Teaching General Chemistry and Making a Difference - Journal of

Mar 11, 2014 - The expectations for introductory chemistry classes involve different priorities for the students and instructors. Students expect a co...
0 downloads 0 Views 133KB Size
Editorial pubs.acs.org/jchemeduc

Teaching General Chemistry and Making a Difference Norbert J. Pienta* Department of Chemistry, University of Georgia, Athens, Georgia 30602-2556, United States ABSTRACT: The expectations for introductory chemistry classes involve different priorities for the students and instructors. Students expect a course to meet their programmatic needs, to maintain their interest, and to be easily achievable. How instructors have met these expectations has probably changed over the last several decades, although instruction has not kept up with our knowledge about how students learn. A change in strategies is suggested. KEYWORDS: General Public, First-Year Undergraduate/General, Curriculum

A

general chemistry course as it was in his career. Perhaps, it is not surprising that a “nuclear-first” curriculum appeals to me: what better way to justify the need to count atomic particles and to justify the existence of neutrons to neophytes of the subject? And 40+ years ago, it was enough to keep me interested in chemistry. One does not specialize in “general chemistry” in graduate school. Organic chemistry became my interest and provided me with a dissertation topic and the opportunity for an academic career. The occasion to teach general chemistry came a short time after observing a heated discussion between two colleagues: an organic chemist blamed the general chemistry instructor because he felt that students knew nothing about acid−base chemistry when they started the organic sequence. How could that be? The best way to find out was to have the experience of teaching both. An opportunity was provided by that colleague at the University of Arkansas, Wally Cordes, another exemplary instructor of introductory courses and faculty mentor in the practical aspects of teaching. (It is with deep sadness that I report his recent passing.) Having been taught the traditional acid−base content in general chemistry, these students were still observed to be deficient in understanding in subsequent semesters when they became students of organic chemistry. Being cognizant of that problem alone apparently did not resolve the issue. Is it sufficient just to have interesting teachers in a successful general chemistry program? In fact, did your Editor learn very much in his general chemistry courses or are these just nostalgic memories of “the way it was”? Did our instructors speak about us in the same way we describe our classes: not sufficiently motivated and struggling to be successful? Readers can scan the Journal for periodic reports and commentaries about each generation of students since the first issues in 1924. One of those reports is likely about your student days, leading me to suggest that it never really was “the way it used to be”. Fortunately, we can move beyond the past and, with the knowledge about how students learn that has been gained in the last several decades, make some changes based on evidence.4 Chemistry is difficult, even the basic concepts and skills in general chemistry. Conceptual understanding and problem solving goes beyond algorithmic exercises.5,6 Students

t a recent seminar visit, an undergraduate student asked several questions about the early years of my education: why choose chemistry as a major, whether general chemistry played a significant role, and whether high school chemistry was an inspiration. Your Editor has previously admitted that an adolescent interest in combustion, particularly colorful railroad flares, had a more formative role than my high school chemistry teacher, who was perhaps even more colorful but from a completely different perspective. Chemistry seemed more interesting than cutting up formaldehyde-soaked “critters” or memorizing taxonomies, the most persistent memories retained from those limited experiences with the biological sciences. Classical physics did not seem as alluring because topics such as motion and gravity offered little mystery; your Editor might be challenged to solve one now, even after what seemed like an entire semester of parabolic motion problems, mostly in the form of airplanes bombing trains. It was not until the latter of the four required semesters that other physical phenomena appeared as questions: exactly where the electricity could be found in a power grid; why people rarely pushed with a rope; and where to buy frictionless pulleys. Student success in general chemistry is relevant to my recent administrative duties in these courses. Indeed, it is currently part of my career. However, my personal experiences in the first semester of this course sequence were not particularly inspired by the circumstances. The text was Mahan’s University Chemistry1 and recent students who complain about current textbooks might liken it to reading the phone book. (If you are young and have no experience with the latter, you can get a reasonable idea from the description on Wikipedia.2) For a voracious reader, pages filled with only text were not the problem; the book or instructor never seemed to convince me that the assigned problems were important or relevant to anything. Nonetheless, an upperclassman in my residence hall convinced me that chemistry was superior to computer science, a subject that had caught my interest. In an alternate universe, this manuscript might be about the importance of curly brackets in writing code. But after careful reflection and boosting some self-determination, the second semester of general chemistry became the next challenge. For this course, an inspired teacher, John Huizinga, made all the difference. Thus, the New York Times report of his recent passing3 is noted here with sadness. Nuclear chemistry was important in that © 2014 American Chemical Society and Division of Chemical Education, Inc.

