New Approaches to Chemistry Teaching. 2005 George C. Pimentel

Apr 1, 2006 - Traditional pedagogy has held sway for 2000 years. The studies of educational psychologists, cognitive scientists, and classroom ...
3 downloads 0 Views 149KB Size
Chemical Education Today

Award Address

New Approaches to Chemistry Teaching1 2005 George C. Pimentel Award, sponsored by Dow Chemical Co. by J. N. Spencer

The George C. Pimentel Award has particular significance for me. My graduate thesis on regular solutions followed the work of Joel Hildebrand, who was the first recipient of this award in 1950. Most of my subsequent research with undergraduates has been on hydrogen bonding, and a copy of George Pimentel’s book on this subject is never far from my workbench (1). His other works on bonding have had a profound influence on me. George Pimentel was actively involved in science education. He was a principal mover in Chem Study (2), a national project designed to improve high school chemistry teaching. The Pimentel Report on “Opportunities in Chemistry” (3) aided many students in a career choice for science. I am humbled to be given the honor to write on a topic so important to George Pimentel. New Approaches to Chemistry Teaching In the ancient Italian university town of Pavia there is a relief dating from the 1400s showing a master lecturing while students absorb and record (4). Just recently at Alexandria, Egypt, a 2000-year-old lecture hall was discovered (5). Thus, depending on your point of view, either the traditional approach to teaching has a long and honorable history, or teaching has a long and not so honorable history of resistance to change. Almost every aspect of science—except pedagogy— has undergone a substantial change in the past 2000 years. This reminds me that Thomas Paine pointed out in an altogether different context—“a long habit of not thinking a thing wrong gives it a superficial appearance of being right” (6).

What’s Wrong with the General Chemistry Course? 1. The course contains too much material. 2. The course content is determined too much by perceived needs of chemistry majors and not enough by the needs of the majority of the students in the course. 3. The course is not a general chemistry course because there is too much emphasis on physical chemistry and not enough on inorganic and organic chemistry. 4. Too much difficult and abstract theory. 5. There is too much emphasis on the solving of numerical problems.

528

Journal of Chemical Education



There are those who have noticed this reluctance to change and urged a reconsideration of age-old pedagogical practices. Cornog and Colbert observed that, “The aim in present texts and in much instruction is a compromise between the largely descriptive chemistry of a few decades ago and the ideas of physical chemistry which have been incorporated in elementary courses during the past 20 to 25 years. The result is an encyclopoedic hodge [sic] of the old and new, too extensive for thorough treatment in the time usually allotted, and so rich in multiplicity of slightly related details as to bewilder the earnest student. Clearly a pruning process is desirable.” (7). This prophetic report appeared in the first issue of the Journal of Chemical Education in 1924. Some 70 years after the Cornog and Colbert report, Ron Gillespie, a member of the Task Force on the General Chemistry Curriculum, noted the continuing problems with general chemistry as indicated in the list on this page (8). Educational Testing Services (ETS) regularly surveys introductory college chemistry courses. In the mid-1980s the ETS survey led to the conclusion that, “[t]he college general chemistry course is crowded with respect to the number of topics it covers. This is true despite the oft-cited criticisms that the curriculum of introductory college chemistry attempts to cover more topics than students can be reasonably expected to learn” (9). Subsequent surveys have shown no significant differences from the conclusion reached in 1990.2 Cornog and Colbert were not unaware of the pace of change expected when they wrote in 1924: “…undignified haste is not to be numbered among the sins of teachers of elementary chemistry” (7). Most of these calls for rethinking—at least for the general chemistry course—focused on content. George Bodner, however, has pointed out that “changing the topics being taught may not be enough” (10). Indeed, the push to include all possible content has forced out process. A common reason cited for refusing to consider change in the general chemistry course, and in courses in general, is that “we need to maintain our standards and rigor”, which is sometimes another way of saying that no change can be considered. In particular, in order to introduce everything a student may encounter in four years of a chemistry major, topics are watered down so they can be presented in a first-year course, but the integrity of the concept is often lost in the reworking. Thus the introductory course can run afoul of claims to standards and rigor and end up with the opposite of what was desired. Twenty years ago Bent asked for a reconsideration of the use of thermodynamics in general chemistry: “Should Thermodynamics Be X-rated? Speaking censoriously, I’d say yes and no. Yes, regarding formal thermodynamics—black magic on blackboards—for beginning students. No, regarding informal

