National Academies Press: OpenBook
« Previous: 2. Overview of Advanced Study Programs in Chemistry in U.S. High Schools
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

overall high school science curriculum.18 The first of these is compression of that curriculum. Although there is variation among schools, the first-year chemistry course has generally been offered during the junior year. However, schools offering a range of AP science courses have helped foster a national shift toward offering biology to ninth-grade students and chemistry to tenth-grade students so the final 2 years of high school can be used for advanced work, including AP Chemistry. Alternatively, this shift allows for making physics available in the eleventh grade, thereby offering opportunities to focus on advanced work in the twelfth grade. The panel believes that such a shift does not coincide well with the academic preparation and intellectual skills of high school freshmen and sophomores, and therefore believes that the science concepts being taught may be too abstract for the age of many students at that level.

Curriculum compression may also prevent students from being able or electing to take the full range of first-year courses in science (biology, chemistry, and physics) during their high school years. Such students may substitute AP Biology or AP Chemistry for first-year physics, thereby leaving them with significant lacunae in their quantitative understanding of physical science concepts. This practice is contrary to the College Board’s recommendation (CEEB, 1999a, p. 2). Moreover, those who take first-year chemistry as sophomores but elect not to take AP Chemistry in their junior or senior year have a 2-year hiatus between their study of high school chemistry and the time at which they enter college and major in a discipline requiring them to take general chemistry. This time lag, extended in chemistry by 1 year because of curriculum compression, can cause difficulty for some students in making the transition from high school to college chemistry.

Advanced mathematics is the gateway to advanced science, particularly advanced chemistry and physics. The College Board recommends that students taking AP Chemistry previously have completed 2 years of algebra and 1 year of geometry in addition to having taken the first-year chemistry course. To take AP Chemistry as a junior, a student must complete these mathematics prerequisites by the end of the sophomore year. Thus, these mathematics requirements also compress the mathematics curriculum. The result is that Algebra I must be completed in eighth grade. Given the small percentage of minority and rural youth who either have access to or enroll in Algebra 1 in grade 8, the need to have completed these course by the end of grade 10 in effect sets up a de facto homogenous grouping of students that continues throughout the 4 years of high school. Such tracking has serious implications with regard to which students are able to take advantage of advanced courses in chemistry and other sciences, and thereby raises issues of access and equity with regard to the benefits of college admission, credit, and placement that currently accrue to students who have such opportunities for advanced study (see Chapters 2 and 10 in the parent committee’s report for more discussion of these issues).

COURSE CONTENT

The IB curriculum, in some sense, flanks the AP course, with expectations that are either higher or lower than those for AP Chemistry. The IB SL course is at a lower level than AP with respect to the content coverage; the IB HL course is at a somewhat higher level than AP in terms

18  

The issues associated with compression of curriculum are discussed in detail in Chapter 10 of the parent committee’s report. The present discussion serves only to raise the issue as it may pertain to chemistry.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

of the topics discussed, especially in light of the HL options. The AP curriculum deals in greater depth than IB with concepts related to kinetic molecular theory, chemical kinetics, and equilibrium. In contrast, the IB curriculum at least introduces concepts from organic chemistry, an area generally not included in the AP course. The IB courses also provide students with some opportunities to develop conceptualizations of the material presented; most such opportunities are lacking in AP courses. In IB courses, chemical concepts are often taught within a context, and examination questions are designed to assess several aspects of a concept, not merely one or two. To cite one example, enthalpy changes in solution are linked to thermal changes, as well as experimental design and sources of error. Bond dissociation energies are associated with environmental changes such as stratospheric ozone depletion. In these aspects, the IB course is broader in content and outlook than its AP counterpart. A comparison of the differences between the two approaches is summarized in Table 1.

TABLE 1. Comparison of AP and IB Chemistry Courses

AP

IB

1 level

2 levels: SL or HL

1 year

1 or 2 years for SL; 2 years for HL

Less conceptualization of topics

Deeper conceptualization of topics

Molecular-level interpretation of KMT [kinetic molecular theory]

More descriptive (macroscale) approach to KMT

More quantitative treatment of kinetics and equilibrium

Less quantitative treatment of kinetics and equilibrium, but better conceptualization of the topics than in AP

No organic chemistry, biochemistry, or environmental chemistry

Organic chemistry, biochemistry, and environmental chemistry concepts covered in SL and HL programs

Consideration of phase diagrams, colligative properties, and Nernst equation

No consideration of phase diagrams, colligative properties, and Nernst equation

Minimum of 90 minutes per week recommended (not required), preferably in one session, to be spent engaged in laboratory work

Laboratory work required, making up 25% of teaching time; required Group 4 project (student-designed research investigation) makes up 10 hours of work. (The project’s intention is that students analyze a topic or problem that can be investigated in each of the science disciplines offered by a school. The topic can be set in a local, national, or international context.)

