Effects of Context-Based Laboratory Experiments on Attitudes of

Jan 1, 1999 - activities in health science courses has led to more positive student attitudes about the activities themselves (11) and to the course i...
1 downloads 11 Views 96KB Size
Research: Science and Education edited by

Chemical Education Research

Diane M. Bunce The Catholic University of America Washington, D.C. 20064

Effects of Context-Based Laboratory Experiments on Attitudes of Analytical Chemistry Students J. Henderleiter* Department of Chemistry, Grand Valley State University, Allendale, MI 49401 D. L. Pringle Department of Chemistry and Biochemistry, University of Northern Colorado, Greeley, CO 80639

The laboratory portion of many analytical chemistry courses introduces students to some aspects of classical chemical analysis. Considerable time is spent practicing skills such as calibration, titrimetry, colorimetric analysis, weighing, pipetting, and sample preparation. While these skills are important to many chemists, they do not represent all that chemists do. A chemist must also be able to communicate with other chemists and nonchemists, work as a member of a team, evaluate data, and make decisions and recommendations based on the data collected (1). These areas are often neglected in introductory courses in favor of practicing physical techniques and calculations, and students are often graded solely on how closely their individually determined results approach an “accepted” value. This fosters the attitude that analytical chemistry is concerned mostly with getting the “right” answer while working alone on an experiment. The focus of this study is on students’ attitudes toward analytical chemistry laboratory and the factors that influence these attitudes. Attitudes are important because they have an impact on what is learned and how it can be used (2–5). We used qualitative methods to examine the effect of contextbased laboratory activities on students’ attitudes toward the perceived usefulness and importance of techniques and skills learned in the laboratory. We define attitude, in this study, as the opinions and beliefs concerning the perceived usefulness and applicability of a subject, method, or technique. This definition is consistent with others; Shrigley (6 ) suggests that attitude is related to “information acquisition”, and thus attitudes are based on the knowledge and beliefs of the learner. The relevance of science is also a component of attitude. Edeani (3), Levin and Klindienst (7 ), Hawkins and Pea (8), and Yager (9) suggest that the perceived usefulness of science is linked to positive student attitudes toward science. Chisholm (10) found this correlation in mathematics. Similar results have also been found in allied health fields, particularly nursing (11–13), and in geological sciences (14, 15). Attitude-change strategies can be introduced into the laboratory in many ways. One is to incorporate “real-world” applications into the laboratory portion of science courses. Students can make judgments about data and collection procedures as would happen in business, industry, and research *Corresponding author.

100

settings. Case studies and role-playing also bring relevance to science laboratory activities. Case studies, used in many health science courses, bring the patient’s perspective into the activities (11, 12). The incorporation of applications into laboratory activities in health science courses has led to more positive student attitudes about the activities themselves (11) and to the course in general (12). Neman (16 ) and Walters (4) describe analytical chemistry laboratories that use teamwork, report writing, data analysis, and written and oral communication skills in addition to introducing chemical analysis techniques. Walters (4) reports on an analytical laboratory course designed around role-playing activities that teach not only analytical chemistry techniques but also cooperation, responsibility, and application of principles. The factors that contribute to attitude change and how to assess such changes are controversial. Much research has focused on instruction and attitude, the development and validation of attitude scales, and cause–effect relationships between attitude and other variables, including beliefs (17–19). However, clearly defined relationships among these variables are not known, and questions surrounding the validity of many attitude measures abound (17, 18, 20). To this end, both qualitative and quantitative methods are used in this study. It is expected that data collected through observation and interviews will enhance that collected by survey. Procedure

Laboratory Activities Modifications were made to seven of nine laboratory activities from the first-semester analytical chemistry course at the University of Northern Colorado, and one new activity was written. The students in this course are mostly juniors— with some sophomores and seniors—majoring in chemistry, biology, or medical technology. Modified laboratories contained a “script” or “real-life” scenario designed to provide a rationale for learning the techniques and methods introduced in the laboratories. The modifications took into consideration research linking attitude with case studies and the results of a survey completed by eastern Colorado industries that employ chemists. The survey consisted of five questions. The first was a checklist of common skills and techniques used in a laboratory and often taught in analytical chemistry (e.g., titration, use of a pipet, calculations). The second asked for