Vol. 83 No. 4 April 2006



www.JCE.DivCHED.org

Chemical Education Today

thermodynamics—descriptive “chemystery” on lecture and lab benches—for all students” (11). Gillespie also questioned the intrusion of thermodynamics into general chemistry, asking, “do all students…need to struggle through…dubious derivations of ⌬G = ⌬H – T⌬S…?” (8). There are several instances of confusion in thermodynamics. The addition of rather subtle content has produced an incorrect explanation or a model that leads to incorrect understanding by students. A common confusion occurs with the units on thermodynamic parameters. The intensive nature of enthalpy, entropy, and Gibbs energy requires the “per mol” unit. This difficulty becomes apparent if the units assigned to ⌬G do not have a per mole designation; e.g., J/mol. The relation ⌬G = –RT lnKp is dimensionally inconsistent if ⌬G has units only of J. Some authors recognized this inconsistency and resolved the problem by supplying the student with information that the “per mol” signifies that the quantities in the chemical equations are on a molar basis or that the mol–1 unit is needed to cancel the mol unit in R. What is not mentioned is that the coefficients in the stoichiometric equation are unitless and that the correct units for thermodynamic parameters for chemical reactions are intensive, that is on a per mol basis (12 a, b, c). Another oft-cited model is that as ⌬G ⬚ becomes more positive, the equilibrium constant becomes smaller. In some cases this is true—for example, if the temperature does not change—but in general this statement is misinterpreted. This particular mental model given to students is particularly difficult to erase. Consider the case where ⌬H ⬚ is positive and ⌬S ⬚ is negative. Clearly, as temperature increases ⌬G ⬚ becomes more positive. Does Keq increase according to LeChâtelier and ⌬H ⬚, or does Keq decrease due to an increasingly positive ⌬G ⬚ ? Does ⌬S ⬚ or ⌬H ⬚ determine the temperature dependence of Keq? It is of course ⌬H ⬚ and it is –⌬G ⬚ /T = ⌬Stot not ⌬G ⬚ that can also be used to make predictions about the temperature dependence of Keq. Phase diagrams, as shown in Figure 1, appear in every introductory text. Unfortunately, the drive to include everincreasing content requires a simplified description, and a misleading interpretation may be given to the phase diagram. Students may extract that at 1 atm between 0 ⬚C and 100 ⬚C only one phase for water exists, and therefore that water has no gas phase existence between 0 ⬚C and 100 ⬚C. Consider colligative properties—a topic for which it is extremely difficult to provide a student-friendly explanation. In the case of vapor pressure lowering by a nonvolatile solute, the explanation offered is often that a nonvolatile solute particle blocks the escape of solvent molecules but has no effect on the rate of return of solvent particles from the vapor to the solution. Peckham has pointed out the falsity of this explanation (13). One explanation is that if a nonvolatile solute is added to a solvent, the entropy of the solvent is increased, which lowers the solvent’s free energy, which in turn disrupts the equilibrium between solvent and vapor. Now in order to restore the equilibrium, the entropy of the vapor must in-

www.JCE.DivCHED.org



crease, which means a lowering of the pressure of the vapor, which lowers the Gibbs energy of the vapor—thus bringing the solvent and vapor back to equality of free energies, that is, back to equilibrium—impossible for students to follow and an example of “black magic”. If colligative properties are to be introduced, a phenomenological basis is best, and this should be done in the laboratory. The student is confused and finds it difficult to do anything other than memorize algorithms. No matter the overwhelming quantity of material presented, no matter how inconsistent the presentation, no matter the use of language unfamiliar to the student, it is not generally considered to be the instructor’s responsibility when the student does not understand. Van Keulen has pointed out that: “Teachers seldom find fault with themselves. What they teach is sound and valid chemistry in their own eyes. When nothing is wrong with instruction, the problem must lie with the student. Consequently, some teachers blame the students: they are lazy, lack motivation, are simply not intelligent enough, or something was wrong with their prior education” (14). In my 40 years of teaching, a few “Great Truths” about teaching have emerged. The first of these is a paraphrase of Van Keulen’s observation. Great Truth #1 Teachers seldom accept responsibility for themselves. People who have sought change in almost any aspect of a discipline have had difficulties; perhaps those advocating change in educational pedagogies have had the most ardent opposition. In part, this seems to be because we know laboratory experiments can be duplicated, but it is not widely believed that educational experiments can be. Thus there are those who refuse to accept any experiments done in an educational setting. Handelsman and colleagues have raised the question: “…why do outstanding scientists who demand rigorous proof

