Archived Newsletters

Comments from the Chair: A Hundred-Year-Old Bad Habit

Rush Holt

You may know the old joke about the teacher who says with exasperation, "I've taught this topic every year for 20 years and the class still does not understand it!" Many teachers will admit, at least privately, that the joke is a little too close for comfort. Although in recent years I have been teaching only an occasional class here and there, I remember all too well occasions when I as a teacher was frustrated by the students' apparent stubborn refusal to see a particular point as clearly as I did, even when I patiently repeated my explanation. Nothing is so hard to break as a bad habit, in this case the teacher's. Sometimes to gain a new perspective one has to climb steep terrain.

You might suspect that my remarks are not meant solely for teachers. The constituents of the APS Forum on Education are varied, coming as we do from industrial, academic, research, and policy communities. Also, the spread of Forum members across the other topical divisions of the APS is remarkably uniform. So, it is clear that the physicists from all perspectives look to the Forum for ideas and action with regard to education. Likewise, the Forum leadership looks to the members in the various divisions to represent the interests of the Forum throughout the practice of physics. This issue of the newsletter focuses on some aspects of undergraduate physics, and the topic is important for even those who do not teach undergraduates. Physics teaching is a reflection of how we view our discipline, and it determines how others view it.

As the APS prepares to celebrate the centennial of our society, we should consider something else that is more than a century old, something not so grand and glorious. I call it the theft of physics. In 1892 the standard setters for college admission met at the University of Chicago and prescribed that high school students should take one year of biology followed by one year of chemistry followed by one year of physics. We practicing physicists have succeeded in that system, but we should question a system where physics is the ultimate course, taught only after the casual and so-called less talented students have been weeded out. We thus exclude 80% of all students from the pleasures and benefits of our science. We proudly continue this tradition in college. We are left with a situation where a lawyer I know who can write a flawless pension plan says he never took physics courses because physics was for the brainy students. We are left with elementary school teachers who should be teaching physical science feeling inadequate and guilty. And we are left with Members of Congress who for the most part have little idea why they are funding (or not funding) physics research.

We have presented physics as declarations from on high, rarely as an activity that a student might want to dabble in after class. Some researchers have looked at our century-long experiment in teaching, and they have concluded that our traditional approach was not working as we thought it was. It comes as a revelation to beginning teachers that what students learn is often not what the teacher teaches. Even more disturbing than students' apparent slowness to understand some concepts is their quickness to learn things that are false. Students' misconceptions often mystify teachers who are certain that they have carefully covered the relevant subject matter. It now appears that those misconceptions are partly a result of our over-arching misconceptions about how students learn. Science is by nature non-dogmatic. Yet, why is it that scientists are so often seen as dogmatic? Why is it that elementary teachers emerge from college with no background in physics, or worse, courses that leave them with the sense that physics is to be done only by specially trained experts? Science Teaching Reconsidered, A Handbook, prepared by the NRC, presents a number of useful teaching suggestions, some obvious, others more subtle or clever. Most important, the handbook effectively challenges the misconception, prevalent among scientists for at least a century, that knowledge of a subject implies the ability to teach it. The handbook shows the error of ignoring what is known about how students learn. There is now significant research about how people learn. The lessons of this research are no less relevant for teaching graduate electrodynamics than for teaching introductory kinematics. Traditional teaching methods based on lectures, assigned readings, problem sets, and formulaic labs are less effective than inquiry-based, cooperative approaches that engage students actively. The changes necessary for truly effective teaching are fundamental and get to the heart of how we view our field and ourselves. I believe that developing more a "student-centered" learning tradition can only help physics as a discipline, but the changes may be difficult for us to accept and to implement.

Changes are underway, and I expect that the pace of change will accelerate greatly. The Board on Physics and Astronomy of the National Research Council is "reconsidering" teaching as it prepares its decadal review of physics. And numerous individuals and some academic departments of physics are also re-examining physics teaching. A few are even beginning to recognize the professional validity of physics education studies. Maybe the students will finally begin to get it -- but not before we do.

Tripling the Number of Physics Majors at a Research University

by Joe Pifer

At Rutgers we have revised our undergraduate program to encompass non-traditional physics majors with a greater breadth of career interests and over the past nine years have seen the number of physics majors triple to 45 graduates in 1997 with another twelve minoring in physics. Our approach to making the physics major attractive has been to introduce a variety of options with widely different math and physics requirements:

  • Professional Option a standard no-holds-barred sequence of courses intended to prepare students for graduate school.
  • Engineering Physics Option a five-year program leading to dual degrees in physics (BA) and engineering.
  • General Option a liberal arts (BA) degree designed for pre-med, pre-law, etc. students, high school teachers, and students with no particular career goal in mind but who are attracted to a science major that gives them the flexibility to try out a variety of different fields or possibly double major.
  • Applied Option a BS degree that emphasizes a breadth of technical knowledge rather than narrow specialization and is designed for students who do not want to pursue a Ph.D., but who are interested in a job in some applied area or possibly a Masters degree.

We also have a minor in physics and a newly introduced minor in astronomy centered on our 0.5 m instructional telescope.

The Bachelor of Science Carrot
One of the ways we differentiate between options is by awarding the professional and applied majors a Bachelor of Science degree and the general and five-year engineering majors a Bachelor of Arts degree. The university defines the BA as a comprehensive degree implying knowledge of a broad range of disciplines with an in-depth of study in one or more disciplines; while the BS is a specialist degree implying intensive technical knowledge of a particular discipline and cognate subjects. In reality BS and BA students are required to take the same distribution of liberal arts courses and the only difference lies in the number of technical credits required, and hence in the number of free electives. It's up to the BA student to make a wise choice of electives to obtain greater breadth than possible with the BS.

Nationally, there is no consensus as to which type of education is preferable for a scientific career. But students interested in a physics career, rightly or wrongly, have no doubt that the BS degree is much preferable. We use this perception to counteract the temptation for students who are struggling with the very challenging professional or applied options to change to the less rigidly prescribed general option. At the same time when advising BA students we attempt to counteract the perception of the BA as an inferior degree by emphasizing the long term importance of a broad education. In particular, we strongly encourage general majors to pursue a second major or several minors as well. In the previous three years 60% of the physics BA students completed a second major in addition to physics, as compared to 29 % of the BS students.

Advising, the Critical Component
Introducing a flexible curriculum will not guarantee an influx of physics majors; good advising is by far the most important factor in increasing the number of majors. Students who are unsure about their career goals are easily discouraged by indifferent, inaccurate, or hard-to-get advising. The problem is compounded if students can choose among numerous options. Our advising system uses carefully written documentation of the requirements, an e-mail alias reaching all majors and touch-tone registration. It is essential that the advisor be knowledgeable about all academic requirements so that students are not sent away frustrated or misinformed. For this reason we have moved from an advising committee to centralized advising with the undergraduate director doing most of the advising, except for professional option students who are assigned a mentor.

We find the most effective way to recruit physics majors and provide meaningful advice is at the first meeting with a student interested in majoring in physics. Such students frequently drop in without an appointment, but even if the student's interest seems tenuous, we find it essential, if at all possible, to immediately take the time - 15 minutes or more - to explore the student's interests, outline the career opportunities, help him or her choose the appropriate option, and to give the appropriate written material. This is the ideal time to discuss the advantages of getting a broad education and perhaps doing a double major. Many students have only a vague or very narrow understanding of what a career in physics might entail and such intensive one-on-one advising helps them solidify their thinking and see the advantages of studying physics. For some students this first meeting is the last time they will seek individual advice except in their senior year when they are job hunting or looking into graduate or professional schools. Others keep closer contact by frequently dropping in for a "quick question" or by e-mail.

But What About the Teaching Load?

There are three ways to broaden the curriculum. The first, allowing students in other options to pick a subset of the courses intended for the professional, costs nothing in teaching load and helps populate courses which otherwise might not have sufficient enrollment to be offered. The second, allowing or requiring courses in other disciplines, again costs nothing. The third is to offer courses specially designed for the new option, which, of course, impacts the teaching load. In order to offer the general and applied options we offer five semesters of "special" junior-level courses and expect to introduce one or two more as we develop the applied major and attract more students.