Journal of Chemical Education • Vol. 76 No. 1 January 1999 • JChemEd.chem.wisc.edu

Research: Science & Education

types of materials analyzed by techniques from the checklist. Questions three through five asked for comments on what knowledge, skills, and techniques recent college graduates performed satisfactorily, which ones needed improvement, and what suggestions the company had for additional analyses or procedures that could be incorporated into the classroom. To illustrate the changes made, three of the original unmodified laboratory experiments and their modified counterparts are described in Table 1. The new laboratory was a determination of total dissolved solids in a community water sample. The redesigned activities were built around a context, or reason why an analysis of this nature might be performed. Questions were added to the report sections, challenging students to interpret their data at a level beyond mere statistical interpretation. Students were asked to make a decision about the scenario described in the beginning of the laboratory and to use their results to rationalize a solution to the problem. Some of the modified experiments specifically required students to share data, and in one case to work in a group to solve a problem. However, formal semester-long cooperative groups were not formed in the laboratory. The expectation for both the control and experimental groups was that each student would do his or her own work, as directed by the experiment, but could discuss results, procedures, and questions with peers and the instructor.

Laboratory Instructors The laboratory instructors were asked to teach their classes as they saw fit. Two different instructors for the two laboratory classes were included in the control group. One of them also taught both laboratory classes included in the experimental group, and he was asked to teach the experimental and control classes in as similar a manner as possible. The time spent on prelaboratory discussion was 30–45 minutes in the control and experimental classes taught by this instructor. The second instructor spent 20–30 minutes on prelaboratory discussion in the control class. The material covered in prelaboratory discussions described the mechanics of performing the experiment and the chemistry involved. The type of assistance the instructor gave students was similar in the control and experimental classes. The instructor responded to questions by probing for more information. When the question concerned the next step or a reason for

performing a step, the instructor asked questions designed to help the students reason out their own answer.

Attitude Surveys, Observations, and Interviews Student attitudes about the practical relevance of skills, methods, and techniques learned in an analytical chemistry laboratory were surveyed using a Likert-style survey developed at the University of Northern Colorado. It is based on a similar survey designed by Hofstein, Ben-Zvi, and Samuel (21), which examines Israeli high school students’ attitudes toward chemistry. It comprises 24 statements related to students’ perceptions of analytical chemistry, its applications, and its perceived usefulness to themselves and others. Statements were rated on a scale of 5 to 1, a score of 5 being assigned to the response expected of students who have a positive attitude toward chemistry. The survey was administered to the control and experimental groups at the beginning (pretest) and again at the end (posttest) of the semester. From the control group, 21 students completed the pretest survey; of these, 16 also completed the posttest. From the experimental group, 21 students completed the pretest survey, and 19 also completed the posttest. A sample of the statements to be rated is listed below; the scoring scale is shown (SA = strongly agree). Analytical chemists are very precise in their work. (SA = 5) A good analytical chemist has no errors associated with his or her work. (SA = 1) I could use skills learned in analytical chemistry in the career I’d like to pursue. (SA = 5) Analytical chemistry is boring. (SA = 1) Work in a chemistry laboratory is an interesting way to earn a living. (SA = 5) It is not usually clear what the purpose of performing a laboratory experiment is. (SA = 1)

Students were also observed as they worked in the laboratory. A written record of how much and what types of interaction took place among students and between students and the instructor was kept. Examples illustrating the types of questions asked of peers and the instructor were specifically noted. At the end of the semester, students were interviewed to assess their thoughts and attitudes about their laboratory experiences. Four of 21 students from the control class (3 men,

Table 1. Major Changes Made to Laborator y Experiments Original Experiment

Modified Experiment

Calibration of Glassware (completed in one 3-hour lab period) Students calibrated a glass pipet and a buret, calculated tolerances on their equipment.