Figure 1. H2O phase diagram.

Vol. 83 No. 4 April 2006



Journal of Chemical Education

529

Chemical Education Today

Award Address for scientific assertions in their research continue to use and, indeed, defend on the basis of intuition alone, teaching methods that are not the most effective?” (15). The answer in part is: • Custom and tradition. We teach what we were taught in the way it was taught to us. • Cognitive dissonance. For any strongly held belief, when evidence is presented that contradicts the belief, we dig in our heels and rationalize the contradictions. • Professional definition. We fear what our colleagues will say. We know we will be accused of lowering our “standards” if we try to change.

Clevenger has pointed out: “We often accept things to be true with little or no evidence and sometimes with evidence to the contrary” (16). The less evidence we have for a position, the more faith we need to believe it. Faith is belief without proof. Should our teaching philosophy be derived from a faith-based pedagogy? Clevenger, again, has provided an important insight: “Sometimes just by acting as if something is true, we make it so…. In chemistry education do we sometimes say, ‘That sounds correct (and matches what I see and believe), therefore it’s true?’ How many of us actually check these assumptions against what is known about how students learn and what content and methods actually work? We might do this, for example, by following the results provided by our chemical education research colleagues” (16). This gives rise to another Great Truth. Great Truth #2 Teachers rationalize away evidence that contradicts their beliefs. Teaching methodology then is based on opinion, not data (17). An opinion is always a claim to knowledge. An opinion is, however, a conclusion that is justified by premises that are hypothetical; that is, based on a temporarily proposed explanation. Knowledge, on the other hand, is a conclusion that has been reached based upon testable and demonstrable data. So why do instructors rely on intuition and opinion? Is it because we have been trained in a narrow way that excluded knowledge of how students learn—a way that worked for us? The attitude seems to be: if it worked for me, why do I need to be concerned with others unlike myself? Consider that in a typical general chemistry course at best only about seven of every 100 students will major in chemistry.3 Of those chemistry majors, some will go to medical school, some to graduate studies in chemistry, and some in other directions. Of those who go on to graduate study in chemistry, some will go into industry, some will go into their own private enterprise, and some will teach. This does not leave many who, like ourselves, will pursue an academic career. Is there evidence that the traditional pedagogy of an expert lecturing and students recording and absorbing is the most effective way of producing a learning environment that benefits all students and not just those like ourselves? 530