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

MATERIALS USED

A list of suggested textbooks is given in the Advanced Placement Program Course Description: Chemistry (CEEB, 1999a, 2001a). These suggested textbooks are drawn from among the textbooks that are most frequently used in introductory college chemistry courses. In many cases, especially in classes taught by inexperienced teachers, the textbook forms the basis of the AP Chemistry course. In an effort to finish the book, many of these teachers may try to cover all of the text before the May examination date. The broad scope of these textbooks contributes to an emphasis on breadth of coverage at the expense of depth of understanding.

Because of the unique structure of the IB Chemistry courses, the IBO maintains that there are no suitable textbooks that can serve as the basis for its courses. Instead, the IBO publishes and disseminates a document, Chemistry Bibliography, which lists general resources. It is divided into Pre-IB, Core/General, Core and Higher Level, Organic, and General/Organic Textbooks. It lists resources for each option and suggests sources for practical work (laboratories), comprehension exercises/review material, post-IB, and demonstrations, as well as periodicals and a list of data/useful sources of information (including Web sites). Teachers are encouraged to select the materials that best align with the common core and optional topics that they teach in their classrooms.

LABORATORY WORK

The average length of a high school class per day (42–58 minutes) poses a challenge with regard to the amount of material that can be presented and actually learned in advanced chemistry courses. Increasingly, schools have adopted block-scheduling options, allowing up to 90 minutes for class and laboratory sessions. Even with this additional time allotted to laboratory work, however, time can be a true barrier to meaningful laboratory activities, especially those that are investigative in nature rather than applying a confirmatory or validation approach to preexisting knowledge. Investigative laboratories can occupy large amounts of time, and the amount of time recommended by the College Board for more traditional laboratories may not be sufficient for these new paradigms and approaches. In addition, college and university chemistry courses typically have affiliated weekly 3- or 4-hour laboratory periods during which students can perform actual investigative experimentation that includes multiple trials, replications, and the examination and evaluation of varying methodologies. This raises questions about the alignment of students’ AP laboratory experiences with those available to college students. The panel notes that increasing numbers of high schools are working to find innovative ways of extending opportunities for laboratory work, which the panel commends.19 The panel applauds the College Board's clearly expressed guidelines that AP Chemistry courses need to

19  

One reviewer of this report pointed out that a serious curricular problem is the increasing tendency for high schools to condense the AP course into one semester through the use of block scheduling. The panel contends that, to benefit most from the AP Chemistry experience, students should take the course throughout an entire academic year. The panel views as counterproductive any attempt to complete the AP Chemistry course in one semester. The panel contends that the material in a college-level chemistry course contains too many concepts to be mastered in this truncated period of time. A one-semester course also does not allow the practice time needed to fully grasp what are for most high school students sophisticated chemical and affiliated mathematical concepts.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

have a weekly, extended laboratory period.20 The College Board explicitly states that “it is not possible to complete high-quality AP laboratory work within standard 45-to-50 minute periods” (CEEB, 1999a, p. 35) (emphasis in original). The College Board recommends that a minimum of 90 minutes weekly be spent in laboratory instruction, preferably in one session. It would be desirable to know what percentage of AP Chemistry courses follow this recommendation.

High school chemistry laboratory experiences are of mixed quality as a result of variations in resources, time available for faculty to set up and maintain laboratory experiments, and the academic background of faculty who teach these courses. When done at all, laboratory work can tend more toward verification than problem-solving investigations. The chemistry panel believes it would be desirable to avoid the former exercises and instead emphasize inquiry-based experiments. Moreover, teachers and administrators would likely pay more attention and commit more time and resources to enhancing laboratory experiences if the culminating AP examination stressed and tested for the knowledge, skills, and techniques that are gained primarily, if not exclusively, through laboratory experiences.

