we teach chemistry conflict with the ideal psychological order? Many years ago David Ausubel argued that logical order does not necessarily correspond to psychological order (2), and I am confident that he is right. The problem is nicely illustrated by research on problem solving.

Anyone who has used the “think aloud” technique to observe the problem-solving strategies used when someone solves a novel problem knows that the process is punctuated by false starts, dead ends, silly errors, redundancy, and occasional insights that eventually lead to solution. Problem solving is a messy process. One idea leads to another in an order that seldom reflects the efficient algorithms presented in textbook solutions manuals. After we thoroughly understand a problem, we reorganize our original thinking and solution efforts to present the solution as efficiently as possible, but that isn’t the way we approached the problem when it was new to us. In a sense, knowledgeable people present what they know from this perspective of having solved a difficult problem and are now reorganizing it as an efficient algorithm. Having sorted out the various pieces of the puzzle and considered ways that they might fit together, the subject is presented “logically”. But to someone who is new to the subject, that logical presentation may not make sense because it often presumes understanding that does not exist. But what is the ideal psychological order? Ausubel has suggested that it starts with concrete, personal knowledge and extends out to the more abstract and less personal. Unfortunately, your personal knowledge isn’t the same as my personal knowledge, so it is impossible to suggest an order that is ideal, psychologically, for everyone. Still, research on student misconceptions suggests that the stumbling blocks to understanding are often the same for many students, and they are often the same stumbling blocks that occurred in the historical development of ideas. For example, Nussbaum has concluded that, in the development of particulate models of matter, it is the idea of nothing—a vacuum—that is difficult for students to grasp, and it was this same idea that caused the greatest difficulty historically (3). Staying with the idea of content, why are some concepts perceived by students to be more difficult than others? Is this intrinsic difficulty or is it a result of some psychological artifacts? Do your ideas of concept analysis help here? There is no single reason that some concepts are more difficult to an individual than others. Were that the case, everyone would have difficulty with precisely the same ideas and everyone would fully comprehend others. That isn’t the case. Whether it is due to the neural wiring with which we are born, the neural connections we construct at an early age, or some combination of the two, there is strong evidence of individual differences in aptitude. Musical concepts certainly come to my wife more quickly than they come to me, whereas spatial relationships that are self-evident to me are a puzzle to her. A complete answer to why this is so undoubtedly resides in a lifetime of experience and a few twists in our respective DNA. Still, there are concepts that are generally more difficult for just about everyone to understand, and concept analysis has helped me rationalize why that might be so. As a general rule we learn concepts like “dog”, “car”, or “test tube” by viewing examples and non-examples of the concept and identifying attributes that are common to examples but missing from the non-examples; dogs have fur and bark; cars have headlights and honk. But there are concepts for which it is impossible to show examples (atom and molecule, for example) and there are other concepts for which there are plenty of visible examples but whose critical attributes

JChemEd.chem.wisc.edu • Vol. 79 No. 1 January 2002 • Journal of Chemical Education

55

Chemistry for Everyone

are not perceptible (element and compound, for example). Concept analysis can reveal such characteristics of concepts that are likely to interfere with learning.

for example, are necessarily murky when the relationships between particles, mass, and amount of substance inherent in the mole concept are not understood.

Could you share with us your reflections on what your experiences in the last 20 years would lead you to do if you were going to rewrite your famous paper on Piaget ( 4)?

What is the “Third Level of Knowing” and how can it be achieved?