Advanced General Physics: This course addresses our basic assumption that at a minimum, a general physics major must have a thorough and deep, mathematically based understanding of the concepts of introductory physics. This is frequently not the case even for students who have done well in an introductory sequence. In addition, the general option students come from a variety of academic backgrounds, so it is difficult to design a lecture course that meets all of their needs. Our solution is actually a leftover from a pedagogical fad that swept through physics 30 years ago self-paced learning the Keller Plan.

Advanced General Physics is a two-semester, three-credit course required of all general and applied option students. The material is broken into modules and students are given a study guide which includes textbook readings and homework. There are no lectures, although video taped lectures are available for a few of the harder modules. Students study on their own with the possible assistance of tutors (undergraduate or graduate, supervised by a faculty member) who are available 6 hours a week. When the student feels ready, he or she asks the tutor for a test from the ten or so kept on file for each module. The test is graded pass/fail and the student can retake the test for a given module as many times as necessary. The course grade is determined by the number of modules passed. The possibility for cheating is reduced by treating the grading as a learning process where the student explains and justifies his or her solutions. There are both advantages and disadvantages to the course:

  • Students can spend as much time as needed on areas where their preparation is weak
  • Students can defer studying on weeks when other courses have exams, which reduces the overall intensity of the curriculum
  • Students whose study skills are weak or who procrastinate do poorly in such an unstructured environment
  • The course provides an effective way to utilize faculty who are weak lecturers.

The Physics of Modern Devices: This course is designed to help students make the connection between the physics they learn and their everyday life. The prerequisites are a year of calculus and of introductory physics. The syllabus is flexible, depending on the instructor's and students' interests, but typically includes a discussion of transistors, solar cells, nuclear reactors, refrigerators, television, lasers, compact disk players etc. One difficulty in teaching the course is a lack of a suitable textbook. Ideally the course should include a lab. This course is becoming increasing popular, with an enrollment last year of 47. One surprising element is that a number of engineers take it as a technical elective, because their engineering courses are too theoretical.

The Physics of Sound: This course builds on student interest in acoustical systems and includes a weekly 80-minute lab. Again the prerequisites are a year of calculus and of introductory physics. Our intent is to position the course between "Acoustics for Music" and a full-blown technical acoustics course. We have not found a completely suitable textbook and the course tends to stray to one direction or the other depending on the instructor.

Experimental Applied Physics: The center piece of the applied option will be a required applied physics lab that we are developing with the assistance of a NSF-ILI grant and grants from several equipment companies. The lab currently has a strong emphasis on lasers and optics and is biased toward giving students experience with the apparatus and techniques they are likely to encounter in an industrial laboratory. At present, in order to offer the course despite having only a few applied majors, we schedule the course at the same time as the modern lab for professional majors, but have the students do a different selection of experiments.

The Future
Despite our large increase in the number of physics majors, we are not content with our program and are working on three areas with the goal of once again doubling again the number majors we graduate:

Evaluation of our program: We need to keep better track of majors as their careers develop in order to get feedback on our program to improve it.

Development of the Applied Option: It's very difficult to start a new option with no students and no resources. We built the applied option around existing courses so that we needed essentially no new resources. Now that we have some students (13 applied option majors in 1997), we need to redesign the option for the bottom up, asking what should these students learn in order to be ready immediately for a applied physics career. This will involve new, specially designed courses, and contact with industry to see what they expect from our students. We also hope to establish a summer industrial internship program open to any of our majors. Our long term goal is to make the applied option an attractive and equally respected alternative to the professional option.

Training more High School Teachers. A few years ago we were graduating no students interested in high school teaching. This year we had five. Our Graduate School of Education has just introduced a five-year certification program leading to a undergraduate degree in a subject area such as physics and a Masters in education, which promises to be very attractive to students.

Acknowledgments
The success of our program is the result of many people's contributions. I'd particularly like to recognize the support of our former Chair, Allen Robbins, the hard work of our Instructional Laboratory Manager, Dr. Michael Molnar, and the dedication and interest of Professors Mohan Kalelkar, Theodore Kruse, Peter Lindenfeld, Ronald Ransome, Joel Shapiro, and all my colleagues in the department.

Abstracted with permission from "Tripling the Number of Physics Majors at a Research University" in Conf. Proc. #399 The Changing Role ofthe Physics Department in Modern Universities, Proceedings of the Undergraduate Physics International Conference, J. Rigden & J. Redish, editors, 1997

Mentoring the Whole Life of a Physics Major: From Recruiting and Introductory Classes to Research and Careers

Neal B. Abraham

I would like to describe a complex mixture of mentoring activities, including some that are not traditionally thought of as mentoring which range from early activities in the recruiting of new students, through strategies in introductory and intermediate courses, to internships and research experiences and career counseling. The opening summary could also be the conclusion: What works? The answer is that many things work, and no one thing works for every student. To make anything work for a new student, the successful old programs often need to be repackaged, personalized, and invigorated with energy and compassion. And to make the task more difficult, what works one year often does not work the next. Successful programs are often forgotten by students from one year to the next and they may not be as successful the next time because the local needs and context have changed. You must listen carefully, act thoughtfully, assume nothing, and bring a renewed personal and friendly touch over and over again.

It is well documented that a far disproportionate share of students earning bachelors degrees in physics (and in mathematics and science more generally) from under-represented groups come from colleges and universities whose student populations have substantial numbers of students from those groups. Additional facts include the following: predominantly undergraduate colleges and universities have a disproportionate number of physics majors; research and career internships help both to attract and retain students; informal and formal peer teaching nurture confidence; teamwork and human-scale faculty members can have an immense impact on the social rewards of doing physics; and there is a synergistic effect of student peers sharing their academic pursuits. That institutions serving traditionally under-represented minority groups carry out their tasks with a certain missionary zeal, cannot be denied. But I think that a close look at these successful programs offers insight that can benefit all students in many different kinds of institutions. Indeed, this has been the consistent message in the studies and findings of Project Kaleidoscope of what works best in undergraduate mathematics and science education The programs leading to this success can be accomplished on many other campuses and they turn out to be equally valuable for women, men and members of under-represented groups.

The New York Times in November 1995 and Physics Today in August 1996 touted the numerical strength of the physics major program at Bryn Mawr College, a private liberal arts college for women which graduates a total of about 300 students each year. Approximately 40% of the undergraduate students take introductory physics in one of four different courses, approximately 30% of the graduates take their degrees in mathematics or science and, over the last two decades, the number of physics majors has grown steadily, albeit fitfully with both two and five-year periodicities in the numbers, until most recently. Currently, five percent of the graduates take (or will take) their degrees in physics, practically 100 times the national average for women as a percentage of the women in their graduating class. In 1995 Bryn Mawr's ten women physics majors were surpassed only by Harvard's 15 and MIT's 12. In recent years about 1-2% of the 150 women earning Ph.D.'s in physics each year earned A.B. degrees from Bryn Mawr and a similar number earned Ph.D.'s in related fields. But these represent barely a third of our majors; others are successfully pursuing medicine, law, high school and secondary teaching, and work in science museums, industries, and research labs. In 1997 we graduated 15 physics majors (five of them double majors in mathematics (3), biology (1) and philosophy (1)). We have an additional 15 senior physics majors enrolled for the Fall of 1997. Figure 1 shows the number of physics majors as a function of year. The circles represent projections.

Bryn Mawr College - Department of Physics Graduating Physics Majors

The specific causes of our surge in majors, countering national trends for men and women, are a little hard to identify. We believe that they are tied to a mixture of factors including recruiting, advising, introductory course strategies, intermediate encouragement, research opportunities, and the large and synergetic relationships that these women form with each other. Many of these features are part of what we might call "generalized mentoring".