Students calibrated and calculated tolerances the same equipment, plus an automatic pipet. Students pooled data for autopipet and compared results with manufacturer's claims.

Gravimetric Analysis of Calcium (completed in two 3-hour lab periods) Students performed a gravimetric analysis of calcium carbonate; precipitated calcium as calcium oxalate. Several samples analyzed to measure technique. Students calculated the mean, average deviation, and relative deviation for their samples.

Students worked in groups to decide how to dissolve a calcium supplement. Students were provided with a CRC Handbook and the supplement bottle. Students dissolved a tablet and analyzed several aliquots from one tablet (by precipitation of calcium oxalate). They compared data with other students who analyzed the same brand of tablet. Calculations were the same as in the original experiment. Students commented on the appropriateness of the method, assuming that "governmental guidelines" required the tablets be within 5% of the bottle label and within 3% based on a (hypothetical) employer's quality control requirement.

Iodometric Titration of Copper in Ore (completed in three 3-hour lab periods) Students performed an iodometric titration for copper in a copper ore sample. Students calculated the percent copper, mean, and average and relative deviations from the mean.

Students performed the titration and calculations described in original. The rationale for performing the experiment was to determine if the ore sample came from a mine worth developing. Students found the price of copper and were asked to calculate the worth of the ore in a mine of a given size and, assuming the copper ore had their percentage of copper, justify why the mine was or wasn't worth developing.

JChemEd.chem.wisc.edu • Vol. 76 No. 1 January 1999 • Journal of Chemical Education

101

Research: Science and Education

1 woman) agreed to be interviewed, and 10 of 34 (6 men, 4 women) from the experimental class were interviewed. These 10–20-minute exit interviews were conducted one-on-one by the researcher, who was known to the students as a chemistry graduate student. The students were informed by both the researcher and the course instructor that the researcher had no input into their grade. They were also informed that the instructor would not have access to the audiotapes or interview transcripts until after grades had been assigned. At the beginning of each interview, the interviewer told the student that a series of questions would be asked, and that the questions were the same for all interviewees. The students knew that the purpose of the interview was to gain insight into their opinions and perspectives on the analytical chemistry laboratory. During the interviews, some students were reassured that their answers were fine—that there were no “right” answers to the questions. The interviews were transcribed and coded along four themes related to attitude: Perception of self as a scientist/chemist; Attitudes toward the laboratory; Expectations about the laboratory; and Beliefs about what is important to know and be able to do as a scientist

Each theme was further divided into categories, or subthemes, showing the range of statements under the main theme. Results Students were compared in several ways to estimate the similarity of the classes. Two analytical chemistry laboratory classes were taught each semester; the two first-semester classes made up the control group, and the two second-semester classes made up the experimental group. Grade point averages of students from each semester’s classes were computed. The total number of chemistry, biology, medical technology, and “other” majors and the gender mix in each class were also determined. The grade point average and gender mix of the two classes were similar (Table 2). The distribution of majors, however, differed. This may be reflected in how students perceive the usefulness of analytical laboratory, since their personal goals may differ. Student majors cannot be linked to survey data; however, interview data do not differ in any significant manner for chemistry, biology, and medical technology majors.

Attitude Surveys Data from the pretest and posttest attitude surveys were analyzed using a two-tailed t-test (α = .05) (22). A few significant differences on individual items were found at the .05 level. For example, on the pretest, more students from the experimental classes than from the control classes disagreed with the statement “the laboratory work I will do in analytical chemistry does not reflect what analytical chemists actually do”. Also on the pretest, more students from the experimental classes agreed strongly with the statement “I can use skills learned in analytical chemistry in other laboratory settings”, although on the posttest the same students disagreed slightly more with this statement. On the posttest, more students from the control classes than from the experimental classes agreed strongly with the statement “I practice proper laboratory techniques in the laboratory”. Finally, the posttest experimental classes were more neutral toward the statement “analytical chemistry in-

102

volves a lot of work” than were the posttest control classes, who tended to agree with the statement. Despite these small differences on individual items, the overall results of pretests and posttests for the control and experimental classes are not significantly different.