The IB examination has more laboratory-based questions than its AP counterpart, but still relatively few. Unlike the AP program, however, IB includes a student’s laboratory grade as a component of the final course grade. It should be noted that the IB program also requires students to prepare portfolios in which they demonstrate that they have the ability to plan, design, and analyze scientific experiments. While these components of the portfolio may draw on experiences in addition to those from laboratory, they contribute to the student’s development of skills and understanding of scientific procedures provided by laboratory experiences. Finally, the IB program requires teachers to submit a detailed description of the PSOW completed by all students, as well as examples of work from each individual student. Guidelines for assessing laboratory work are detailed extensively in the IB Diploma Programme Guides.

The chemistry panel notes that in 1999, the AP examination introduced a required laboratory-based question in the free-response section (CEEB, 1999b). Should this question continue to be structured so as to assess students’ understanding of laboratory techniques and data acquisition, this will be a positive step. The panel believes, however, that all of the subparts of the laboratory-based question should be directed to laboratory techniques, experimental design, and data acquisition and interpretation, rather than to theoretical constructs. Because the AP examination contains only one question about laboratory work, including experimental techniques, an AP student could score well enough on the other parts of the examination to earn college credit without ever having undertaken any of the suggested laboratory work.

The chemistry panel heard anecdotal accounts of AP Chemistry teachers who omit or defer laboratory activities to review previous AP examinations. In such circumstances, laboratory work, if done at all, is crammed into the course after students have taken the AP examination. This deplorable practice appears to result from teachers recognizing and taking advantage of the lack of laboratory questions on the examination. Such practices disregard the validity of laboratory work as a means of introducing, reinforcing, consolidating, and amplifying chemical principles. The fact that a number of colleges and universities require proof of sufficient chemistry laboratory work before granting credit or placement out of the general

20  

The College Board recommends but does not require that students participate in particular types of laboratory experiences.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

chemistry course indicates that laboratory work is a weak link in the AP course because of the variability in the way the laboratory component is administered.

In contrast, the IBO requires that teachers assess student work, with samples being sent to the IBO for moderation. The IBO allows only 2–3 hours of work to be carried out after the deadline, which prevents teachers from leaving laboratory work until the end of the course. Further, the IBO provides detailed guidelines and rubrics for assessing practical work (IBO, 2001, pp. 15-32).

CURRICULUM AND METHODS OF INSTRUCTION

Given the broad scope of the AP and IB chemistry curricula, it can logically be inferred that large components of AP and IB Chemistry courses may be taught by the traditional method of “filling the open vessel”—lecture, note taking, and assessment by recall. These practices are in stark contrast with the constructivist model of learning recommended by the National Science Education Standards (NSES) (NRC, 1996) and other recent reports (NRC, 2000b). Those studies clearly demonstrate that students learn more deeply when they develop an understanding of the material while undertaking inquiry-based, problem-centered activities. Based on the materials examined, the chemistry panel agrees that neither AP nor IB appears to emphasize such approaches to learning. Rather, the breadth of topics included in AP and IB course outlines and syllabi indicates that far too much emphasis is being placed on covering a large body of information deemed necessary for success on the examination.21

The kinds and levels of questions that appear on both the AP and IB examinations reinforce the emphasis on broad but shallow coverage of topics. Thus, an overarching, largely unintended, but nevertheless real and perverse effect is that the exam-driven nature of both programs may cause the development of intellectual curiosity in students to fall victim to the pace of the courses—all in the name of “rigor.”

The chemistry panel also is concerned that the current system of basing the AP Chemistry course on typical or average general chemistry courses (using information gathered through surveys of chemistry departments at colleges and universities that receive the greatest number of AP candidates [see, e.g., CEEB, 1999a]) precludes incorporating emerging best practices that are beginning to appear in some college-level chemistry courses. Further, because AP courses are based on average college general chemistry courses, these changes will not be reflected in AP courses until changes in college-level chemistry become widespread,.