Journal of Chemical Education



Are there emerging patterns or practices to show how to obtain a better classroom environment? The AAAS has suggested a beginning point. “Commissions, panels, and working groups have agreed that reform in science education should be founded on ‘scientific teaching’ in which teaching is approached with the same rigor as science at its best” (18). We are beginning to perceive that traditional ways have not been tested and may not be the best practices, and that, indeed, we need to learn about learning. With that perception we now begin a movement toward process. What have we learned about learning? Alberts has pointed out that knowledge about teaching is as important as is knowledge in research. “Research has taught us a great deal about effective teaching and learning in recent years, and scientists should be no more willing to fly blind in their teaching than they are in scientific research, where no new investigation is begun without an extensive examination of what is already known” (19). Mazur has cautioned that instructors may be misled by performance on exams. “[I]t is possible for students to do well on conventional problems by memorizing algorithms without understanding…it is possible for a teacher, even an experienced one, to be completely misled into thinking that students have been taught effectively” (20). Studies on the ability of students to remain focused during a class have provided vital information about student learning. Horowitz has shown that 50% of students tune out after a few minutes of a lecture and never again during the lecture is more than half of the class attentive (21). Bonwell and Eison cite several similar studies and note that “the exclusive use of the lecture in the classroom constrains students’ learning” (22). Verner, in a study on the lecture, notes that “research studies indicate that…30 minutes appear to be the optimum length” (23). These studies point out that the attention span of students during a lecture is short. The classroom should be a place where the students can become involved in handson activities. Classroom experiences that rely on rote learning and drill do not provide an opportunity for students to engage in the learning process. The movement toward active student participation in learning has been progressing as more is discovered about how students learn. Alberts noted that “teaching is a skilled profession, which can only be learned through much study and experience. This view took 20 years to acquire…we can all be expected to teach as we ourselves were taught, which explains why I only lectured at the students as a Princeton professor” (19). The physics community came to an understanding of the importance of interactive engagement (IE) some years ago. “In recent years, physicists and physics educators have realized that many students learn very little physics from traditional lectures” (24). The study by Hake of some 6500 physics students showed conceptual gains for IE students over the traditionally taught students. “The conceptual and problemsolving test results strongly suggest that the classroom use of interactive-engagement methods can increase mechanicscourse effectiveness well beyond that obtained in traditional practice” (25).

Vol. 83 No. 4 April 2006



www.JCE.DivCHED.org

Chemical Education Today

Award Address Fagan et al. came to the same conclusions: “the Peer Instruction (PI) results indicate that most of the assessed PI courses produce learning gains commensurate with interactive-engagement pedagogies, and more than 300 instructors (greater than 80%) consider their implementation of PI to be successful” (26). In addition, Halloun and Hestenes (27) have demonstrated that basic knowledge gain was essentially independent of the professor. The traditional passive introductory physics course, even those taught by the most talented and popular instructors, imparted little conceptual understanding. McDermott (28) has generalized 30 years of research ranging from physics undergraduates to graduates. Her results parallel results found by Bodner (29) for graduate students in chemistry at Purdue University. 1. Facility in solving standard quantitative problems is not an adequate criterion for functional understanding. 2. Connections among concepts, formal representations, and the real world are often lacking after traditional instruction. 3. Certain conceptual difficulties are not overcome by traditional instruction. (Advanced study may not increase understanding of basic concepts.) 4. A coherent conceptual framework is not typically an outcome of traditional instruction. 5. Growth in reasoning ability often does not result from traditional instruction. 6. Teaching by telling is an ineffective mode of instruction for most students.

These studies and others lead to the next Great Truth. Great Truth #3 Telling is not teaching. I cannot transfer an idea intact from my head to the head of a learner. Even seeing may not be sufficient for learning. Although students may see a demonstration that contradicts their beliefs, they are unlikely to recognize the disconnect. Their mental model can always accommodate any disequilibrium. These findings on how students learn have been used to develop a new approach to teaching. The Process-Oriented Guided Inquiry Learning (POGIL) pedagogy (30) is based on constructivism and the use of cooperative group learning, coupled with the three-step learning cycle (31 a, b), which consists of a hands-on exploration of data followed by a series of questions designed to guide the student to the development of a concept and finally an application of the understanding of the concept. The fundamental tenets of this approach are: • What goes on in the learner’s head is dramatically influenced by what is already there. • The instructor needs to know what the students already know and what is going on in their minds. • Students construct their own knowledge.

Such an instructional strategy provides for social interaction, an important component in the learning process, and gives www.JCE.DivCHED.org