Observations during Laboratory Classes Observations during the first few weeks of the course suggested that there were no distinct groups of friends in the class at the beginning; at least, none were noted in the laboratory. Each of the two control classes was taught by a different instructor (A and B). Both experimental classes were taught by instructor A. All four classes were fairly quiet at the beginning of the semester; there was little interaction among students and their peers, although interactions increased as the semester progressed. One difference between the two control classes involved the role of the instructor. Instructor A was nearly always present in the classroom. Students asked him questions throughout the laboratory, although direct answers were not always forthcoming. Some questions were answered with a series of probing questions that elicited the answer from the student. For example, when students asked about the endpoint of a titration, they were often told, “Well, record the volume and add another drop and see what happens.” Instructor B had a less structured class. He was not in the classroom as often as instructor A. When he felt that students should know the answer to their question, he’d often shrug and ask them what they thought. He would then nod to their responses or ask them questions like, “Really?” He answered questions directly when he felt the student really did not know an answer or needed information. His laboratory introductions and lectures were shorter, generally 20–30 minutes (versus 30–45 minutes for instructor A), possibly because he had less experience teaching this particular course. These differences in teaching styles led to some differences between the classes. Instructor B’s control class had slightly more peer interactions inside the laboratory than did instructor A’s control class. The increased interactions were mostly among peers who were seen working together on chemistry problems outside of the laboratory. Students in the experimental classes (both taught by instructor A) were fairly passive during the prelaboratory lectures; they took notes and asked few questions. Except for the new laboratory experiments, instructor A taught the control and experimental classes in essentially the same manner. The biggest difference was that for some activities, he asked more leading or guiding questions in the experimental class. Table 2. Class Characteristics for Analytical Chemistry Group Characteristic

Control (23 students)

Experimental (36 students)

Mean GPA ± SD

3.06 ± 0.72

3.02 ± 0.54

11 (48%) 4 (17%) 5 (22%) 3 (13%)

25 (69%) 5 (14%) 5 (14%) 1 (3%)

12 (54%) 11 (48%)

21 (58% ) 15 (42%)

Major Chemistry Biology Medical technology Other (graduate, chemistry minor) Gender Male Female

Journal of Chemical Education • Vol. 76 No. 1 January 1999 • JChemEd.chem.wisc.edu

Research: Science & Education

This may have been done to accommodate the changes made to the laboratory experiments. He may have felt that more guidance was needed to help the students figure out how to solve the problems brought up in the new activities, especially since these activities were field-tested in this class. He may also have helped initiate student communication with peers, as some laboratory activities required students to share and interpret data. It appeared that something changed in how students in the two groups perceived the analytical laboratory. The differences centered on the level of interaction and the types of questions asked. Both control classes asked the instructors questions about procedures and calculations. At first there was very little interaction between students as they worked; but after about a month, students began to ask their peers questions like “does this [titration] look like it’s the right color?” Most students seemed to be more comfortable watching their peers than asking questions of them while in the laboratory. By the second month more questions were being asked of peers. These questions focused mostly on procedures and what results might be expected. Students still asked the instructor questions about why various things happened, but peers became a source of information as well. Some students responded to questions with answers that were more than just yes or no; explanations and advice were also offered. The experimental classes began in much the same way. Students watched each other but didn’t say much; they relied on the instructor to answer questions. This pattern began to change during the third and fourth weeks. One new activity done in early September involved learning how to use and calibrate an automatic pipet. Owing to equipment limitations, students had to share the automatic pipets. Students at each of the three laboratory tables shared knowledge about the pipets and the proper technique for their use. The more knowledgeable students seemed willing to share their expertise instead of letting the instructor field questions. Most of what was shared was “this works for me” strategies, but this level of interaction between peers was not seen as early in the control classes. The level of communication continued after this activity was completed. The experimental classes had to work together to complete the gravimetric calcium laboratory, which began in late September. This activity required students to devise a plan to dissolve a commercial calcium supplement, then analyze the supplement according to a set procedure. The level of student interaction increased dramatically with this activity. Students discussed how they would try to dissolve the sample. Students at each laboratory bench delegated authority and reported back to the group. Even though each student was responsible for analyzing his or her own sample, the interaction necessary to begin the laboratory continued past the initial shared portion of the experiment. Students discussed problems they encountered in the laboratory with their peers as well as with the instructor. They also began to ask their peers to explain and to justify their answers and procedures. As the semester progressed, students in the experimental class began to ask each other more in-depth questions than did students in the control group. In both groups, many questions focused on “what did you do” and “how did you do it”. However, the experimental classes asked more “why” questions. They challenged each other to explain why one person’s experimental results differed from theirs. They also asked each other to justify what they did, not just explain it. The control classes may