“Piaget for Chemists” struck a responsive chord because it provided a language to describe a condition that experienced teachers had observed many times in their students. It explained what they knew to be true. The basic premise of the paper is still true: A substantial number of students enter college without much facility with the reasoning patterns that Piaget called formal operations and others have described more generally as scientific reasoning. As a consequence, students have difficulty comprehending a variety of science concepts that depend on that reasoning for their meaning; for example, until a person understands the reasoning that leads someone to say that 70 parts out of 100 is equivalent to 105 parts out of 150, there is little chance that they can comprehend stoichiometry, solution concentration, density, and a host of other ideas that involve proportional relationships. The only thing that I would attempt to do if I were rewriting “Piaget for Chemists” is to emphasize a little more the process by which knowledge is constructed and reduce the focus on the stages of development that Piaget described. Too many people got the impression that the stages simply unfold as a consequence of maturation and that nothing can be done to alter that. I believe that much can be done to enhance intellectual development, but it takes time and persistent effort on the part of the learner. Much of your work concerns problem solving. Where and why do students go wrong here? All learning is, in effect, problem solving, and there are as many ways to go wrong in learning as there are learners— actually more, since each of us goes wrong in different ways at different times. Still, two general categories cover a majority of failures at problem solving: One of the difficulties originates with teachers rather than with students. We convey the notion that there is but one way to solve a problem and that the path to solution should be transparent. Consequently, students look at a problem, realize that they do not know how to solve it, and quickly give up. I see no fundamental difference in what must take place when teenagers face a new chemistry problem and what takes place when infants learn to fit wooden shapes into holes. The infant goes through a great deal of trial and error, trying to force round pegs into square holes, turning half moons the wrong way, and such; all the while gathering information about the pegs and the holes until the infant eventually coordinates perception and action to unerringly place the pegs into the proper slots on the first try. All true problem solving is similar, and it takes time to discover what information stored in our brains may be relevant and how the pieces fit together. The various metacognitive strategies and heuristics taught to enhance problem solving are just devices that help us sort through information and find relationships. The other large source of failure at problem solving— particularly of the type commonly found in chemistry—is poor conceptual understanding. Stoichiometric calculations, 56

Schooling is often concerned with getting right answers, with the result that students often learn algorithms for getting right answers with no appreciation of what the answers mean or why the algorithm works. Take percent, for example. Just about every adult knows how to calculate a percent by “dividing the little number by the big number and multiplying by 100.” But many who apply (and misapply) the algorithm have little understanding of what the resulting number implies. Being able to apply rules, computer fashion, to produce right answers that have little meaning represents a primitive level of knowledge. Being able to apply rules to produce answers that one understands represents a second, and decidedly higher, level of knowing. But there is a third, more powerful level. It occurs when one can not only apply rules and understand the result but can understand why the rule produces a sensible result. That understanding may take a variety of forms and reveal a number of relationships that could be expressed in similar rules. A person who understands percent at this “third level of knowing”, for example, should easily see the relationship between percent and parts per million or parts per billion and realize that any base—10, 12, or 144—could be used to define a ratio called “perdecca” or “perdozen” or “pergross” with the same properties as the common percent. Achieving this third level of knowing is largely a matter of striving for it, but it can never be achieved as long as teachers just focus on right answers! Far more attention must be given to why an answer is right, what the answer means, and how the logic used to arrive at this answer might be used to answer other questions that have yet to be asked. Referring to your paper on Rutherford (5), science often progresses through heuristic principles. Why do textbooks favor the “rhetoric of conclusions” approach so consistently? For publishers, the bottom line is the bottom line. They are in the business of selling, not teaching. Market research influences decisions, not pedagogical research. Furthermore, teachers and administrators make purchasing decisions; students do not. Those decisions are made on the basis of superficial reviews. Teachers do not adopt books that omit their favorite topics, so publishers insist that all teachers’ favorite topics receive some attention. The result is a tome that covers everything, often superficially. You have firsthand knowledge of several educational systems in other countries at high school level. From this point of view, how would you advise chemical education practitioners in the United States on ways of improving things in high schools, or is everything fine already? Everything is not fine, and it never will be, thank heavens! We should always assume that we can improve, but we should be clearer about our goals and we should be more diligent in seeking evidence that the changes we so readily adopt actually move us closer to those goals. Accountability is the current watchword, and it is a wonderful idea. The problem is that far too much time is wasted on the mechanics of checking

Journal of Chemical Education • Vol. 79 No. 1 January 2002 • JChemEd.chem.wisc.edu