Published: March 11, 2014 305

dx.doi.org/10.1021/ed500136y | J. Chem. Educ. 2014, 91, 305−306

Journal of Chemical Education

Editorial

need multiple exposures to ideas, and every one of those opportunities has to involve the best chances for learning; they need to be challenged, to think about the ideas, and to discuss them.7 As a community, we need to revise our strategies and practices to replace traditional lecture, laboratories that simply confirm facts, and assessments based on exercises chosen simply because the questions are easy to create.8 We can strive to make general chemistry interesting, but the next generation of students will judge us as instructors based on best practices that involve active learning and critical thinking.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

Views expressed in this editorial are those of the author and not necessarily the views of the ACS.



REFERENCES

(1) Mahan, B. H. University Chemistry; Addison-Wesley Publishing Co.: Reading, MA, 1970. (2) For the description of a “phonebook” or telephone directory, see: http://en.wikipedia.org/wiki/Telephone_directory (accessed Feb 2014). (3) For the New York Times obituary of John R. Huizinga, see http:// www.nytimes.com/2014/01/30/science/john-r-huizenga-physicist-atfore-of-nuclear-era-dies-at-92.html?_r=0 (access Feb 2014). (4) For one summary of evidenced-based practice, see the National Academy of Sciences Discipline-Based Education Research Report: http://www.nap.edu/catalog.php?record_id=13362 (accessed Feb 2014). (5) Wheatley, G. H. Problem Solving in School Mathematics, MEPS Technical Report 84.01; School Mathematics and Science Center, Purdue University: West Lafayette, IN, 1984. (6) (a) Bodner, G. M.; Domin, D. S. Mental Models: The Role of Representations in Problem Solving in Chemistry. Univ. Chem. Educ. 2000, 4 (1), 24−30. (b) Domin, D.; Bodner, G. Using Students’ Representations Constructed during Problem Solving To Infer Conceptual Understanding. J. Chem. Educ. 2012, 89 (7), 837−843. (c) Bhattacharyya, G.; Bodner, G. M. “It Gets Me to the Product”: How Students Propose Organic Mechanisms. J. Chem. Educ. 2005, 82 (9), 1402−1407. (7) For a relevant discussion, see: (a) Cooper, M. M.; Underwood, S. M.; Hilley, C. Z. Development and Validation of the Implicit Information from Lewis Structures Instrument (IILSI): Do Students Connect Structures with Properties? Chem. Educ. Res. Pract. 2012, 13 (3), 195−200. (b) Cooper, M. M.; Grove, N.; Underwood, S. M.; Klymkowsky, M. W. Lost in Lewis Structures: An Investigation of Student Difficulties in Developing Representational Competence. J. Chem. Educ. 2010, 87 (8), 869−874. (c) Lafarge, D. L.; Morge, L. M.; Méheut, M. M. A New Higher Education Curriculum in Organic Chemistry: What Questions Should Be Asked? J. Chem. Educ. 2014, 91 (2), 173−178. (8) Content maps and tools for programmatic assessment relating to undergraduate general and organic chemistry courses are available; see: (a) Holme, T.; Murphy, K. The ACS Exams Institute Undergraduate Chemistry Anchoring Concepts Content Map I: General Chemistry. J. Chem. Educ. 2012, 89 (6), 721−723. (b) Raker, J.; Holme, T.; Murphy, K. The ACS Exams Institute Undergraduate Chemistry Anchoring Concepts Content Map II: Organic Chemistry. J. Chem. Educ. 2013, 90 (11), 1443−1445.

306

dx.doi.org/10.1021/ed500136y | J. Chem. Educ. 2014, 91, 305−306