also have thought about why and how to justify results, but they did not verbalize this to their peers in the laboratory. These changes in student interactions suggest some attitude differences between the groups. Early communication in the experimental classes may have been induced by the need to share equipment and data. However, the fact that communication continued and broadened in scope suggests something more. Their willingness to share information, discuss problems, and justify responses suggests that students gained confidence in their abilities and knowledge of the laboratory. This indicates an attitude change, which may be based on perception of self as competent and able. Along with increased communication came comments about the laboratory itself— what was enjoyable and what was not. Comments expressing dissatisfaction with the laboratory also surfaced in the experimental classes. These often centered on the desire to do more or to have more control over the laboratory experiments. It is here that the interview data suggest some attitude changes as well.

Interviews The interviews were coded on the basis of the four main themes (Tables 3–6), which emerged after multiple readings of the interview transcripts and after looking for patterns among the topics, thoughts, and feelings that came through in the interviews. Each theme was further divided into subthemes, and assignment of thoughts and statements from the interviews was made to these subthemes. The coding scheme and some of the coded interviews were discussed with two other science educators. One of them had a strong background in chemistry; the other had expertise outside of chemistry. Both educators examined the coding scheme and coded portions of two to four interviews. Discrepancies were few; those that arose were discussed between the researcher and the science educators, and a final code was assigned that satisfied all parties. Inter-rater reliability data are not available. For the purposes of this study, a “significant” subtheme constitutes three or more occurrences of an idea expressed by at least half the students interviewed within the control or experimental classes. The responses are shown in Tables 3–6 as a ratio of the number of times an idea was expressed to the number of students expressing that idea. For example, to be a significant subtheme for the experimental classes, at least 5 students had to make 15 or more statements described by that subtheme. Responses were also examined by gender. Subthemes were regarded as significant if 3 of the 5 women interviewed (control and experimental classes combined) mentioned the subtheme at least 9 times. Similarly, if at least 5 of the 9 men (control and experimental classes combined) brought up a subtheme a total of 15 times or more, it was regarded as significant. In the following discussion of themes and subthemes, only significant responses are included. Theme 1 The majority of the responses falling under Theme 1, student’s self-perceptions in the analytical laboratory (Table 3), suggested that students from all classes feel competent, aware, and able to perform the techniques expected of them. Familiarity through review and repetition was a significant subtheme for both the control and experimental classes. Students from all classes expressed feelings of competence and understanding much more frequently than feelings of confusion or lack of

JChemEd.chem.wisc.edu • Vol. 76 No. 1 January 1999 • Journal of Chemical Education

103

Research: Science and Education

Table 3. Theme 1: Students' Self-Perceptions of the Analytical Laborator y Significancea

Subtheme

Group

NO

NS

Examples of Students' Statements Related to Subtheme

Techniques and laboratory skills

C

11

3

In this class you learn to titrate.

Reasoning and metacognitive awareness

E

17

6

[I need to] try and figure out why, where they went wrong, and then go back and correct it and see if that helped and if that didn't help, try and figure something else out.