So what do I mean by whole life mentoring? The answer is that we must seek to intervene and provide counsel, comment and insight at each stage of a student's thinking about physics. Most of all they are clever and attentive: they have read the publicity about the employment malaise, heard about the long hours, the difficulties of balancing families and careers, the abstract courses and the arcane testing hurdles. We start early by helping the admissions office during recruiting with posters, scripts for tour guides, and handouts for students and their families. We work with all students who want to take physics, entice some to take physics earlier, convince others to take physics at some point, recognize good work and encourage good students to continue. We provide a rich set of educational, learning and teaching experiences (both those in formal class and lab settings and those in informal consultations with faculty and fellow students. We affirm a variety of learning styles and a variety of demonstrations of mastery, encourage and arrange internships and research experiences throughout the four years of the undergraduate experience) counsel and encourage pursuit of a wide variety of careers, and, at each level, demand excellence and insist on involvement. We encourage all undecided students to consider taking our departmental placement exam which serves as a basis for assessment and counseling about starting points in the curriculum. We also work hard to make early contact with those qualifying for advanced placement by external AP exams or International Baccalaureate degrees, since some of those students are daunted by the maturity expected in sophomore courses. One early message in mentoring the whole life of a student is that you must stay in contact in order to provide advice and support. Waiting to talk to those who spontaneously enroll in the second year course may reduce potential majors to a third or less of those who might have continued successfully.

Introductory courses are most successful when they have a minimum of prerequisites, a combination of applications and an emphasis on conceptual understanding, and when students are encouraged to talk, write, discuss, and think about physics in more than chalkboard presentations, formalized homework problem solving and textbook-based memorization and regurgitation. Sometimes we use a conceptual and thematic approach. Sometimes we require oral reports and diary entries on readings of current events of scientific or science policy significance. Demonstrations can often be distracting and inconclusive but we are convinced that they have an important pedagogical value as a supplement to textbook descriptions and illustrations, blackboard sketches, and professorial narratives. We find that mixing demonstration apparatus with laboratory equipment gives students a sense of continuity and participation that improves their mastery. Our labs are relatively conventional, but we often try to see that they have a twist. Ours are rarely "prove the theory by experiment", or "fit the theory to the experiment", since some aspect of the idealized problem is tampered with to give anomalous experimental results. The student teams and the teaching assistants and instructors then search for explanations, reducing a larger class to only two or three investigators. We also try in lab to have different subgroups of students doing different things a design that is hard on instructors but challenging for the students. We use demonstration apparatus in the laboratories for "conceptual labs" (the "instructions" might be: take this miscellaneous collection of apparatus, figure out some interesting phenomena and questions, and write us an essay about the issue and the evidence). We also find it is important to build student confidence, especially in the introductory courses. It helps enormously to give encouragement in writing (on homework and exams) and in person. It can also help to use neutral colors for grading and comments: red is often taken to be harsh and critical. Sometimes we give midterms back in person to take the chance to offer a few comments or words of encouragement. Of course there is always positive benefit to returning exams quickly, with comments and interpretations, and will prompt debriefing to make testing and test reworking a part of the learning experience. Sometimes we approach students in lab or in the corridors to assure them that they are doing well enough to major. In short, we find that the best way to expand the pool of majors beyond the "hard core", self-selected and hard to deter, is to provide advice and encouragement.

We also have a vigorous program of research opportunities during the academic year and summers for students. Through a complex web of fundraising Bryn Mawr is able to support from thirty to fifty students doing summer research in mathematics and the sciences. Though opportunities are numerous, our population of majors far exceeds our ability to offer them summer internships. Hence we have a vigorous program to encourage students to look for research internships in industry and government labs and at other colleges and universities. In physics we have a faculty member who is designated to advise students about such job hunting, much of the Fall is spent in helping students identify options, prepare resumes that emphasize their interests and practical experiences in computing, electronics, and instrumentation. We find that students gain considerable maturity and confidence from working with those who had not taught them more elementary subjects and from returning to campus with summarized accomplishments which their local mentors had not seen pass through the foibled stages.

Our academic year student research program has also grown and evolved. We ask interested students to talk with each faculty member to gain an idea about research opportunities. We ask students to apply to the department, indicating their choices, and then use departmental discussions to assign students and to help faculty members balance their desires to be supportive and encouraging of eager students with practical issues of time management and student needs. We have often had trouble getting students to complete thesis writing at the end of a year-long project in a timely fashion, but this has gotten easier since we moved from a single end-of-the-year report to an alternative pioneered by our chemistry department. We now have students doing research give oral reports at the end of the Fall semester, giving them motivation to work through a formulation of the background, motivation, and goals for their ongoing work. This sometimes painful process, compelled midway through the project, makes the writing of introductory and background chapters of a thesis far easier to begin early the Spring semester. We then join with the chemistry departments of Bryn Mawr and Haverford colleges for a student poster session in a public forum late in the Spring semester. This is a good opportunity for the students to see the work of their peers and for underclass students to learn more about on-campus research opportunities. It is also a good time for faculty and students to mingle and to learn of the career choices of the seniors.

We also have a program sponsored by college funds to "apprentice" students as faculty members, so that they can see the whole life of a faculty member. In this program we are encouraged to help the students participate in the design and uncertain phases of a research project, in the assessment and ordering of equipment and apparatus from the instrument shops, and in regular reassessment of the goals and accomplishments. In some contexts it is argued that it is important to make a research project "successful" or "conclusive". Instead we have found that it is equally valuable for students to have some insight into the doubts, despair, and indecisions that are natural parts of our professional lives. We have a similar program for teaching apprentices and involve those students in preparing and assessing assignments, examinations and class presentations.

Giving students teaching opportunities is another form of mentoring. The coaching of student teachers helps to give them perspective on learning and pedagogy. The work of students as laboratory teaching assistants, tutors, monitors of a physics clinic for problem solving, graders, or in any facet of the program provides an opportunity for students to see other students making progress in their mastery. The students mentor each other. They seek advice and guidance from each other. They work together in teams both as students and as teachers. Much is made of the importance of role models, and we have found in the educational environment that one of the most important forms of role modeling is having peers, or near peers, fill the pipeline of success. When a student can see a role model at every level of success, she is hard pressed to make generalizations about gender difficulties and she finds many encouraging looks and words that she may soon be like them.

We have made some interesting choices about our intermediate and advanced curriculum. One goal is to provide a rich set of laboratory experiences, so students in our intermediate electronics course each have a bench with a full set of apparatus. While the students are encouraged to cooperate, visit each other's benches, and to ask questions and share answers, no one makes the experiment work (no other hands are on the knobs and dials) at each bench except the student herself. This may make some things slower, but it builds confidence over the semester that each student can operate oscilloscopes, design circuits, and analyze data. Other labs emphasize teamwork; some labs involve one-week projects, others are multi-session projects (some laboratory courses meet three afternoons a week) with freedom for students to explore different alternatives. And many labs are well equipped with research-grade apparatus which introduce students to sophisticated technologies and research style and strategies, both of which students may see again in research projects. When labs run long, we often take 20-minute breaks for tea (faculty supplied) and snacks (student supplied). Conversations are wide-ranging and vigorous and help to build camaraderie. The lost time is more than made up in the added efficiency that a break and a little sugar can provide.

In our junior and senior level course work we teach most courses in alternate years, in a coordinated program with the Haverford College physics department, so that most courses are offered every year at one campus or the other. Haverford's number of physics majors is about the same size as ours, though predominantly male as is typical in co-ed schools, and these courses, when students choose to mingle, give them a co-educational environment and a wider variety of instructors. Beyond the commingled courses, we also emphasize in these courses that work and material is not simply defined by texts and in-class presentations. We use library reserves and frequently ask students to prepare written or oral reports on supplemental topics. Nonetheless, both because of numbers and on principle, we have resisted merging our second year courses with those at Haverford., those prior to the declaration of the major. Experience teaches us that confidence is fragile for some women physics majors, and the added pressure and competition in larger and co-ed classes in the first two years contribute to a noticeable reduction in the number of women majoring in physics. We also find that students must be acculturated to making use of office hours, planning ahead to work on problem assignments, and developing teams for initial out of class discussions. Smaller sophomore-year classes make it easier for the faculty member to meet regularly with each student, inviting them for midcourse assessments in addition to handing back exams personally with short discussions about their performances and ways to improve.