A far better model would be to base AP Chemistry courses on best practices in the teaching of college chemistry, even if the resulting course were less similar to typical college courses than is now the case. The panel notes that neither the AP nor IB Chemistry course as yet accurately reflects recent efforts to restructure and change teaching and learning in general chemistry at the college/university level. Among such changes are emphases on “less is more” in terms of course coverage, a wider variety of assessment techniques, small-group and inquiry-based learning, and inquiry-based laboratories (National Research Council, 1999). The AP and IB

21  

The panel notes that the IB’s IA component is not taught in a traditional manner and is consistent with the recommendations and emphases for teaching and learning in the NSES.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

Chemistry courses also do not yet recognize the increasingly interdisciplinary nature of modern chemistry; its incorporation of highly important related fields, such as materials science and biochemistry; and the opportunities presented by such fields to teach related chemical concepts in a contextual manner. It is important to note that the College Board established the Commission on the Future of the Advanced Placement Program to make recommendations regarding the future of the AP program. The commission recommends in its recent report, Access to Excellence: A Report of the Commission on the Future of the Advanced Placement Program (Commission on the Future of the Advanced Placement Program [CFAPP], 2001), that leaders in the subject disciplines, pedagogy, and research be engaged to ensure that current reforms and best practices are reflected in AP examinations

QUALITY CONTROL

To be successful, an AP Chemistry course should be initiated only after a school’s administrators and faculty have carefully considered the valid reasons for offering such a course. Such consideration should be followed by a thorough analysis of the resources—personnel, facilities, and supplies—available at the school to support the course. While these considerations would appear obvious, high schools are not bound by them (CEEB, 1999a). A school can offer an AP Chemistry course by administrative fiat, even when the personnel, facilities, and supplies available for the purpose are not up to the expectations of the College Board, as noted in the Acorn Book for chemistry. It is not unreasonable to expect the College Board to exercise some control over where AP Chemistry courses are offered and who teaches them. It is therefore encouraging to note that Access to Excellence, completed after the chemistry panel had conducted its deliberations, contains the recommendation to develop and implement standards for AP programs in schools and school systems, for AP teachers, and for professional development workshops and institutes for AP teachers. Access to Excellence also recommends that the College Board take a more proactive approach to leading educational reform by changing the emphasis of the AP program from one of replicating typical college courses to one of emulating best practices in the discipline. The report further recommends that leaders in the subject disciplines, pedagogy, and research be engaged to ensure that current reforms are reflected in AP examinations. The chemistry panel endorses these recommendations fully and is pleased to learn that the College Board is willing to serve as a forum and vehicle for stimulating educational reform.

The College Board’s expectations for teacher qualifications are explicit: “if the objectives of a college-level general chemistry course are to be achieved, the teaching should be done by a teacher who has completed an undergraduate major program in chemistry including at least a year’s work in physical chemistry” (CEEB, 1999a, p. 3). However, the College Board does not have a means of verifying that AP Chemistry teachers have these minimum qualifications or certifying them as competent to teach AP-level courses and associated laboratory activities. This lack of oversight and control is in addition to the College Board’s lack of a mechanism for determining whether a school has adequate facilities and supplies before allowing it to offer an AP Chemistry course. Nor does a mechanism exist for making such a determination once an AP Chemistry course is being taught, or for preventing a school from continuing to offer the course if the school is found lacking until the shortcomings are addressed. The Commission on the Future of the Advanced Placement Program (2001) does not recommend

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

certification of schools or teachers, but recommends instead that the College Board develop and disseminate AP quality standards for teachers, schools, and school systems. The commission (2001) also recommends that the College Board undertake a rigorous and systematic program of research to validate that (1) AP examinations actually test what is covered in the corresponding college courses and (2) students who earn specific scores on AP examinations have indeed mastered subject matter at a level equivalent to that of students who take these courses in college.

In contrast, before a school is authorized to offer an IB Chemistry course, the school must be authorized to offer the complete IB Diploma Programme. One factor in this authorization process is an evaluation of the qualifications of the people who will teach the individual IB courses. In the sciences, a school that is seeking authorization also must demonstrate a plan for an acceptable set of laboratory activities relative to its resources. Sample laboratory reports must be sent periodically to IBO headquarters for critical review (IBO, 1999). It should be noted, however, that IB program administrators provide minimal oversight to ensure that the program standards are maintained over time. However, assistant examiners who encounter schools experiencing difficulties with the program try to inform the IBO of these situations. Further, in the sciences, ongoing evaluation of a school’s program continues through the IBO’s moderation of sample laboratory reports and teachers’ PSOWs that are submitted annually. Feedback is provided to the schools and teachers to promote improvements.