the learner the opportunity to test new knowledge, while allowing the instructor to mentor and follow group discussions. Students work in small groups on specially designed activities that follow the learning cycle and are intended to develop mastery of both course content and key process skills. Targeted process skills are: information processing, critical thinking, problem solving, communication, teamwork, management, and assessment. Do POGIL methods work? At Franklin and Marshall College, nine years of class data for the same three instructors were analyzed to compare the POGIL approach to the traditional approach. The W, D, F rate fell from 22% to 9.6%, the lower half of the class showed the most improvement in grades, absenteeism was cut, and the withdrawal rate was 2.3% compared to 9.3% prior to implementing POGIL (30). At a regional liberal arts college with POGIL and traditional sections, with different instructors but the same texts, common exams, and randomized enrollment, the W, D, F rate for a veteran instructor in the traditionally taught organic lecture course was 33% as compared to 12% for a beginning teacher in the POGIL section; again the lower half of the class showed the largest gains.4 At a large public university the withdrawal rate in the traditional organic chemistry section taught by an experienced instructor (n = 109) was 47% compared to 12% in the POGIL section (n = 75) taught by an inexperienced instructor. For the top half of the students, the average on the final exam prepared by the traditional instructor was the same for the two sections.4 At a small liberal arts college the 1993 ACS General Chemistry final exam was administered for 10 years; the average percent correct was 55.5. The first year a POGIL class took the exam the average percent correct was 68.5 (32). At the University of South Florida, a combination of PeerLed Team Learning (PLTL) and POGIL replaced the three, one-hour lectures per week with two, one-hour lectures and one Peer-Led Guided Inquiry (PLGI) session per week (33). The instructor, texts, and exams all remained the same. On each of four exams, the PLGI students scored higher than the control group; these data are comparable to those obtained by other PLTL studies (34). The PLGI group performed better than the control group overall in spite of experiencing fewer lectures each week; thus fears were groundless that students who had less exposure to lecture would be disadvantaged. For the last 15 years it has seemed reasonable to expect that increased understanding about how students learn by cognitive scientists, educational psychologists, and classroom experiments would lead to a substantial change in the teaching of chemistry. Indeed, some changes have been made. There are some texts that now have dared to break the standard format and there are cottage industries promoting change all across the country. In response to ever-increasing content, process had been allowed to be pushed out of the classroom, but is now beginning to return. However, bringing back process does not mean ignoring the content. Although we expect change to continue to occur slowly at colleges and universities, there is evidence that it will occur more rapidly at other levels. The National Science Teachers

Vol. 83 No. 4 April 2006



Journal of Chemical Education

531

Chemical Education Today

Association (NSTA), for example, recommends that science teachers plan an inquiry-based science program and implement approaches to teaching science that cause students to question and explore. The learning cycle approach is listed as one of many effective strategies (35). The College Board is now undertaking a study of AP courses and exams in chemistry, physics, environmental science, and biology with the intention of having them reflect the principles of learning with understanding. The College Board’s science standards, curricula, and assessments will emphasize inquiry-based learning.5 In this respect the AP program is ahead of many college curricula in complying with an emphasis on conceptual understanding. The College Board has made a commitment to the following guiding principles for the AP course, among others. • Learners use what they already know to construct new understandings. • Learners have different strategies, approaches, patterns of abilities, and learning styles that are a function of the interaction between their heredity and their prior experiences. • The practices and activities in which people engage while learning shape what is learned. • Learning is enhanced through socially supported interactions.

Thus, the College Board has concluded that constructivism, previous knowledge, classroom structure, and group learning all play an important role in the learning process. If instructors are to engage in scientific teaching—paying as much attention to research in teaching as their area of specialty—they should follow several tenets supported by an overwhelming body of study (see articles cited in this work and references therein): • Active modes of learning and teaching are better for learning than are passive modes. • Students teaching students in an instructor-guided inquiry is more effective in promoting learning than are lectures. This is not to suggest that putting students in a group and asking them to solve problem number 34 from their text will result in increased conceptual understanding, but materials carefully prepared by instructors and based on current cognitive principles of how students learn will produce increased teaching gains. • A philosophical basis is needed that allows teachers to learn, to acquire new skills, and to enhance creativity. • It is easier for students to become lost after the first few minutes of class than teachers like to think. • Barriers to change are high. For new teachers it is more reassuring to follow custom and tradition than to consider innovation. For experienced teachers the pull of “this is the way I learned it” is very strong.

532

Journal of Chemical Education



• While we have models for science research, we fail to apply these models to science teaching. Without a model, teaching comes down to trial and error. • To engage students fully, knowledge of how students learn, as well as the knowledge in our own specialized content, is required (36).