Another way to mentor is through providing advice in a variety of media: printed handbooks, posters, and websites are among the ways we try to make information available. Our brochures, posterboards and website range over such topics as careers, preparing for teaching, preparing for the GRE, where recent graduates are working, and how to plan for different and flexible futures. We update them often and discuss them with students over pizza and soda in the evenings. We also frequently mix with students to discuss time and stress management, to review our curriculum, and to have meals with our colloquium speakers. Our evening (dinner time) speaker program held in the cafeteria sideroom has been the most successful way to draw students and has given them the opportunity of dinner with other physicists. We have found that our own contacts, alumnae, and the CSWP speakers list give us a nearly inexhaustible supply of women with diverse careers, talents, topics and stories. We rely less on the infrequent visits by outside women than on the daily support networks that develop among students within the department; programs and facilities range from a "majors' room" with computers and lockers, key access to kitchenette and computers and classrooms, desks in research labs for students doing research, mailboxes for messages, homework solutions in the conference rooms as well as in the more distant library, and student-run evening physics clinic for answering of questions. With a little luck and lots of synergy, our majors have come to think of the physics department as the place where they can and will find each other for teamwork or companionship, for problem solving or relaxation, daytime, nighttime and weekends. They mentor each other as much, indeed even more sometimes, than we mentor them.

There are other little touches which give "whole life" added meaning. There are bulletin boards in the department corridors which contain information about new science, women in science, summer jobs and careers, and activities of current majors and alumnae. The students design T-shirts, celebrating such things as "women in classically forbidden regions" and "strangely attractive women", demonstrating their scientific humor and their awareness of their unique participation as members of an under represented group in physics. The majors often organize field trips to industrial, university, and national labs and we arrange class schedules to support these trips.

And we provide current students with space to call their own, not only classrooms to use in off-peak hours, but rooms for majors with computers and lockers, mailboxes for departmental messages and for the return of exams and homework, and access to departmental facilities including kitchenette, microwave, and evening telephone (for local calls). As a result, the physics department is the "home office" and social center for many of our students. They come to find friends (and physics colleagues), they come to relax, as well as work, and they feel comfortable in their teamwork and in their differences.

Let me reiterate an earlier point that makes the task for all of us that much more difficult when it comes to mentoring. Our students are agile and active readers, they listen well and they hear our code words and our occasional despair. Every word about job shortages, the absence of women, chilly climate discrimination in laboratory work and research groups, hiring crunches, horrors of graduate school teaching, difficult working conditions for teaching assistants, and competitive testing and grading practices are heard and magnified. We must work hard to spread reality and we must caution students where appropriate. One of the reasons we post and report on our alumnae career paths is to assure current students that employment options remain open. We must counsel students against applying at some graduate schools where teaching is abysmal or where the expectations are for one prior year of graduate-level work, an inappropriate standard but one that is not uncommon when Masters-level foreign students and advanced placed undergraduate majors at research universities might gain that qualification. Poor teaching in graduate level courses is not a secret, so it behooves graduate schools to invest in burnishing their programs and thereby their reputations. And, in my considered view after three terms on the GRE Board of Examiners for the Physics test, continued use by graduate schools of GRE cutoffs (for admission or for continuation) are unforgivable kowtowing to indefensible statistical information. Nearly all of the top 1/3 of GRE physics scores are captured by foreign test takers. Students taught in small classes (in the liberal arts colleges where physics majors are disproportionately numerous) have essentially no experience with and no interest in the mental gymnastics of multiple choice examinations with a minute or two per problem. I know of no demonstrated value, valid correlation, or justifiable reason for the use of such a test as a barrier of any sort for pursuit of a career in physics. Except that most of us who are now physicists once survived or excelled in such a test, I challenge any user of the GRE to demonstrate a research or teaching skill that is adequately or appropriately measured by multiple choice gymnastics. Given the system, we necessarily coach our students and rehearse them to improve their scores and therefore to widen their opportunities for graduate school, but we consider it a relatively demeaning exercise that has minimal value in their physics education. When you know that the US medians are less than thirty right answers out of 100, that women at all kinds of institutions do one standard deviation worse than their male counterparts, that liberal arts college students do one standard deviation worse than their university trained counterparts, and that these differences are not correlated with any other measure of student performance, then I trust you will conclude as I have that the GRE exam is better discarded than reformed. And, in the meantime, if you do not mentor your students to do as well as they can and to be prepared for the worst score they have ever received on a standardized test, you will lose them in droves, particularly, as it turns out, the women and minority students. If any other proof were needed, the continued admission to and success of our majors at graduate schools around the country speak volumes about the limited value of the GRE score, as most would have been excluded from admission based on their scores alone.

In conclusion I suggest that mentoring has three primary tasks: giving honest advice, instilling confidence, and leaving room for growth. Among the best ways to do this are:

  • share secrets of successful teaching and learning strategies;
  • validate student mastery and career choices;
  • ensure a personal and socially supportive atmosphere;
  • be aware of the fragility of success.
  • And finally, for good and effective mentoring, keep asking, keep trying, and keep listening.

Neal B. Abraham is Chairperson of the Department of Physics at Bryn Mawr College.

Longer Version

Spring 1996 photo of most of the 40+ senior, junior and sophomore physics majors at Bryn Mawr

Browsing Through the Journals

Thomas Rossing

If you don't like the news, don't shoot the messenger. Several communications circulating on electronic forums have complained that by quoting different points of view in these columns, I have given tacit approval to these points of view. It is not the intent of this column, of course, to offer approval of any particular point of view. It is rather to call attention to news and views about physics education in various journals that FEd members don't necessarily read regularly. Hopefully, readers will find some of these snippets interesting enough to read the original articles and discuss them. Perhaps some readers will respond by writing letters to the editor for publication in the newsletter. We encourage this; all we ask is that they be kept reasonably short.

According to a speech entitled Attracting and Preparing Teachers for the 21st Century" by U.S. Secretary of Education Richard W. Riley, reprinted in the May 1997 issue of Community Update (DOE publication), the entire context of American education is changing. "We need teachers skilled in using computers as a powerful teaching tool, and many more teachers versed in English as a second language." State teachers of the year, asked to comment on the new teachers they had mentored, expressed strong concerns that new teachers are unprepared to manage classroom discipline, he warned.

The thesis that scientific knowledge is a cultural construct is challenged by Kurt Gottfried and Kenneth Wilson in a commentary in the April 10 issue of Nature. According to Andrew Pickering, who typifies what the authors call the "Edinburgh school of sociology", "The quark-gauge theory picture of elementary particles should be seen as a culturally specific producta communally congenial representation of reality...The preponderance of mathematics in particle physicists' account of reality is no more hard to understand than the fondness of ethnic groups for their native language." This view comes about, the authors suggest, because historians of modern physics have, until recently, paid little attention to experiment while exploring theory in great detail. Proponents of the Edinburgh school, however, tend to overlook predictive power, the strongest evidence that the natural sciences have an objective grip on reality.

According to an article in Science (Oct. 4, 1996), Japan has cut back its science requirements in high school from fifteen hours per week, in the 1960s, to eight and from 1048 hours in elementary school to 735 (compared to 450 hours in the United States). The Japanese authorities are concerned that students do not understand and enjoy science but rather engage in rote learning and cramming for university entrance examinations. The percentage of Japanese high school students taking physics has dropped from more than 90% in the 1970s to 20% at the present time.

On the other hand, a note in the May issue of Physics World relates the happy news that more students than ever have

applied to read physics at UK universities next year. About 80% of students now take GCSE physics or double award science, compared with 30% of boys and 10% of girls in 1980. On the other hand, the number of students taking A-level physics in schools continues to decline. In response to this, the Institute of Physics is launching a £1 million program to invigorate physics education for 16- to 19-year olds.

Australian universities have appealed to the government to increase funding of higher education to avoid the country falling behind its counterparts in Southeast Asia, according to a note in the May 15 issue of Nature. The appeal was made against a backdrop of increasing unrest on university campuses, partly because of cuts in funding abruptly implemented by the coalition government of Prime Minister John Howard last year. Recent student demonstrations in both Melbourne and Sydney protested the introduction of full fees for Australian students. Fay Gale, vice-chancellor of the University of Western Australia pointed out to the Government that Taiwan and Singapore now outstrip Australia by factors of two and three, respectively, in proportion of GDP spent on education.