Familiarity via review and repetition

C E

10 20

4 8

I especially got much better at it towards the end and then I really liked it.

Applications outside academia

C

6

3

Basically the thing that I learned was, in the experience was that, it was kind of like if I were going to be working for a company.

General concepts about chemistry and the laboratory

E

28

9

...your errors, and indeterminate and your determinate errors.

aThe subtheme was significant for the group indicated (C = control; E = experimental). N is the number of occurrences of statements related O to the subtheme (the total number of times ideas related to the subtheme were expressed). N S is the number of students in the group who made statements related to the subtheme. The criterion for “significance” was that at least half of the students interviewed should make one or more statements related to the subtheme, and that there should be at least 3 occurrences of ideas related to the subtheme. Four students from the control group and 10 from the experimental group were interviewed.

Table 4. Theme 2: Positive and Negative Feelings toward the Laborator y Significancea

Subtheme

Group NO NS

Examples of Students' Statements Related to Subtheme

Positive and negative feelings toward specific laboratories

C E

11 41

4 10

Expressions of familiarity, fun and learning

E

30

8

There was so much chemistry going on. It was kind of fun to try and keep everything straight, and plus it worked out well [water hardness lab].

Learned from the experience but didn't enjoy it

C E

8 15

3 7

I actually learned the most from the copper, which is one that I didn't like.

Positive peer interactions in the laboratory (only experimental group)

E

38

9

We had a really good group at the bench. So it was a really positive interaction. If we didn't understand something we could ask the person next to us. And they would explain it.

aSee

I liked all the titrations. It's just that, it's [Eriochrome black T] a lousy indicator on that one [water hardness].

footnote to Table 3.

Table 5. Theme 3: Students' Expectations about the Laborator y Significancea

Subtheme

Group NO

NS

Examples of Students' Statements Related to Subtheme

What students anticipated learning

C E

8 26

4 10

...how to do things that I figured we'd be doing if we were working for a chemical company. ...but then after a week or so of class I expected good lab techniques.

What students say they learned

C E

9 30

2 10

In the very beginning I wasn't real confident at it, and messed up some of them [titrations], and as you get more experience you gain more confidence in your ability.

What students wanted to learn but didn't learn

C E

12 28

4 7

aSee

I'd probably like to spend more time on the, doing the NMRs and IRs and, and the UV–vis, too...[we] got plenty of theory on it. We didn't get enough real practice on it.

footnote to Table 3.

Table 6. Theme 4: What Students Feel They Need To Know and Do as Scientists Subtheme

Group NO

NS

Examples of Students' Statements Related to Subtheme

General knowledge needed

E

21

8

A lot of basic chemistry…like the complex ion formation reactions and acid–base chemistry and precipitation chemistry and oxidation– reduction chemistry and then a little bit of spectroscopy, but not much.

Laboratory techniques and skills

C E

8 28

4 8

…you can really get a lot of lab technique and a lot of chemistry out of that class, which obviously is going to help you with research.

Personal qualities of scientists

C

9

4

Hm, to never be satisfied that you yourself reached an endpoint. That you continue to grow, continue to learn.

aSee

104

Significancea

footnote to Table 3.

Journal of Chemical Education • Vol. 76 No. 1 January 1999 • JChemEd.chem.wisc.edu

