Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

JOURNAL OF RESEARCH IN SCIENCE TEACHING VOL. 31, NO. 2, PP.

197-223 (1994)

Experimenting in a Constructivist High School Physics Laboratory

Wolff-Michael Roth
Faculty of Education, Simon Fraser University, Burnaby, BC, Canada V5A IS6

Abstract
Although laboratory activities have long been recognized for their potential to facilitate the learning of
science concepts and skills, this potential has yet to be realized. To remediate this problem, researchers
have called for constructivist learning environments in which students can pursue open inquiry and frame
their own research problems. The present study was designed to describe and understand students’
experimenting and problem solving in such an environment. An interpretive research methodology was
adopted for the construction of meaning from the data. The data sources included videotapes, their
transcripts, student laboratory reports and reflections, interviews with the students, and the teacher’s
course outline and reflective notes. Forty-six students from three sections of an introductory physics course
taught at a private school for boys participated in the study. This article shows the students’ remarkable
ability and willingness to generate research questions and to design and develop apparatus for data
collection. In their effort to frame research questions, students often used narrative explanations to explore
and think about the phenomena to be studied. In some cases, blind alleys, students framed research
questions and planned experiments that did not lead to the expected results. We observed a remarkable
flexibility to deal with problems that arose during the implementation of their plans in the context of the
inquiry. These problems, as well as their solutions and the necessary decision-making processes, were
characterized by their situated nature. Finally, students pursued meaningful learning during the interpreta-
tion of data and graphs to amve at reasonable answers of their research questions. We concluded that
students should be provided with problem-rich learning environments in which they learn to investigate
phenomena of their own interest and in which they can develop complex problem-solving skills.

Access to laboratories and experience of inquiry have long been recognized as important
aspects of school science. Most of the curricula developed in the 1960s and 1970s were
designed to make laboratory experiences the core of the science learning process (Shulman &
Tamir, 1973). Science in the laboratory was intended to provide experience in the manipulation
of instruments and materials, which was also thought to help students in the development of
their conceptual understanding. Yet teaching and learning in high school laboratories are
plagued with problems. The study of science through laboratory-centered programs is still
falling short of the expectations science educators held since the inception of these programs
(Hofstein & Lunetta, 1982; Tarnir, 1989). Although teachers’ lack of enthusiasm, problems in
classroom management and student achievement, and problems in the implementation of inqui-
ry are cited as factors (Stake & Easley, 1978; Welch, Klopfer, Aikenhead, & Robinson, 1981),
alternative explanations for the lack of success of laboratory instruction are seldom explored.

0 1994 by the Nationai Association for Research in Science Teaching


Published by John Wiley & Sons, Inc. CCC 0022-4308/94/020197-27
198 ROTH

A vivid description of the situation in science laboratories was provided by ethnographic


studies of high school science in Australia and the U.S. (Gallagher & Tobin, 1987; Tobin &
Gallagher, 1987). These authors discuss several elements as constitutive for the problems with
laboratory teaching. Experimental tasks often embody a cookbook approach. Students follow
recipes gathering and recording data without a clear sense for the purposes, procedures, and
their interconnections. These tasks have low cognitive demands and provide a context that
precludes reflective thought and concentration. Consequently, students engage in activities not
intended by the curriculum planners. They spend much of their laboratory time in off-task
activity with short periods of attention to get the work completed. The time off task is used for
nonscience related socialization with peers all over the classroom. To make up for the time lost
in these activities, teachers set a high pace in regular classroom periods to transmit factual
information.
Many of these problems related to teaching and learning have been linked to the objectivist
epistemology underlying most of current science teaching (Tobin, 1990a; Tobin, Espinet, Byrd,
& Adams, 1988; Tobin & Ulerick, 1989). Objectivists attribute to knowledge an existence
independent from the knower. Accordingly, such knowledge, which matches reality in a one-to-
one correspondence, can be transferred from an authority to a more passive learner. From this
perspective, lectures are the preferred modes of transmitting knowledge. Objectivism has been
subject to substantial criticisms and an alternative epistemology, constructivism, is being ac-
cepted as a more appropriate conception of knowledge in such fields as education, epistemol-
ogy, history and philosophy of science, cognitive and social psychology, philosophy, and
sociology of knowledge (Bruner, 1986; Gergen, 1985; Goodman, 1978; Knorr-Cetina, 198 la;
Toulmin, 1982; von Glasersfeld, 1987). Constructivists hold that learning is an interpretive
process, as new information is given meaning in terms of the student’s prior knowledge. From a
constructivist point of view, each learner actively constructs and reconstructs his or her under-
standing rather than receiving it from a more authoritative source such as a teacher or a
textbook. As a consequence, constructivism implies that learners must be given
opportunities to experience what they are to learn in a direct way and the time to think and
make sense of what they ure learning. Laboratory appeal as a way of allowing students to
learn with understanding and, at the samc time, engage in the process of constructing
knowledge by doing science.” (Tobin, 1990b. p. 405. emphases added)