"Active Learning in the Lecture Hall" is the title of a paper by biologist Elaine Anderson in the May 1997 issue of Journal of College Science Teaching. Essays, special team presentations, a field project, and concept mapping are among the active learning experiences successfully applied in large lecture sections for non-science majors. She cites the advice of Raymond Orzechowski (J. College Science Teaching 24, 347 (1995)), however: "Don't try too much in the beginning, especially with students who lack prior experience and skills in group learning."

The National Research Council (NRC) has issued a series of reports called "Preparing for the 21st Century" to inform policy makers and the public about conclusions by the NRC, the National Academy of Sciences, the National Academy of Engineering, and the Institute of Medicine. One of the reports entitled "The Education Imperative" urges that education be the Nation's top priority. Objectives include requiring that college entrance exams include a section that evaluates scientific inquiry, providing more grant money for graduate students, and setting institutional standards for graduate schools. The six reports can be found on the web site: http://www2.nas.edu/21st.

Science Teaching Reconsidered, a recent publication from the National Research Council, explores the teaching experiences of undergraduate science educators in an effort to "facilitate change in the way science is taught to students in U.S. colleges and universities." Science teaching, according to the NRC, is a mixture of creativity, imagination, innovation, and practical application. The handbook is available from National Academy Press (800-624-6242) for $10 plus $4 shipping/handling.

Parallel Parking an Aircraft Carrier: Revising the Calculus-Based Introductory Physics Sequence at Illinois

David K. Campbell, Celia M. Elliot and Gary E. Gladding

For many decades, calculus-based introductory physics course sequences - henceforth, "University physics" -have provided the foundation of the outstanding science and engineering programs at many of our nation's large research universities, including the University of Illinois at Urbana-Champaign (UIUC). These courses, many substantially unchanged since the post-Sputnik science education initiatives of the late 1950s to early 1960s, have allowed most of our students to master the necessary skills to succeed in careers in science and engineering, as well as in law, medicine, and education. Given this apparent success and the immutable nature of most of the core concepts of basic physics - Newton's laws of mechanics, electricity and magnetism, geometrical optics -the obvious question is, "Why change?"

This question becomes even more telling when the costs of changing are considered, especially for a research university. First, the sheer scale of the project is daunting -at the UIUC, we teach nearly 2000 students per semester, 70 percent of them engineering students. Second, as at other research universities, all our faculty, postdocs, and graduate assistants must maintain active, ongoing research programs, and these responsibilities must be balanced with their teaching assignments. Finally, to provide the necessary continuity for the students who take University physics each year, we must implement these changes on the fly, without the luxury of phasing them in gradually. This means that each new course must be introduced immediately after the old course ends. Overall, the process of fundamentally recreating University physics seems a very challenging (and perhaps unnecessary?) exercise - akin to "parallel-parking an aircraft carrier".

Nonetheless, despite the very real difficulties and expense inherent in undertaking major curriculum reform, we have elected to go ahead and rock the boat. (Or in our case, the aircraft carrier.) Why? First, as recent physics education research has made painfully clear, traditional physics pedagogy has often been surprisingly ineffective in conveying fundamental conceptual understanding,1 as distinct from rote learning and formulae manipulation. While successful students are able to solve problems by the "tyranny of technique", recent studies have shown that even they frequently misunderstand the most fundamental concepts, partly because they maintain deeply entrenched misconceptions about basic physics that conventional pedagogy has failed to dislodge.2 Second, there are many individuals for whom traditional pedagogy has proven to be woefully counterproductive, leading neither to conceptual understanding nor to calculational dexterity, but rather to an utter disenchantment with physics.

In addition to these well-documented shortcomings in traditional instructional methods, the motivation for change comes from other needs that are also often unmet by traditional University physics - promoting collaborative learning, teamwork and communication skills, motivating research faculty to employ new instructional methods, training graduate teaching assistants to be effective teachers, and standardizing effective pedagogy so that successful learning is independent of the efforts of one inspired teacher.

For all these reasons, over the past two years our Department has completely restructured University physics at the UIUC. A timeline for this project is presented in Fig. 1. Importantly, this effort arose simultaneously from the convictions of an energetic group of faculty dedicated to instructional change, and the vision of a departmental leadership committed to innovation to meet the explicit needs of our students and faculty, our client departments in the College of Engineering, and the organizations and institutions that employ our graduates. Our guiding philosophy, deriving from the strong theoretical base of recent physics research3 has been: (1) to stress conceptual understanding as well as problem-solving skills, (2) to provide many paths to this understanding in order to accommodate diverse student learning styles, and (3) to make the students active participants in each path.1,4 It has been an exhilarating and exhausting experience.

Figure 1: The three bars illustrate the four-year transition from the former three-course University physics sequence to the revised four-course sequence. The numbers in parentheses below each course name denote the number of faculty assigned. One semester's worth of faculty time is devoted to the initial design of each new course during the last semester the old course is taught, and the new courses are initiated immediately thereafter. The three-semester refinement/standardization stage allows for adjustments in response to testing and evaluation before the courses go into the production phase.

Objectives for our Curriculum Revision
In planning our curriculum revision, we first developed a set of objectives, which included:

  • Adopting new "best-practice" instructional techniques, based on physics education research, that emphasize conceptual understanding.5,6,7
  • Utilizing state-of-the-art instructional media, including multimedia presentations, World Wide Web-based interactive course materials,8 and laboratory computer data acquisition and analysis.
  • Promoting student opportunities for collaborative learning and teamwork.9
  • Standardizing meaningful course content and effective pedagogical methods, so that good teaching is not dependent on a single inspired instructor but is integral to all sections of all classes,10 while allowing room for faculty creativity and continuous improvement.
  • Building an administrative/management infrastructure to support and sustain continued curriculum development as new methodologies evolve.

A significant consideration in redesigning the courses was to develop a comprehensive curriculum in modular units that could be combined in different ways by departments in the College of Engineering, according to their students' perceived needs. We have thus taken the earlier four-credit-hour courses - Physics 106 (Mechanics), 107 (Electricity and Magnetism plus Thermodynamics), and 108 (Geometrical Optics, Waves, and "Modern Physics") - and reworked them into Physics 111 (Mechanics) and 112 (Electricity and Magnetism) - still four-hour courses - and Physics 113 (Fluids and Thermal Physics) and 114 (Waves and Quantum Physics), which are now two-hour courses. Thus, the courses can be flexibly combined to form one 12-hour or two 10-hour sequences.

Physics 111 replaces Physics 106, with the addition of a unit on transverse waves (previously presented in Physics 108). With the deletion of the thermodynamics section of Physics 107, Physics 112 now provides a unified topical coverage of electricity and magnetism. Practical treatments of electromagnetic waves, polarization, and geometrical optics (previously in Physics 108) now round out Physics 112. Physics 113 presents an introduction to fluid mechanics and augments the coverage of thermodynamics (previously in Physics 107) with some ideas of the microscopic origins of the basic concepts. Physics 114 contains a practical treatment of wave interference and an introduction to quantum physics, topics that were previously covered in Physics 108. We invite you to examine our new course outlines and materials.

Fine-Tuning the Changes: Refinement and Standardization
As shown in Fig. 1, after the initial one-semester design phase, we plunge immediately into teaching the course. We have planned a three-semester "shakedown cruise", when the curriculum and instructional changes can be tested and refined. The goal of this phase of development is to arrive at a basic core curriculum that is both meaningful and effective -ot dependent for success on the heroic efforts of one inspired and inspiring teacher - while at the same time offering the flexibility to accommodate individual faculty creativity and continuing improvement.

For example, in response to student mid-term and end-of-term evaluations in Physics 111, we have added more "conceptual" problems to the discussion sections and hourly exams, added some true/false and 3-part multiple-choice problems and some "optional" problems to the homework, in order to provide increased practice with these types of questions. We have reduced the number of activities in several of the new labs, having initially underestimated the time it would take students to complete them.