Students cannot take the IB examination without having taken the course. In contrast, students can register to take the AP Chemistry examination without ever having taken an AP course, although this is not typically the case. The obverse is also true: students who take an AP Chemistry course are not required by the College Board to take the examination. The College Board tracks only those students who take the examination. Although it does not have precise data about the number of students enrolled in AP courses, the College Board estimates that approximately 63 percent of students who are taking a course designated as AP sit for the AP examination in that subject. This, however, is the estimate for AP courses in toto. Currently, data are not available from the College Board for individual subject areas, such as chemistry. Consequently, the percentage of those students taking an AP Chemistry course who do not take the AP exam is not known.22 Accordingly, the panel believes that statements by the College Board about the quality of AP Chemistry courses are suspect because they fail to account for the many AP courses nationwide in which large numbers of students may not take the examination. In addition, the AP examination results are not available until after the end of the academic year. Thus the final grade received in an AP Chemistry course is not specifically linked with the student’s performance on the AP examination. Furthermore, since AP examination scores are not available until students have completed the course, there can be no grade-related consequence in the course from not taking the AP examination or, if taken, from doing poorly on the examination.

The panel also notes that commercial vendors, some of them linked with a university (e.g., the Michigan State University program), now offer AP courses on the Internet to individual students. This new development allows students who pay the necessary fees to take an AP course independently of whether their school offers an AP course in the same subject. Internet-based courses are self-pacing. The pace and calendar are established by the examination date,

22  

The College Board recently began to gather such data from school AP coordinators.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

should the student choose to take the examination. In courses with a laboratory component, such as chemistry, arrangements for laboratory work may be irregular. In some situations, the laboratory work is done at a nearby campus or high school during weekends or other periods when several laboratory experiments are conducted in an intense burst of work. In these circumstances, the timing of laboratory activities may not correlate well with the chemical principles being studied at that point. Simulations of laboratory experiments are also used, although current simulations offer no opportunity for students to gain facility and dexterity in the proper manipulation and use of laboratory equipment.

COURSE CONTENT AND EXAMINATIONS

The 75 multiple-choice questions on the AP Chemistry examination (Section I) match closely the objectives stated in the Acorn Book for an AP Chemistry course. Despite the rather long list of chemical topics and concepts to be included in the AP curriculum, Section II of the examination asks questions evaluating a more limited set of topics. In addition, the panel finds that the topics tested in both parts of the examination are rather predictable from year to year and formulaic in the way in which the concepts are queried,23 especially in Section II. Certain topics and the style of questions addressing them change little from year to year. One example is the standard question in Section II, Part B of the AP examination that is related to reactions and writing of their affiliated chemical equations. (Although this question could perhaps be described as a laboratory-based question, the panel believes that students can answer the question correctly by sheer rote memory without ever having done any of the laboratory work.) Given this type of predictability in the examination’s coverage from year to year, together with teachers who are very familiar with the general content and structure of the test, students may be able to earn high scores on the examination without actually having mastered all the material. This is especially true when students are given a choice of which questions to answer in Section II.

Some of the IB questions are identical in content to the chemical systems described and covered in the syllabus, thus requiring only rote learning with little mastery or understanding. In fact, there is an injunction from the IBO to IB instructors against using chemical systems other than those on the examination when teaching a given topic.

AP examinations ask less than IB exams about the application of concepts, especially with respect to new contexts or chemical systems. There is heavy emphasis on algorithmic solutions, rather than the extension or application of concepts to new, unfamiliar but equivalent situations. Few of the questions test students’ abilities to predict or explain observations. For example, students could successfully write and balance sets of chemical equations with little understanding of the principles of chemical reactivity involved and no knowledge of the associated phenomena. The tests assess primarily the acquisition of information, as opposed to analyzing students’ understanding, application, and extension of concepts. The examinations thus do little to encourage inquiry-based learning.

The chemistry panel drafted samples that demonstrate how questions taken from the 1999 AP Chemistry examination could be modified to accomplish several objectives: requiring

23  

In large part, this level of predictability is based on the Educational Testing Service’s emphasis on and close monitoring of exams to maximize their reliability from year to year.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

critical thinking, combining qualitative and quantitative aspects of a chemical system within a given question, and applying concepts to chemical systems and situations not previously encountered. The suggested modified questions are presented in Appendix D. By using the approaches illustrated by these types of questions, the AP and IB programs would encourage teachers to teach in less algorithmic ways and students to learn in a different, more inquiry-based manner as recommended by the NSES.