In the final analysis, it is not our beliefs and opinions about teaching that should determine the pedagogy, it is what the evidence supports. Upon entering a scientific investigation, any competent scientist must to some extent be willing to accept the possibility that his or her previously held assumptions, opinions, or beliefs may be in error. That same skepticism is required when evaluating a pedagogical philosophy. It is always necessary to ask for and insist upon the evidence that leads to a particular belief. Has this new model been tested against another model and has it been shown to be more effective in promoting learning? The same questions that would be asked in any scientific investigation must also apply to teaching philosophies. Current pedagogical practices must be subjected to the same challenges and questions that are raised in chemical research. Opinion in designing a pedagogical approach is not sufficient, just as opinion is not sufficient in any research. Assumptions about teaching must be questioned. Intuition and opinion do not constitute valid arguments. When considering course content, this question needs to be asked: Why do they need to know that? (37, 38). Content is out of hand in all courses while there is little attention paid to process. The studies referenced in this article have shown that students in classes where process is emphasized do as well on traditional exams as students who were instructed traditionally; on exams more attuned to conceptual understanding, students taught traditionally do not do as well as students taught in a conceptual learning environment. Thus, attention to process and conceptual understanding will, in fact, lead to higher standards and a more rigorous course. Notes 1. This article is based on the award address for the year 2005 George C. Pimentel Award in Chemical Education, sponsored by Dow Chemical Company. The address was presented at the American Chemical Society Meeting in San Diego, CA on March 14, 2005. Information about the nominating procedures for this award (as well as a list of recipients of it) may be found on the ACS Web site at http://www.chemistry.org/awards (accessed Feb 2006). 2. These data from Educational Testing Service (ETS) were available to the author in his role as the current chair of the AP Chemistry Test Development Committee. 3. According to textbook publishers, about 300,000 plus students enroll in a general chemistry class each year. About 10,000 bachelor’s degrees in chemistry are awarded each year. (C&EN Feb. 7, 2005). At small liberal arts colleges, such as that at which the author teaches, about 7% of the students in the first-year class major in chemistry. According to a survey by Houghton-Mifflin,

Vol. 83 No. 4 April 2006



www.JCE.DivCHED.org

Chemical Education Today

Award Address about 7% of the enrolling students intend a chemistry major, but as the figures show, not all carry out their intent. Based on the author’s experience in the Council on Undergraduate Research (CUR) and as Chair of the ACS Division of Chemical Education Task Force on the General Chemistry Curriculum, the best estimate is that fewer than 7%—and perhaps as low as 3%—of students in general chemistry graduate with a bachelor’s degree in chemistry. 4. Data were obtained by the POGIL Project and may be found online at http://www.pogil.org/downloads; then choose POGIL Data Presented at POGIL Workshop at Science and Math Symposium April 9th, 2005 (accessed Feb 2006). 5. The author is currently chair of the AP Chemistry Development Committee and the AP Redesign Panel for the reworking of the AP exam and curriculum.

16. 17. 18.

19. 20. 21.

22.

Literature Cited 1. Pimentel, G. C.; McClellan, A. L. The Hydrogen Bond; Freeman: San Francisco, 1960. 2. Abraham, M. R. Inquiry and the Learning Cycle Approach. In Chemists Guide to Effective Teaching; Pienta, N. J., Cooper, M. M.,Greenbowe, T. J.; Prentice Hall Series in Educational Innovation, Prentice Hall: Upper Saddle River, NJ, 2005; pp 41–52. 3. Pimentel Report. Opportunities in Science; National Academy of Sciences: Washington, DC, 1985. 4. University Brochure, Universita Degli Studi Di Pavia, Strada Nuova 65–I 27100 Pavia, Italy. 5. Whitehouse, D. Library of Alexandria Discovered. BBC News World Edition May 12, 2004. 6. Paine, T. Common Sense; Philadelphia, 1776; Foner, P. S. Introduction. In The Complete Writings of Thomas Paine; Citadel Press: New York, 1969. 7. Cornog, J.; Colbert, J. C. J. Chem. Educ. 1924, 1, 5–12 (unnumbered folios). 8. Gillespie, R. J. J. Chem. Educ. 1991, 68, 192–194. 9. Taft, H. L. J. Chem. Educ. 1990, 67, 241–247. 10. Bodner, G. M. J. Chem. Educ. 1992, 69, 186–190. 11. Bent, H. A. J. Chem. Educ. 1985, 62, 228–230. 12. (a) Craig, N. C. J. Chem. Educ. 1987, 64, 668–669. (b) Spencer, J. N. J. Chem. Educ. 1974, 51, 577–579. (c) Spencer, J. N. The College Board: AP Central Web site is available at http:// apcentral.collegeboard.com/chemistry (accessed Feb 2006). 13. Peckham, G. D. J. Chem. Educ. 1998, 75, 787–788. 14. van Keulen, H. Making Sense: Simulation-of-Research in Organic Chemistry Education; CD-β Press: Utrecht, The Netherlands, 1995. 15. Handelsman, J.; Ebert-May, D.; Beichner, R.; Bruns, P.;