We are also using this time to develop, test, and refine a large bank of problems for discussion sections, homework sets, quizzes, and exams. We track student scores problem by problem and then use the results to either eliminate a problem or assign it to a better usage. For example, some questions clearly are not amenable to true/false or multiple choice solutions, but are well-suited for discussion sections. Other problems that appear to require considerable pondering might not be appropriate for time-limited exams but would offer good exercise for extra-credit homework questions.

It is important to note that we maintain a distinction between "standardization" and "fossilization". While our goal is to develop uniform, tested, effective curriculum modules, we also must support faculty innovation and creativity. Thus, faculty teaching the courses for the first time are encouraged to develop their own problems, design their own ACTs, and create their own lecture presentation materials. We are then able to test the effectiveness of this new material in our 2000-student "laboratory" keeping the best of it and thus incorporating incrementally improved material each semester.

"Institutionalizing" the Changes
Before discussing the five components of each of the new courses, we should emphasize our view that the long-term success of any curriculum revitalization can only be guaranteed if the changes can be institutionalized, i.e. can be designed to remain in place after faculty members initially involved in the revised courses move on to other endeavors, and new faculty take over.

Our Department has a tradition of instructional collective ownership: that is, our faculty do not have property rights to specific courses, but rotate to new teaching assignments every three to four semesters - reducing faculty burnout by distributing the burden of the more time-consuming and difficult courses. We find this approach increases interactions among the faculty in our large and broad department, promotes departmental collegiality, and improves overall instruction, while allowing us to maintain some "quality assurance" over the course content. We thus maintain a prescribed core curriculum and comprehensive files of instructor's notes, lesson plans, homework assignments, special projects, and exam questions for each course, so that a first-time instructor has a substantial set of pre-tested material to begin from. Despite these efforts, the elementary courses, in which many students view themselves as unwilling conscripts, have remained the most demanding and least satisfying teaching experiences. Not surprisingly, faculty have accepted their assignments to these courses stoically but not enthusiastically and were very glad when their tours of duty were complete. Clearly, unless teaching University physics were to become more rewarding for the faculty, the prospects for institutionalizing the improvements that curriculum reform offered were dim.

Our solution for Physics 111-114 has been to employ team teaching, in which a group of faculty divides up the responsibilities for the various course components - i.e. each course has main and back-up lecturers, labmaster(s), and homework master(s) - and the team members are also rotated through these courses, such that each new instructor is matched with experienced faculty who have either participated in the initial course design or taught the revised course. This shared responsibility ensures that faculty assigned to these 700+ student courses regard this as an ordinary teaching assignment, not one requiring superhuman effort, and enhances institutionalization of content and methodology by providing on-going contact between experienced faculty and new team members.

From the faculty perspective, we have already noted two indicators of faculty satisfaction: (1) the original Physics 111 developers have chosen to remain in the course (having already served their original commitment of three semesters) but have exchanged course responsibilities, a practice we would like to encourage, and (2) seven new faculty members, in addition to the six original developers, have successfully and enjoyably taught on Physics 111 and 112 teams. Given that research faculty have historically disliked teaching traditional University physics courses nearly as much as students have hated taking them, we think this is remarkable.

Course Components
Lectures: We have altered the traditional lecturing format substantially, using interactive, multimedia lecture presentations that incorporate active learning segments (ACTs), typically three per presentation. These ACTs, which are motivated by research that has shown that students must be intellectually active in order to develop "functional" understanding,1 are patterned after the ConcepTests developed by Eric Mazur.4 Importantly, many ACTs involve demonstrations to illustrate correct physics intuition and to reinforce the basic concepts being presented. In addition to the instructor, two "stairmasters" (typically senior graduate teaching assistants [TAs] or faculty who will be teaching the following semester) are assigned to each session; their job is to circulate through the auditorium and provide guidance and facilitate discussion during the ACTs. The interactivity promoted by the ACTs, both among the students, between the students and the presenter, and between the students and the stairmasters, results in classroom dynamics that are quite different from conventional large-audience lectures, and substantial increases in student attendance under the new format, compared to historic norms, have been observed.

Discussion Sections: The two-hour discussion sections for these courses feature group work on problems emphasizing conceptual understanding that have been created by senior faculty, not TAs. Our original intention was to create "context-rich" problems patterned after those of the U. Minnesota group.11 We have found that this approach works quite well for the Physics 111 (Mechanics) course, but we have thus far had to abandon it in Physics 112 (Electricity & Magnetism), because of its more abstract content. While we are hopeful of eventually altering this situation, our experience mirrors that of the physics education community, where the well-tested Force Concept Inventory12 and the Mechanics Baseline Test13 provide accepted standards for assessing the knowledge of mechanics, but no similarly broad tools currently exist for testing knowledge of electricity and magnetism.

The format of the new discussion sections has required significantly increased attention to the training of our TAs. Instead of merely working calculational problems as the students watch passively, the TAs are now expected to act as facilitators for group learning and to emphasize conceptual knowledge-based problem solving, i.e. to guide students to the solution instead of telling them the answer. Increased emphasis has necessarily been placed on effective instructional methods, and we now separately train discussion TAs and lab TAs.

Laboratory Sections: Our former University physics labs - like those at many peer institutions - were "observe and measure" only, often with too much emphasis on the details of measurement (e.g., error analysis). The new two-hour labs feature experiments based on the "predict-observe-explain" approach of Thornton & Sokoloff 14 to more actively engage students in the learning process and to promote mastery of concepts by manipulation of experimental apparatus. We have also adopted the use of pre-lab assignments, consisting of several questions designed to prepare the students for the concepts and exercises presented in each lab. Scripted lab reports, which can be finished within the class time, are employed.

Homework Assignments: Our Department's more than twenty-five years' experience with computer-aided physics education provides considerable evidence of the effectiveness of requiring students to interact with a set of carefully constructed, incremental homework exercises, which progressively build both the student's sum of factual knowledge and his or her abilities to synthesize and apply this knowledge in practical problem-solving.15 Homework sets developed for the new curriculum consist of problems that the students solve using CyberProfTM (CP), an interactive Web-based learning environment created by one of our faculty as an outgrowth of his complex systems research. CyberProfTM offers a number of advantages over conventional computer-assisted homework sets because of its Web-based implementation, platform independence, and comprehensive feedback. In its current version, CP is able to recognize even a mathematically ambiguous or partially correct answer and, using an interactive series of Hints and Helps, to guide the student to the correct answer. A drawing tool records and analyzes graphical input, and a what-if feature is planned that would allow a student to change a problem's variables to generate additional related questions, thus promoting self-testing.

Exams: Student attention inevitably focuses on the exams, since their performance on exams is the dominant factor in their final grades. Consequently, "conceptual" questions must be included on exams, if we wish to convey the importance of functional understanding.1 When only traditional calculational problems are given on exams (as in our previous Physics 106-108 sequence), students develop problem-solving "routines" that are often based on shortcuts, learned through repetition, and applied unthinkingly, rather than being derived from the concepts the problems were designed to test. We have thus adopted an exam format that includes approximately 25 percent conceptual questions.

In order to grade answers fairly and uniformly for the large number of students taking these classes, we have adopted a machine-gradeable, true/false and multiple-choice question format, which we believe tests conceptual understanding. Our initial experiences with this exam format have been quite positive, but we plan to undertake rigorous professional assessment of this method to validate it as an accurate tool for assessing a student's conceptual understanding.

The Last Word(s)
Has it been easy? No! Would we do it again? YES!! We believe that physics is the "liberal arts education for a technological society" and that excellence in physics is critical to maintaining scientific, technological, and economic vitality in a world where U.S. leadership can no longer be taken for granted. Thus, we must do a better job for our students in conveying conceptual understanding, in promoting teamwork and communication skills, and in recapturing the interest and enthusiasm of those who have too often before been disillusioned and left behind. Our preliminary assessment of the University physics revision undertaken at UIUC is strongly favorable. We have glowing testimonials from students and faculty that express great enthusiasm for the changes, and evaluation surveys that show statistically significant improvements in student satisfaction with the new courses, compared to their forerunners. In the years to come, we will continue to work to achieve our final objective - to support and sustain continued curriculum development as new insights emerge and methodologies evolve.

This effort has been supported in part by the Hewlett-Packard Company, the AT&T Foundation, and the Shell Foundation, and their vision is gratefully acknowledged.