Research: Science & Education

knowledge. This may be due to actual self-perception, or to unwillingness to talk about incompetence or lack of knowledge. The control classes made more statements classified under subthemes relating to techniques and laboratory skills and applications outside of academe. It is surprising that the experimental classes had less to say about applications, as the laboratory experiments they used were designed to incorporate real-world applications. The experimental classes made more statements classified under subthemes related to reasoning and metacognitive awareness and general concepts about chemistry and the laboratory. A gender difference was found in the subtheme related to reasoning and metacognitive awareness. Women stated almost twice as often as men that they felt laboratory helped them develop metacognitive and reasoning skills. Theme 2 Theme 2 addressed positive and negative feelings toward the laboratory (Table 4). All of the classes expressed many statements classified as subthemes 1, positive and negative feelings toward specific laboratories, and 3, learned from the laboratory but didn’t always enjoy it. Subtheme 2, expressions of familiarity, fun, and learning, was significant for the experimental classes and for the women interviewed. It is possible that this reflects a more positive attitude in the experimental classes in general. The difference in responses coded as subtheme 2 was large—3 students in the control classes made 4 comments reflecting this subtheme, compared to 30 comments made by 8 members of the experimental classes. Men were more likely to verbalize subtheme 3. Theme 3 Theme 3 addressed student expectations about the laboratory (Table 5). Three subthemes were significant for both the control and experimental classes. There was also a gender difference in one of the subthemes. Three women made 18 comments classified as subtheme 3, things they wanted to learn in the laboratory but didn’t. Eight men made 22 remarks within the same subtheme. This may suggest that the women were more likely than men to express their opinions about what they wanted to learn. Perhaps women had different expectations for the laboratory and expressed them as mismatches between expectations and what actually happened. There were no gender differences in responses for the other two significant subthemes, what students anticipated learning and what they said they learned. Theme 4 Theme 4, what students feel they need to know and do as scientists (Table 6), showed some differences in responses between the control and experimental classes. Subtheme 2, laboratory techniques and skills, was significant for both the control and experimental classes. Subtheme 1, regarding general knowledge, was expressed far more often by the experimental classes (21 statements, 8 people) than by the control classes (3 statements by one person). This difference may be due to an increased awareness of the importance of knowing a lot of “stuff”—not just specific techniques. Perhaps the inclusion of real-world issues in the experimental classes’ activities contributed to this difference; or perhaps the experimental classes were more willing to express their concerns. Students in the control classes made more statements about

the personal qualities of scientists (subtheme 3), although the difference is not great. Four control-class students made 9 responses, compared to 11 responses by 4 experimental-class students. A gender split was seen in statements about future needs, which overall was not a significant subtheme. women were less likely than men to speculate on future needs. Conclusions Students who did the revised laboratory experiments were expected to have more positive attitudes toward the usefulness and applicability of analytical chemistry than students who did not experience the revised experiments. Although survey data do not suggest attitude changes, interview and observational data do suggest changes. Observations indicate that student–student interactions were at greater depth and more prevalent in the experimental than in control classes. This may indicate a deeper level of involvement with the material and suggests more positive attitudes about the laboratory. Deeper involvement may also suggest more confidence with the course material. The interview data show that experimental classes did express more confidence than control classes about their selfperceived abilities. Moreover, a greater number of positive statements about the laboratory, relative to the number of students interviewed, arose from the experimental classes. The experimental classes also expressed more confidence about reasoning and metacognition, familiarity, and general concept knowledge than did the control classes. The experimental classes also identified more expectations that were not fulfilled by the laboratory. This may be interpreted in two ways. The unfulfilled expectations may be seen as a negative attitude, or as expressions of students’ increased awareness of their needs and learning strategies and how they can change them. None of the data strongly support or refute the belief that students enrolled in first-semester analytical chemistry classes will develop a better understanding of how practices and techniques associated with analytical chemistry can be applied to many areas of chemistry and to science in general. The experimental group talked more than the control group about general knowledge and techniques and skills being important to scientists. Self-perceptions seemed a little stronger in the experimental group as well, though neither self-perception nor techniques and skills were linked to an understanding of how analytical chemistry fits into a “big picture.” Summary Much previous research on science attitudes and attitude development among science educators has been quantitative, and controversy still surrounds the measurement of attitudes. This study provides a different way to understand students. A deeper understanding of what and how students think and feel can be gained through carefully constructed interviews and observations. Qualitative techniques may thus provide deeper and more focused information to enhance an understanding of student attitudes and their effects on learning. This study has important implications for laboratory design. Laboratory activities that integrate application seem to correlate with positive student attitudes. The inclusion of applications can be useful in designing laboratory activities. This method of instruction may also increase interest in ana-