Chemistry for Everyone

up on students, teachers, and schools and far too much attention is focused on test scores and the like. What we need to know is the extent to which students can make sensible statements about the physical world in which they live and how adept they are at using logical analysis to make sense of new phenomena. Using whatever devices a teacher finds effective with particular students in a given school, students must be made to believe that school is about learning and that learning is worthwhile as well as fun. It is not about getting grades and credits and degrees and lucrative jobs. Those things may be indicators of sorts, but they are not the purpose of school. School is about learning. The real question is how one can get students interested in learning—more correctly, interested in learning those things that adults deem worthwhile. Seduction, I think. The only success I had came when I detected some glimmer of interest in students and then found ways to channel that interest toward things that seemed important to me. That requires that teachers interact with students on an individual basis and that they be flexible enough to capitalize on “teachable moments.” One of my best experiences grew out of an argument in a high school physics class. A farm boy asked why hot water freezes faster than cold water, and others laughed at his preposterous suggestion. But he knew what he was talking about. He had placed hot water and cold water on the frozen ground in winter for the chickens to drink, and he knew that the hot water froze sooner than the cold. “No way!” his classmates insisted, so we had an honest debate that kept the entire class investigating for about six weeks. In the process students learned that the density of water varies with temperature as does its specific heat; that the energy required to vaporize a gram of water is much larger than the energy required to raise its temperature from freezing to boiling; that answering even simple questions requires us to make somewhat arbitrary decisions such as what we mean by “hot” or “freezing”; that successful experimentation depends on developing practical techniques—finding a way to observe water while it cools, cooling it fast enough to freeze before you get bored but slow enough that you can time the process accurately, for example. In another class students who had their own band were invited to bring their instruments and play a song. That led to a discussion of what makes the music “good” or “bad”. We looked at the wave pattern formed on an oscilloscope when the same note was played on different instruments and talked about the similarities and differences in the patterns. In order to do that we had to talk about the characteristics of waves, how they are generated, how they interact, and all of the other topics that have fascinated scientists for decades. There is absolutely no inherent value in the particular activities that I have described. They just happened to be activities that were of interest to particular students at a particular time. I was able to use them to teach because I was willing to deviate from a set syllabus and nobody insisted that I shouldn’t. How much importance do you attach to the place of laboratory work in the teaching of chemistry at all levels? As long as you understand laboratory work to be the kind of activity I have just described, I would say that it is abso-

lutely essential. Percy Bridgeman described science as “doing one’s damnedest with one’s mind, no holds barred” (6 ). That is what laboratory work should be about. It should be seeking answers to real questions, not arbitrary exercises. The best laboratory program that I know is the one developed for the Chemical Bond Approach Project under the direction of Tony Neidig. Each lab began with a prelab discussion, during which a question was posed and students suggested how it might be answered. The next day data were collected, and the following class period was devoted to a discussion of the data and their meaning. In my experience, students sometimes disagreed about the best procedure to use, and I encouraged proponents of each technique to follow their instincts and report their results. Postlab discussions often revealed flaws in a technique that were not obvious from the start, and what was lost in terms of unusable data was insignificant compared to what was learned about experimentation and the vagaries of the experimental method of answering questions. I often went on fishing expeditions during laboratory work. Looking over shoulders, I would call attention to some unexpected result and ask why the student thought it had occurred. Occasionally students took the bait and we would set up an impromptu experiment to see if we could find the answer. Results of these off-the-cuff experiments would be reported as part of the postlab discussion the following day. Quite often the answer to the question would remain unclear and we would pursue it through reading in the textbook or in the library. Because the question was rooted in the students’ own experience, the library research was undertaken with some enthusiasm, and students found it easier to understand what they were reading because they could relate it to their own experience in the lab. It is this ability to make connections between what one reads or what one hears and what one has experienced directly that makes laboratory work an essential part of learning chemistry. But if the laboratory work is nothing but a series of exercises that are divorced from other aspects of the course, it is rather useless. What observations do you have upon the quality of the preparation and in-service support of chemistry teachers? Are they born good teachers or can good teaching be taught? I recall discussing your second question with fellow teachers in 1958. I thought the answer was obvious then, and I still do. Why would anyone presume that teaching is different from other complex activities? Do we presume that great actors or singers or doctors or automobile mechanics are just born with their talent, full blown? I certainly don’t! They acquire their talent over time. A few may acquire their talent without benefit of instruction, but not many. They are taught. But there are still differences in aptitude, part of which is undoubtedly determined by genetics. So it is with teachers. Some inherit characteristics that predispose them to excellence, but they still benefit from instruction, as do all who wish to teach well. The primary problem with the preparation and in-service support of chemistry teachers is that there isn’t any. Chemistry courses are designed for chemists, chemical engineers, students in allied health fields, and occasionally for liberal arts majors, but not for chemistry teachers. Education courses, both