It is evident that constructivism and its implications are in contradiction with the current
teaching practices reported above.
A related criticism of the reigning, objectivist epistemology comes from research spanning
and interconnecting cognitive anthropology, cognitive science, and social psychology (Collins,
Brown, & Newman, 1989; Lave, 1988; Newman, Griffin, & Cole, 1989; Rogoff, 1990; Roth &
Roychoudhury, I993a). These researchers hold that learning, rather than being a transfer of
knowledge, is a cognitive apprenticeship during which students are introduced into the practices
of a culture. The crux of the matter is that current school practices constitute a culture in itself.
Researchers in this area have indicated that the current problems of schooling are related to the
distance between practices at school and those of everyday life, whether these are everyday
arithmetic practices people demonstrate while shopping, working in a dairy factory, tailoring
clothes, or selling lottery tickets in street markets (Carraher & Schliemann, 1982; Lave, 1988;
Lave, Murtaugh, & de la Rocha, 1984; Schliemann & Acioly, 1989; Scribner, 1984). In
addition, schools proclaim to prepare students for future careers in fields such as mathematics,
science, or history. Yet the skills fostered at school differ from those needed in these fields
(Brown, Collins, & Duguid, 1989).
A CONSTRUCTIVIST PHYSICS LAB I99

Key to the learning environments these critics propose as alternatives are contexts where
students work with others on common, genuinely problematic tasks. These others preferably are
more advanced adults and peers but could also be peers of equal or less advanced standing. By
working together on problems, new knowledge is first constructed collaboratively in the joint
problem space from which the learner subsequently appropriates it, that is, individually con-
structs his own representations (Newman et al., 1989; Rogoff, 1990). These critics also high-
lighted a second issue: Learning is essentially situated in and thus is a function of the physical
and social context. As a consequence, learning from projects is more advantageous than learn-
ing from isolated problems because “students can face the task of formulating their own
problems, guided on the one hand by the general goals they set, and on the other hand by the
‘interesting’ phenomena and difficulties they discover through their interaction with the environ-
ment” (Collins et al., 1989, p. 487). Such projects then set the stage for constructivist problem-
solving situations as advocated by Wheatley (1991), where individuals have to decide on what to
do to arrive at a solution when it is not clear what needs to be done.

Problem Solving
One can identify three distinct lines of research which study the forms and contents of
human problem solving. The first is a cognitive approach, taken largely by such fields as
cognitive psychology, cognitive science, and artificial intelligence. This approach locates cogni-
tion in the individual, and identifies individual differences or differences between groups of
varying expertise. The second line of research takes a cultural approach. It has grown largely
out of the work of the Russian psychologist Vygotsky and the work of anthropologists. In the
view of this line of research, cognition is not merely located in the individual, but to a large
extent in the dialectic between individuals and their culture. A third line of research has emerged
from the sociology of scientific knowledge. It recognizes that cognition goes beyond the indi-
vidual and the culture, and includes the physical context of the acting individual as well. Thus,
cognition is largely situated, including individual, social, and physical contexts. Each of these
three approaches provides us with a distinct view of problem solving. These views will serve as
a reference in our study of students’ problem solving and thinking in an open-inquiry laboratory.