  1. L.C. McDermott, Am. J. Phys. 61, 295-298 (1993); and Am. J. Phys. 59, 301-315 (1991).
  2. Ibrahim Abou Halloun and David Hestenes, Am. J. Phys. 53, 1043-1055 (1985); Am. J. Phys. 53, 1056-1065 (1985).
  3. Arnold Arons, Homework and Test Questions for Introductory Physics Teaching (New York, J. Wiley, 1994); Edward F. Redish, Am. J. Phys. 62, 796-803 (1994); Jose Mestre and Jerold Touger, Phys. Teach. 27, 447-456 (1989).
  4. E. Mazur, Peer Instruction: A User's Manual, Prentice Hall, Upper Saddle River, NJ (1997); Sheila Tobias, Revitalizing Undergraduate Science: Why Some Things Work and Most Don't, (Tucson, AZ, Research Corporation, 1992).
  5. Edward F. Redish, "New Models of Physics Instruction Based on Physics Education Research", invited talk presented at the 60th meeting of the Deutschen Physikalischen Gesellschaft, Jena, Germany (14 March 1996).
  6. Alan Van Heuvelen, Am. J. Phys. 59, 891-897 (1991).
  7. William J. Leonard, Robert J. Dufresne, and Jose P. Mestre, Am. J. Phys. 64, 1495-1503 (1996).
  8. A.W. Hübler, "CyberProf: A New Way of Teaching and Learning", keynote speech presented at the National Science Foundation LACEPT: Teaching Mathematics and Science at the Undergraduate Level, Baton Rouge, LA (26-27 January 1996); see also Alfred W. Hübler and Andrew M. Assad, "CyberProf: An Intelligent Human-Computer Interface for Asynchronous Wide-Area Training and Teaching", invited talk presented at the Fourth International World Wide Web Conference, Boston, MA (12-13 December 1995).
  9. Patricia Heller, Ronald Keith, and Scott Anderson, Am. J. Phys. 60, 627-636 (1992); Patricia Heller and Mark Hollabaugh, Am. J. Phys. 60, 637-641 (1992).
  10. J.W. Harrell, "Freshman Physics in the NSF Foundation Coalition", Forum on Education of the Am. Phys. Soc., 6-7 (Spring 1997).
  11. Kenneth Heller, Patricia Heller, and Mark Hollabaugh, Cooperative Group Problem-Solving in Physics, (Minneapolis, MN, University of Minnesota, January 1994).
  12. David Hestenes, Malcolm Wells, and Gregg Swackhamer, Phys. Teach. 30, 141-151 (1992); and Douglas Huffman and Patricia Heller, Phys. Teach. 33, 138-143 (1995)
  13. David Hestenes and Malcolm Wells, Phys. Teach. 30, 159-166 (1992).
  14. D.R. Sokoloff, R.K. Thornton, Motion and Force Laboratory Curriculum and Teachers' Guide, Vernier Software (1992); D.R. Sokoloff, P.W. Laws, R.K. Thornton, Real Time Physics: Active Learning Laboratories, Electricity (1993); R.K. Thornton and D.R. Sokoloff, Am. J. Phys. 58, 858-867 (1990).
  15. L.M. Jones and D.J. Kane, Am. J. Phys. 62, 832-836 (1994).

The authors are from the Department of Physics at the University of Illinois at Urbana-Champaign, Loomis Laboratory, 1110 West Green Street, Urbana, IL 61801-3080

Preparing Physics Majors for Secondary-Level Teaching: The Education Concentration in the Haverford College Physics Program

By Lyle Roelofs

It is easy to document both the strong demand for physics teachers at the secondary level1 and the fact that not all individuals currently in those positions are well qualified.2 Many undergraduate physics majors who might otherwise be interested in teaching high school physics, however, do not pursue that career option because the requirements for certification are quite strenuous in many states. We have accordingly developed at Haverford College a concentration in education for physics majors which provides experiential preparation for teaching physics but requires fewer courses beyond the standard physics major than does the typical curriculum leading to certification. I described this program at the recent Conference of Chairs held in College Park, MD and it is discussed in more detail in a forthcoming issue of the American Journal of Physics.3 This brief summary may be of interest to members of the Forum on Education. Readers desiring further information are directed to the AJP article or to our website.

The 'concentration' is a structure in the Haverford College curriculum consisting of a total of 6 courses, 2 or 3 of which also may and must count toward the student's major requirements. It is thus similar in weight to a 'minor', but differs in being more closely tied to a particular major. Our Education Concentration consists of: four courses offered through the Education program at the college providing a general introduction to education and a final semester summary seminar; and two novel courses developed by and offered in our department in which the student learns, by doing, how to teach physics. These latter courses are typically taken by advanced undergraduate physics majors and involve participation in the instruction of our introductory course for non-majors. One of the two involves the student in teaching laboratory physics -- activities include presentation of pre-lab comments, a critique of an existing experiment, and the development and testing of a new experiment. The other course involves the student in the classroom portion of the introductory course. He or she attends and critiques class sessions, participates in the development and grading of exams, leads sessions providing individualized assistance in problem solving, leads one class session during the semester using peer instruction techniques4, develops a demonstration to use in that class presentation and becomes familiar with the modern literature on physics pedagogy.

Although a program leading to certification in secondary education is available at our institution, most of our majors who are interested in teaching have opted for the concentration route described here. The career options for a B.S. physics major afforded by the concentration include proceeding directly to a teaching position in a situation that does not require certification. (Most private schools do not require starting teachers to be certified, and in addition many states--19 as of this writing--have approved so-called Charter Schools which operate with public funding, but under charters that relax many of the strict mandates that govern teacher appointment in public education.) Or a student may enter an M.A. program in teaching and obtain both that degree and certification in a little over a year, thus becoming highly qualified--and also highly sought after--for teaching positions in any school setting. Since 1993 eleven of our graduating majors have gone on to education careers: one obtained certification as an undergraduate here; eight moved directly into teaching positions with just their B.S. in physics, most having taken the Association courses; and two obtained Masters degrees with certification before beginning to teach.

References

  1. "Teacher Supply and Demand in the United States" published by the American Association for Employment in Education, 820 Davis St., Suite 222, Evanston, IL 60201-4445.
  2. See "The Condition of Education 1996" issued by the National Center for Education Statistics, a department of the US Education Department. http://www.ed.gov/NCES
  3. L. D. Roelofs, Am. J. Phys. (to appear).
  4. Eric Mazur, Peer Instruction: A User's Manual (Prentice-Hall, Upper Saddle River, NJ, 1997).

Lyle D. Roelofs is Professor of Physics at Haverford College

The 1997 Conference of Physics Department Chairs

By Roger D. Kirby

The 1997 Conference of Physics Department Chairs was held May 9-11 at the American Center for Physics in College Park, MD. More than 170 representatives from a wide variety of educational institutions attended, making it the largest such conference in recent memory. The conference topic - Undergraduate Education in Physics: Responding to Changing Expectations - accurately reflects the sense of the conference. Undergraduate physics programs are under increasing pressure from university and college administrations, industry and funding agencies to better educate and train our students at all levels. The expectations of our programs have changed, and evidence is mounting that they need revitalization; in particular, most programs have a small number of majors with respect to faculty size, and many faculty and students have expressed dissatisfactions with their experiences, particularly in the introductory courses.

The recent NSF report "Shaping the Future" recommends that each science department "Set departmental goals and accept responsibility for undergraduate learning, with measurable expectations for all students; offer a curriculum engaging the broadest spectrum of students; use technology effectively to enhance learning; work collaboratively with departments of education, the K-12 sector and the business world to improve the preparation of teachers (and principals); and provide, for graduate students intending to become faculty members, opportunities for developing pedagogical skills."

This meeting was intended to help Department Chairs provide the leadership needed to advance their programs along these lines. The program included invited talks, breakout sessions, and informal opportunities for participants to benefit by sharing ideas and experiences informally chairs from other institutions. Some participants provided brief summaries of innovations from their own institutions.