The AP examinations have yet to address the shift in increasing numbers of introductory college/university chemistry courses toward including applications in biochemistry, materials science, or environmental chemistry. The IB program and examinations include some applications in biochemistry and environmental chemistry, but still lack attention to materials science.

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

4
Teachers and Teaching

TEACHER PREPARATION AND PROGRAM QUALITY

It should be self-evident that to teach a subject well requires at least a solid academic background in and working knowledge of the subject matter being taught. In addition to knowing the subject matter, however, a skilled chemistry teacher has the pedagogical insights necessary to present chemical concepts most effectively to different students. Clearly, no less should be expected of those who teach AP or IB Chemistry. To provide a chemistry course in line with the criteria for an advanced study course in chemistry at the high school level noted in Chapter 2, those who teach such a course must be adequately prepared. The chemistry panel takes this minimum level of preparation to mean a B.S. or B.A. degree in chemistry that includes a full year of physical chemistry. This view of the panel is congruent with that of the College Board, which recommends that an AP Chemistry teacher have completed an undergraduate major in chemistry, including a year’s work in physical chemistry (CEEB, 1999a, 2001a). This expected level of preparation also is consistent with recommendations from the National Science Teachers Association (1998) and the finding of the National Commission on Mathematics and Science Teaching for the 21st Century (NCMST, also known as the Glenn Commission) that “the most consistent and powerful predictors of higher student achievement in mathematics and science are (a) full certification of the teacher and (b) a college major in the field being taught” (2001, p. 18). The members of the chemistry panel would extend this recommendation to state that high school teachers who offer advanced courses in chemistry should also have earned an MA. or M.S. degree in chemistry, although the panel recognizes that realizing this goal would be difficult given the current and expected future shortages of science teachers.

The reality is, however, that there is enormous variability in the subject matter preparation and backgrounds of chemistry teachers in the United States. There are well-prepared and experienced teachers who teach AP Chemistry at the level expected by the College Board. Yet despite growing demands for AP and IB Chemistry courses in more schools and in more states, the pool of teachers actually qualified to teach the courses is small (Blank, 1992). As the number of schools desiring to offer these courses increases while the pool of qualified teachers fails to keep pace, a point will be reached at which the ability of many schools to offer quality AP or IB Chemistry courses will be seriously compromised. If such is the case, the very credibility of many AP and IB Chemistry courses may be called into question.

Within the past decade, a significant effort has been made nationally to change the way in which chemistry and other science teachers are certified (National Research Council, 1997). A

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

shift in an increasing number of states has led to the requirement that teachers have either a college major or a defined number of designated academic credits in the subject matter of the field to be taught (National Science Board, 1998). Currently, 29 states require that high school teachers have a degree in the subject matter they teach, rather than an education degree. As of 1993, 63 percent of all grade 9–12 teachers of science had a major in science, while nearly three-quarters (72 percent) had a major in science or science education. Although encouraging, these data fail to take into account the fact that two-thirds of all states permit a “broad-field” secondary science certification, which certifies teachers to teach across a range of science subjects (chemistry, physics, biology, and earth or space science). In at least eight states, the broad-field certification permitting teachers to teach more than one science encompasses the same number of semester credit hours as the states’ single-science certification (Blank, 1992). Of the nearly 65 percent of those with a major in science who are teaching science in grades 9–12, almost one-third are doing so under a broad-field secondary science certification. More than half of teachers teaching physical science classes (chemistry, physics, earth science, or space science) do not have an academic major or minor in any one of the physical sciences (NRC, 2000a, pp. 50-51). Additionally, data from the Council of Chief State School Officers (1997) indicate that approximately half (53 percent) of those teaching chemistry do so as their primary teaching assignment.

These data raise serious questions concerning the adequacy of academic background preparation for many of those who teach even the first-year course in chemistry, let alone advanced courses in the subject. Additional research is required to document the relationship between teaching effectiveness in introductory courses and student performance in advanced courses. However, it appears logical that students in first-year courses who are taught by underprepared chemistry teachers are more likely to lack the foundational concepts and skills needed to complete advanced study work successfully, as compared with students whose teachers have mastered content knowledge and effective pedagogical approaches to teaching chemistry. The effect is compounded if the same teachers who teach the first-year courses inadequately are assigned to teach the advanced study chemistry course as well.