JChemEd.chem.wisc.edu • Vol. 79 No. 1 January 2002 • Journal of Chemical Education

57

Chemistry for Everyone

preservice and in-service, are designed for science teachers, not chemistry teachers. There are never enough chemistry teaching majors at any one school to justify courses that are designed specifically to meet their needs. The resulting education isn’t really bad. The chemistry courses provide the content knowledge that teachers need, and the education courses provide the (general) pedagogical knowledge that they need. What is missing is the pedagogical content knowledge that is most important for effective teaching. By pedagogical content knowledge I mean knowledge about how to teach particular content. It addresses questions like these: What is the best sequence for topics in chemistry? How does one decide? What chemical system(s) is (are) most useful in introducing equilibrium (or enthalpy or kinetics or …)? Are there particular demonstrations that are more effective than others? If so, what makes them effective? How should you do demonstrations to make them most effective? When is a demonstration more useful than an experiment and vice versa? How do you organize chemicals in the stockroom? What solutions should you prepare well in advance and which must be prepared just before use? What safety precautions must one take? What legal obligations do you have for students’ safety? What is your liability exposure? What does it cost to run a chemistry lab? How can you keep costs down? What are the advantages and disadvantages of semimicro experiments? Can simulations replace experiments without sacrificing learning? And on and on! As far as university teachers are concerned, there is a feeling abroad that they should be trained in the same way as the present generation were trained with a basis in classical chemical research. The teaching is then picked up by experience. What are your views on the acceptability of research in chemical education? All of us can and do pick up a great deal of what we need in our jobs through experience. That isn’t going to change because we don’t want to stay in school all of our lives. It makes sense to focus schooling on those things that we will do most and leave to experience the peripheral tasks that we will occasionally do. If one accepts this rationale, then chemists who plan careers in industry, and possibly those destined for research universities, are served well by traditional chemical research training. But the large number of Ph.D. chemists who end up in teaching colleges are poorly served by the traditional training, and their teaching shows it. The Ph.D. program for those chemists should include about 12 semester hours of course work related to teaching, and they should do research in both chemistry and chemical education. (I would recommend a master’s thesis in chemistry and a Ph.D. thesis in chemical education, but that probably just reflects my bias.) As a prolific contributor to journals, can you give us any advice about presenting chemical education papers in an acceptable way? The goal should be to inform, not to impress the reader or the promotion committee. Don’t publish until you have something to say. (Far too many articles are published because of promotion considerations.) All good writing is reader based rather than writer based. Think about your audience. What do they know? What must you tell them so that they can understand what you think you have learned? Tell them what 58