www.JCE.DivCHED.org



23. 24. 25. 26. 27. 28. 29. 30. 31.

32. 33. 34. 35. 36. 37. 38.

Chang, A.; DeHaan, R.; Gentile, J.; Lauffer, S.; Stewart, J.; Tilghman, S. M.; Wood, W. B. Science 2004, 304, 521–522. Clevenger, J. ACS Division of Chemical Education CHED Newsletter; Fall 2004. Johnstone, A. H. J. Chem. Educ. 1997, 74, 262–268. American Association for the Advancement of Science. Liberal Art of Science: Agenda for Action; AAAS: Washington, DC, 1990. National Research Council. Science Teaching Reconsidered: A Handbook; National Academy Press: Washington, DC, 1997. Mazur, E. Peer Instruction: A User’s Manual; Prentice Hall: Upper Saddle River, NJ, 1997. Horowitz, H. M. Student Response Systems: Interactivity in a Classroom Environment. In Proceedings of the 6th Annual Conference on Interactive Instruction Delivery; Warrenton, VA, 1988; 8–15. Bonwell, C. C.; Eison, J. A. Active Learning: Creating Excitement in the Classroom. ASHE-ERIC Higher Education Report 1991, 1, ED340272. Verner, Coolie. Adult Education; Center for Applied Research in Education: New York, 1964; p 78. Crouch, C. H.; Mazur, E. Am. J. Phys. 2001, 69, 970–977. Hake, R. R. Am. J. Phys. 1998, 66, 64–74. Fagen, A. P.; Crouch, C. H.; Mazur, E. The Physics Teacher 2002, 40, 206–209. Halloun, I. A.; Hestenes, D. Am. J. Phys. 1985, 53, 1043–1055. McDermott, L. C. Am. J. Phys. 2001, 69, 1127–1137. Bodner, G. M. J. Chem. Educ. 1991, 68, 385–388. Farrell, J. J.; Moog, R. S.; Spencer, J. N. J. Chem. Educ. 1999, 76, 570–574. (a) Abraham, M. R.; Renner, J. W. J. Res. Sci. Teach. 1986, 23, 121–143. (b) Lawson, A. Science Teaching and the Development of Thinking; Wadsworth: Belmont, CA, 1995. McKnight, G. Salem College, Winston-Salem, NC. Private communication. Lewis, S. E.; Lewis, J. E. J. Chem. Educ. 2005, 82, 135–139. Tien, L. T.; Roth, V.; Kampmeier, J. A. J. Res. Sci. Teach. 2002, 39, 606–632. NSTA Position Statement, October 2004. available at http:// www.NSTA.org/positionstatement&psid=43 (accessed Feb 2006) Spencer, J. N. Thought & Action, The NEA Higher Education Journal Winter 2001–2002, XVII, No. 2, 93–100. Hawkes, S. J. J. Chem. Educ. 1992, 69, 178–181. Spencer, J. N. J. Chem. Educ. 1992, 69, 182–186.

J. N. Spencer is in the Department of Chemistry, Franklin & Marshall College, P. O. Box 3003, Lancaster, PA 176043003; [email protected]

Vol. 83 No. 4 April 2006



Journal of Chemical Education

533