TEACHER SELECTION, DEVELOPMENT, AND SUPPORT

Many high schools have only one chemistry teacher. As noted, this individual also is likely to teach other science courses (CCSSO, 1997). For an AP Chemistry teacher in this situation, the AP assignment increases the number of required course preparations as well as the number of different laboratory preparations, along with their attendant difficulties. If this teacher also teaches nonchemistry courses (biology or physics, for example), he or she must remain informed about developments in those subject areas as well. Perforce, this reduces the amount of time the teacher has to remain abreast of developments in chemistry and thus adequately prepared for the AP Chemistry course.

As noted earlier, the chemistry panel believes strongly that an AP or IB Chemistry course cannot be taught appropriately by an instructor whose background in the content of the subject is insufficient. For example, it is the panel’s position that a B.S. in science education does not adequately prepare an individual to accept an AP or IB Chemistry assignment without completing further coursework in chemistry. This view, too, is consonant with the

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×

recommendation of the College Board that a teacher of AP Chemistry have an undergraduate major in chemistry.

Data are unavailable about the exact qualifications and academic preparation of current AP or IB Chemistry teachers. The way in which those who teach AP or IB Chemistry are selected or assigned to do so also appears to vary considerably among schools. First-year teachers, veteran teachers, and those whose experience levels fall anywhere between these extremes may be given this assignment. In many school districts, seniority appears to play a significant role in the selection—even when the individual selected through seniority is not as well prepared as less senior colleagues.

PROFESSIONAL DEVELOPMENT

The College Board offers a variety of professional development opportunities to support AP Chemistry teachers. Such support includes summer institutes, workshops, and seminars. New AP Chemistry teachers are invited to attend 1- or 2-week summer institutes to learn the rudiments of teaching an AP course, its laboratory component, and the College Board’s expectations for the course. There are also 1-day seminars held regionally to update AP Chemistry teachers on teaching developments and changes in the AP courses or examinations. In addition, teachers from schools that participate in AP courses are invited to videoconferences where they can learn from those who develop AP examinations and review AP courses. And an online discussion group is available for AP Chemistry teachers to share syllabi, compare teaching strategies, and discuss substantive issues related to the content of the course or chemistry in general. Frequently, college instructors participate in these discussions, providing a forum for discussions between high school and college faculty around the teaching of chemistry. Participation in the regional workshops and online discussion groups is free.

AP summer institutes are not free. The College Board supports some of them, and schools sometimes bear the costs for their teachers to attend, but not always. Absent such support, the length of the institutes (1 or 2 weeks) and their associated costs (travel, registration, meals, housing), if not covered by a school, district, or state, are a disincentive for teachers to attend, particularly when such attendance is voluntary and outside of paid contract hours. The structure and format of these institutes vary. Most offer sessions on pedagogy and an orientation to the AP Chemistry course, its laboratory, and the AP examination. However, the institutes do not address the major shortcoming of some AP teachers noted above—an inadequate background and the lack of a deep understanding of the chemical principles necessary to teach the AP course material successfully. As noted above, a deep conceptual understanding of the content and unifying concepts of chemistry is a critical requirement for effective teaching in the discipline.

The panel encourages the College Board to provide a substantial and sustaining level of guidance and oversight for the preparation of teachers, student learning, and support by school systems so that high school AP Chemistry courses can be of the high quality espoused by the College Board. Such efforts might include expanded and enhanced support for teachers, the establishment of creative partnerships for teacher professional development, and the development of comprehensive services for schools and school systems to help them design and implement

Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 11
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 12
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 13
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 14
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 15
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 16
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 17
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 18
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 19
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 20
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 21
Suggested Citation:"3. Quality and Content of the Learning Experience for Students." National Research Council. 2002. Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry. Washington, DC: The National Academies Press. doi: 10.17226/10364.
×
Page 22
Next: 4. Teachers and Teaching »
Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry Get This Book
×
 Learning and Understanding: Improving Advanced Study of Mathematics and Science in U.S. High Schools: Report of the Content Panel for Chemistry
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

READ FREE ONLINE

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!