conclusions you have drawn from the data you present, but give them enough information to allow them to form their own conclusions as well. Don’t be as wordy as I am! In the latter part of your career you have become an administrator. How hard is it to reconcile ideal educational practice with the “limitations” of financial and administrative restrictions? What I found difficult to reconcile was the desires of all the faculty. Everyone has a pet project that requires money and space, and there is never enough to go around. Publicly supported institutions have an obligation to spend tax monies wisely, and “wisely” will be defined by the public served. When I moved to Indiana from Kentucky, I was struck by how posh the public schools were, and when I lived in Israel I was struck by the austerity of the school buildings. But the teachers I knew in Israel were bright, dedicated, and comparatively well paid. I think that the Israelis were better stewards of public funds. How can research in chemical education be more effectively transformed into improved teaching and learning? Perhaps if we quit using textbooks, it would help. Then we would not feel obliged to throw out our material every three or four years and reorganize. We could begin to systematically build a course based on materials that have been demonstrated to be effective in teaching a particular concept or principle, and we could continue using those materials until someone demonstrates that other materials do the job better. It will also require that new faculty members taking over a course be informed about why the course is designed as it is, and they must not be allowed to change materials until they can demonstrate that the changes they wish to make lead to better learning. A final and very personal question: It seems that your dedication to your students and the whole tenor of your work is driven by some inner force. How much has your religious outlook affected the nature and direction of your career? My religious faith is the most important thing in my life, and it influences everything that I do. I am a Christian, and I have consciously tried to follow the example set by Christ as I understand what is reported of that life in the New Testament. Put simply, I try to leave the world a little better than I found it. I try to respect all people and believe that every person has value—even those whom I dislike. As a teacher I have tried to help each student be the best person he or she can be; I have seen my job as helping them become what they wish to be rather than coerce them into becoming what I think they should be. When I was a sophomore in high school, I felt “called” to Christian service, but I wasn’t at all sure what that meant. The counselor with whom I spoke suggested that I should set out to do the greatest thing that I could do for God, and if God didn’t stop me, assume that I was on the right track. It seemed like reasonable advice, and I have tried to follow it. Most of what I have done has been at someone else’s suggestion. People have offered me positions of responsibility, and when I felt that I could do the job and the job was worthwhile, I accepted. At times I have accepted too much and had to pull back in order to do anything well. In general, others have been more impressed with my accomplishments than I

Journal of Chemical Education • Vol. 79 No. 1 January 2002 • JChemEd.chem.wisc.edu

Chemistry for Everyone

have been, but I am quite satisfied with my life. In April of 1999 the ACS Division of Chemical Education honored me with a special symposium. The speakers were former students and colleagues who talked about various aspects of my career. I had dreaded the affair for almost a year, but I thoroughly enjoyed it. The speakers convinced me that my years of hard work had produced good results, and that’s nice!

University of Glasgow, Scotland; Mansoor Niaz of the Universidad de Oriente, Cumaná, Venezuela; and Mary Virginia Orna of the College of New Rochelle, New Rochelle, NY, for the advice and suggestions they gave me for improving the questions for this interview and the assistance during the interview. Note

Conclusion And you deserve this honor because your research and your many writings, both papers and books, have made a difference in the way many people around the world teach chemistry. You have had an influence that few other instructors and research scholars can claim. You leave a rich heritage to all of us, but particularly to your former students, some of whom have become distinguished scholars in their own right. Your scholarly contributions are not only relevant to the chemistry education community, but also to practitioners of chemistry in many other fields. On behalf of the chemistry education community, I wish to thank you for sharing your insights, the fruit of a life dedicated to chemistry teaching. Acknowledgments I would like to thank George M. Bodner of Purdue University, West Lafayette, IN; Alex H. Johnstone of the

1. Arnold Arons discusses proportional reasoning and strategies for developing it in Chapter 1 of A Guide to Introductory Physics Teaching (Wiley: New York, 1990) and his later book Teaching Introductory Physics (Wiley: New York, 1997).

Literature Cited 1. Herron, J. D. Understanding Chemistry: A Preparatory Course; Random House: New York, 1981. 2. Ausubel, D. The Psychology of Meaningful Verbal Learning; Grune and Stratton: New York, 1963. 3. Nussbaum, J. In Teaching Science for Understanding: A Human Constructivist View; Mintzes, J. J.; Wandersee, J. H.; Novak, J. D., Eds.; Academic: San Diego, 1998; pp 165–194. 4. Herron, J. D. J. Chem. Educ. 1975, 52, 146-150. 5. Herron, J. D. J. Chem. Educ. 1977, 54, 499. 6. Bridgeman P. W. Yale Review 1945, 34, 450.

JChemEd.chem.wisc.edu • Vol. 79 No. 1 January 2002 • Journal of Chemical Education

59