JChemEd.chem.wisc.edu • Vol. 76 No. 1 January 1999 • Journal of Chemical Education

105

Research: Science and Education

lytical chemistry and improve understanding of its uses and limitations, although more research needs to be done on this issue. The depth of student learning as well as teaching practices that facilitate learning using this laboratory style also need to be researched further. The observed gender differences were unexpected. Men were more likely to express familiarity with the course material, whereas women were more likely to express confidence in their techniques and skills and in their metacognitive abilities. Women were more likely to make positive statements about the laboratory and were more likely to talk about problems with the procedures. They were also more likely to verbalize unmet expectations in the laboratory. These gender differences may be related to differences in how male and female students view the laboratory, or may be artifacts of the sample size. More work needs to be done to determine the significance, if any, of the gender differences found in this study. Owing to the narrow focus and small sampling of students in this study, several important limitations must be noted. First, the class size in the laboratory sections studied is small. This may prevent application of broad generalizations about attitude changes to larger groups or to courses with different structures. A second limitation arises from the homogeneous grouping of students in the classes. Since most students taking analytical chemistry are chemistry, biology, or medical technology majors and probably possess some initially positive attitudes about chemistry, perhaps not much attitude change should be expected to be shown by the surveys. The interview data likely reveal more about the nature of student attitude change and the attributes of the laboratory program that may influence change.

106

Literature Cited 1. Harwood, J. J. J. Chem. Educ. 1989, 66, A268–A271. 2. Walker, M. Educ. Chem. 1974, 1, 58–60. 3. Edeani, D. O. Presented at the Annual Meeting of the Association for Education in Journalism, Athens, OH, July 1982; ERIC Document Reproduction Service No. ED 218640. 4. Walters, J. D. Anal. Chem. 1991, 63, 977A–985A. 5. Roth, W.-M.; Roychoudhury, A. J. Res. Sci. Teach. 1994, 31, 5–30. 6. Shrigley, R. L. Sci. Educ. 1983, 67, 425–442. 7. Levin, J.; Klindienst, D. Presented at the Annual Meeting of the National Association for Research in Science Teaching, Dallas, TX, April 1983; ERIC Document Reproduction Service No. ED 244837. 8. Hawkins, J.; Pea, R. D. J. Res. Sci. Teach. 1987, 24, 291–307. 9. Yager, R. E. J. Coll. Sci. Teach. 1989, 18, 273–275. 10. Chisholm, C. Correlates of Math Avoidance Responsible for Filtering Individuals from Math/Science Areas; M.S. Thesis, Towson State University, 1980. 11. Frechette, M.; Farina, J. J. Chem. Educ. 1979, 56, 597–598. 12. Dori, Y. Presented at the Annual Meeting of the National Association for Research in Science Teaching, San Francisco, CA, March 1989; ERIC Document Reproduction Service No. ED 319605. 13. Strube, P. Res. Sci. Educ. 1991, 21, 300–305. 14. Kern, E. L.; Carpenter, J. R. J. Geol. Educ. 1984, 32, 299–305. 15. Schmidt, G. L.; Carpenter, J. R. J. Geol. Educ. 1976, 24, 61–66. 16. Neman, R. L. J. Coll. Sci. Teach. 1983, 13, 16–18. 17. Krynowsky, B. A. Sci. Educ. 1988, 72, 575–584. 18. Shrigley, R. L. & Koballa, T. R. Sci. Educ. 1992, 76, 17–42. 19. Weinburgh, M. H.; Englehard, G., Jr. School Sci. Math. 1994, 93, 118–123. 20. Munby, H. J. Res. Sci. Teach. 1983, 20, 141–162. 21. Hofstein, A.; Ben-Zvi, R.; Samuel, D. Sci. Educ. 1976, 60, 401–411. 22. Shavelson, R. J. Statistical Reasoning for the Behavioral Sciences, 2nd ed.; Allyn & Bacon: New York, 1988.

Journal of Chemical Education • Vol. 76 No. 1 January 1999 • JChemEd.chem.wisc.edu