Problem Solving in Limited, Well-DeJned Domains


Most cognitive psychologists assume that all the resources for solving problems are located
in the reasoner’s head. This individual has available specific declarative knowledge and a range
of procedural skills that aid in the solution of a problem (Heller & Reif, 1984). Declarative
knowledge is usually domain specific about concepts, principles, and theories, but can also be
knowledge of specific algorithms. A range of procedural skills are thought to be domain
independent. Among these are such heuristics as the difference-reduction method, means-end
analysis, working backward, or problem solving by analogy (Anderson, 1985) and defining the
problem, designing a problem solution, or implementing the solution (Eylon & Linn, 1988).
Typically, researchers from this perspective observe small numbers of individuals, describe their
behaviors in detail, and build models, often to be implemented as a computer program, which
simulate the observed performance (Larkin, McDermott, Simon, & Simon, 1980). They con-
trast the behaviors of experts, normally professors, with those of novices such as undergraduate
students as they solve well-defined problems from well-structured knowledge domains. In some
instances, successful computer programs have been developed that assist novices in learning
about a domain and in becoming successful problem solvers.
200 ROTH

Problem Solving in Everyday Activity


Evidence suggests that problem solving is not an abstract context-free competence that can
be applied across varying domains. Dairy workers, grocery shoppers, market vendor children,
carpenters, bookies taking street-comer bets, and Liberian tailors are highly accurate (-99%) in
solving problems that emerge out of their everyday practices (Carraher & Schliemann, 1982;
Lave, 1977; Lave, 1988; Lave, Murtaugh, & de la Rocha, 1984; Schliemann & Acioly, 1989;
Scribner, 1984). On the other hand, their performance drops significantly on structurally equiva-
lent paper-and-pencil problems, although the problems were embedded in familiar content.
Most of these studies also showed that the performance on situated and practice-related prob-
lems was independent of the level of schooling. On the other hand, experiments with dairy
workers and carpenter apprentices showed that students with more schooling but no practical
experience performed on situated tasks at levels lower than those at which individuals with less
schooling but with working experience performed. Researchers in this field have concluded that
in practical situations the social and physical contexts provide information and resources the
actor uses to establish an appropriate solution to the problem. They also believe that the
development of these effective problem-solving strategies is due to the fact that knowledge
constructed in practice “is the locus of the most powerful knowledgeability of people in the
lived-in world’ (Lave, 1988, p. 14).

Problem Solving by Scientists at Work


Recent ethnographic studies in scientific laboratories revealed an image of scientists that
has little to do with the rational reasoner who applies a set of special science process and
problem-solving skills. Rather, this area of research has shown that the reasoning of scientists is
characterized by its local nature, depending on the research context and on the concrete research
situation (Knorr-Cetina, 198la; Latour & Woolgar, 1979). Thus, “the contingency and contex-
tuality of scientific action demonstrates that the products of science are hybrids which bear the
mark of the very indexical logic which characterizes their production, and are not the outgrowth
of some special scientific rationality to be contrasted with the rationality of social interaction”
(Knorr-Cetina, 198la, p. 33). Accordingly, local idiosyncrasies, know-how, interpretations,
physical (equipment and material), and social (co-workers) resources determine the accomplish-
ments within scientific laboratories. Scientists constantly refer to this circumstantial nature of
their inquiry. Projects take certain turns because they had pieces of equipment developed for
another project but that they could use, because certain apparatus was available, or because they
happened to come across a specific paper (Knorr-Cetina, 1981b). The myth of the scientific
rationality and problem solving is established when the scientists strip the contextual factors and
report their new constructions as if they were the product of unaltered intentions. Finally,
scientific reasoning is heavily affected by a web of transscientific social relations in which the
scientists locate their laboratory reasoning and action (Knorr-Cetina, 198la; Latour, 1987).

Purpose
At present, there is a lack of interpretive studies on learning in science laboratories, and an
even greater paucity of research on constructivist laboratory environments that emphasize both
the individual and collaborative construction of knowledge. The purpose of this study was to
add to the fundamental understanding of the cognitive processes in constructivist laboratories.
Specifically, we were interested in finding answers to questions such as “How do students
communicate with peers to identify a problem?’ “How do students design strategies for solving
A CONSTRUCTIVIST PHYSICS LAB 201

the problem?” and “How do students implement a plan to collect and analyze data to answer the
problem?’ Our intent was to provide a thick description of the students’ cognitive processes
rather than a quantification of their prevalence.

Research Design
In order to arrive at answers to our questions, we used an interpretive methodology based
on naturalistic inquiry (Lincoln & Guba, 1985) and constant comparative analysis (Strauss,
1987). Because we wanted to understand how meaning arises out of the laboratory work of
students, our methodology was informed by microsociological approaches to understanding
everyday work practices, planning behaviors, institutional analysis, and classroom discourse
(Burrell & Morgan, 1985; Cazden, 1986; Garfinkel, 1967; Suchman, 1987). Direct observa-
tions, videorecordings and their transcripts, student laboratory reports, student interviews, the
teacher’s reflective notes, and students’ reflective notes as well as a reflective essay constituted
the data sources for this study.