The conference began with a Friday evening dinner followed by talks by Bob Hilborn (Amherst College) and Duncan McBride (NSF), to set the stage for the Saturday and Sunday portions of the program. Hilborn discussed the need and justifications for attempting to revitalize our undergraduate programs in physics while McBride discussed NSF's role in undergraduate education.

The Saturday morning session was devoted to the role of physics education research in improving undergraduate education and active engagement methods of instruction. Lillian McDermott (University of Washington) and Joe Redish (University of Maryland) discussed recent advances in physics education research, and in tutorials demonstrated ways to implement the results of such research. There is now considerable evidence that so-called active engagement methods offer the possibility of substantially better student learning and attitudes. Eric Mazur (Harvard University) demonstrated how active engagement techniques can be effectively implemented.

A session entitled "Flexible Curricula" dealt with changes in undergraduate major programs which can help attract more majors to our departments and which can better serve the needs of the students. All speakers emphasized that physics as a profession cannot accommodate large numbers, so departments wishing to attract more students must actively work to build links to other professions and disciplines. Joe Pifer discussed Rutgers University's development of several different "tracks" for majors and noted that it had led to a tripling of the number of majors. Vijendra Agarwal (Moorhead State University) discussed his department's development of several tracks with "concentrations" in other disciplines and internships with local industries. Lyle Roelofs described Haverford College's successful concentration in secondary education.

In a session on curricular innovations at the introductory physics level, David Campbell (University of Illinois) discussed his department's reworking of the introductory calculus-based physics sequence. Louis Bloomfield (University of Virginia) described a course called "How Things Work", for non-science majors. The course uses everyday objects such as bicycles, microwave ovens, etc., as vehicles for introducing and discussing physics concepts. The course is currently taught to hundreds of students each semester.

After dinner on Saturday Bob Eisenstein (NSF) discussed The Future of Physics: A View from Washington. Bob emphasized the great excitement that is evident in the physical sciences, but noted that funding uncertainties for physics research are real, and that we have to more effective in taking science to the general public in a variety of venues.

Successful undergraduate programs that include women and minorities require substantial attention to mentoring and advising. James Stith (Ohio State University), Neal Abraham (Bryn Mawr College), and Priscilla Auchincloss (University of Rochester) discussed programs that can make a real difference in recruitment and retention. Stewart Smith (Princeton University) showed how a universal requirement of undergraduate research can succeed with students having a wide range of abilities and interests.

Many other issues were discussed through breakout sessions of about 20 participants. If undergraduate education is to be taken seriously, reward systems for faculty members should reflect an institutional commitment. We need to test our efforts by becoming better informed about student assessment and measurement of learning, and we need to utilize undergraduate research more frequently as a way of facilitating student intellectual and personal development.

The 1997 Conference of Physics Chairs was co-chaired by Roger D. Kirby (University of Nebraska) and Jerry P. Gollub (Haverford College). A more complete report on the Chairs Conference can be found at the American Physical Society website.

The AAPT Workshop for New Physics Faculty

Kenneth S. Krane

The first AAPT Workshop for New Physics Faculty was held from October 31 - November 3, 1996, at the University of Maryland and the American Center for Physics. The Workshop's purpose was to promote expertise in teaching among recently hired faculty, especially those at the research universities, where developing expertise in teaching often receives less emphasis (and less reward) than developing expertise in research. Yet it is at this stage of a new faculty member's career that teaching habits are formed, often by emulating senior faculty at the institution or the faculty member's own undergraduate instructors, who may not necessarily provide effective role models and who may not be following the curricular and pedagogic changes that are emerging in the physics community. Moreover, the lack of attention to good teaching at the research universities sends a subtle but significant message to graduate students, and in this way poor teaching practices infect the next generation of physics teachers at all types of institutions. It was for this reason that the research university faculty were selected as the primary audience for this program. In particular, we targeted faculty in the first year or two of their initial tenure-track appointment.

To enable the development of a national program for improving physics teaching, the AAPT was awarded a grant by the Undergraduate Faculty Enhancement program of the National Science Foundation. Nominees were selected based on letters from department chairs. The NSF grant supported all expenses of the participants except the cost of travel to the Workshop site. Staff support for the Workshop was provided by the AAPT.

The 1996 Workshop began with an afternoon at the NSF headquarters and meetings with research program directors in physics, astronomy, and materials science. The workshop itself was designed to be highly interactive, with the format consisting of plenary presentations followed by interactive breakout groups. The opening speaker was Lillian McDermott of the University of Washington, who spoke on "Learning about Conceptual Misunderstandings Bridging the Gap Between Teaching and Learning." After her talk, Lillian led the participants through a working session based on the tutorials in introductory physics being developed by her research group. Other speakers included:

  • Eric Mazur (Harvard University) on "Active Learning and Interactive Lectures"
  • Bob Eisenstein (Director of the Physics Division of NSF) on NSF interest in research and education
  • Cathy Olmer (Indiana University) on "Undergraduate Research"
  • Jim Stith (The Ohio State University) on "Recruiting and Retaining Physics Majors."
  • Bob Beichner (North Carolina State University) on "Using Technology to Teach Physics"
  • Sara Majetich (Carnegie Mellon University) and Luz Martinez-Miranda (University of Maryland) on "Women and Minorities in the Classroom,"
  • Wolfgang Christian (Davidson College) on "Using the World Wide Web in Teaching,"
  • Ken Heller (University of Minnesota) on "Being a Role Model for Your Teaching Assistants."
  • Duncan McBride (NSF's Division of Undergraduate Education) "Shaping the Future: New Expectations for Undergraduate Education"
  • Diandra Leslie-Pelecky (University of Nebraska) on "Outreach Programs"

A highlight of the workshop was the "Physics IQ Test" demonstration show presented by Dick Berg of the University of Maryland.

The Workshop evaluation generated extremely enthusiastic responses from the participants, despite their exhaustion from the full schedule. Many commented about the great variety of topics of which they were previously unaware but which they expected to have immediate or long-term impact on their teaching. They were also appreciative of the opportunities to interact with peers and with the discussion leaders in the small breakout groups.

The participants were also asked to respond to two additional questions: "What excites you about being a faculty member?" and "What is the biggest frustration you face as a new faculty member?" Nearly all replied that teaching and communicating about physics were particularly exciting to them. They seemed to be frustrated over the lack of time (and support) available to develop the research and teaching expertise they see as necessary for promotion and tenure. When asked what they would say if given the opportunity to address the next physics department chairs conference on the subject of improving the environment for junior faculty, nearly all replied that there was need for a more effective mentoring system and for clearer statements of the expectations for advancement in rank.

The NSF grant also provides funds for a follow-up to the Workshop to be held at the summer AAPT meeting in Denver in August 1997. At this meeting there will be a session devoted to the challenges and opportunities for new faculty, at which several participants in the Workshop will give talks describing how they have applied the lessons learned at the Workshop. The talks will be followed by an open "cracker-barrel" discussion, which will allow all new faculty attending the meeting to share their experiences and perceptions.

The second Workshop for New Physics Faculty will be held at the ACP in College Park on October 30 - November 2, 1997. The primary target audience is newly hired faculty from the research universities; however, if space is available participants from other institutions will be included. Department chairs are invited to nominate their recently hired faculty with a letter to the AAPT Executive Office indicating the principal teaching and research expectations of the nominee as well as the department's willingness to support the nominee's travel to the Workshop site.

Ken Krane of Oregon State University is Head of the Workshop for New Physics Faculty Steering Committee.

ERRATUM

The authorship of the article entitled "Freshman Physics in the NSF Foundation Coalition" in the last issue of the FEd newsletter should have been credited to J.W. Harrell and Jerry Izatt.

Call for Nominations

Nominations for APS Fellows through the Forum on Education are needed. Send suggestions to Bev Hartline, chair of the Nominations Committee.


This Newsletter, a publication of The American Physical Society, Forum on Education, presents news of the Forum and articles on issues of physics education at all levels. Opinions expressed are those of the authors and do not necessarily reflect the views of the APS or of the Forum. Due to limitations of space, notices of events will be restricted to those considered by the editors to be national in scope. Contributed articles, commentary, and letters are subject to editing; notice will be given to the author if major editing is required. Contributions should be sent to any of the editors.