Participants
Forty-six students from three sections of an introductory physics course for high school
juniors participated in the study. The institution, a private school which at that time was for boys
only, is located in central Canada. The physics course, in spite of being an elective, drew a
higher-than-average proportion of the students, 42 of the 67 juniors at the school. Four grade 10
students also took the course. However, although generally college bound, few students from
this school (around 5 % ) eventually select university programs in science or mathematics, but
predominantly opt for careers in business, law, and engineering. The author taught all three
sections of the course. In addition to graduate degrees in physics and science education, he has
had 10 years of experience teaching science using nonlecture, laboratory-based approaches. He
used a graduate student advisor metaphor as referent for the planning of, and acting in, the
learning environment. A university-based researcher facilitated and critiqued the analyses
emerging from this project to ascertain the trustworthiness, quality of the hermeneutic process,
and the authenticity of the constructions.

The Classroom Setting


The teaching-learning context of this study was grounded in a constructivist epistemology
that emphasized both the individual and collaborative construction of knowledge. Students were
encouraged to take individual responsibility for their learning and to participate in the decision
making with respect to such issues as assessment, organization of the learning environment,
access to resources, and the establishment of research teams.
At the beginning of each unit, the students received a handout, outlining the suggested and
compulsory activities for a given period. There were four types of activities. First, the experi-
ments constituted the core of the course around which all other activities revolved. Second,
students read relevant sections from the main textbook and at least one additional source, then
summarized the key ideas by constructing concept maps (Novak & Gowin, 1984). Third, the
students submitted each week six to eight textbook problems of their own choice, provided that
problems came from different textbook sources but were on the current topic. By assigning
readings and a minimum number of textbook problems we ascertained that the formal cumcu-
lum requirements set by the province were satisfied. Finally, students prepared short notes on
202 ROTH

the biographies of scientists or wrote essays on special topics not normally covered in the
textbook.
Students were assessed in groups for their laboratory work, laboratory reports, and the
concept mapping. Each group member received the same grade. Students were assessed on an
individual basis for textbook problems and writing assignments. In addition, individual under-
standing was assessed with four in-class tests distributed over the year, and during two formal
examinations, one in December, the other at the end of the school year, in June. The December
examination also included a laboratory component worth 20% of the exam mark. To complete
this component, the students planned and executed an experiment in their laboratory groups, but
submitted individual reports.
Students spent much of their class time on experiments (about six to seven of the nine
periods in a two-week cycle), although they were free to do the problems, library research and
writing assignments during class. One period per week was used for whole-class discussion,
review, sharing research results with other students, and introducing new tools such as measure-
ment instruments or computer programs. Outside of the scheduled classroom time, the students
used the lab and its facilities during the spare periods, after school, in the evening, or during the
weekend. The classroom was always open in the sense that students from any section could
work on their experiment unless it interfered with the work of the students scheduled for that
period. This arrangement provided yet another avenue for students to see and talk about
different ideas.
The students were introduced to new units with a range of demonstrations during which
they also encountered new instruments, apparatus, or software. They were encouraged to
research those questions that arose out of these demonstrations. The demonstration materials
were then made available for the students to familiarize themselves with the equipment and
tools. Subsequently, they began an investigation, in groups of three or four, by formulating a
focus question and planning the data collection (one period)-for the first experiment in a unit,
the teacher often suggested a research question. The students then set up the apparatus, collected
the data, and submitted the data to a computer-based mathematical and graphical analysis (two
to four periods). Each group discussed its results, consulting other groups and the teacher (one
period), and then prepared a report.
Each group received feedback from the teacher that outlined both the strengths and weak-
nesses of the report. This feedback included alternate interpretations, different graphical and
statistical analysis procedures using the student data set as an example, suggestions for error
sources, suggestions for everyday applications, general encouragement, and praise. Marks were
assigned, but they were used to encourage students to improve, and students were provided with
the opportunity to rewrite a report. These marks accounted for about 70% of a students’ term
mark; the other 30% was derived from essays, homework questions, and tests.

Data Sources and Analyses


To allow a microanalysis of the processes in a constructivist laboratory, we videotaped 30
laboratory sessions during the fall term of 1990. During this period, the syllabus included topics
from Newtonian mechanics and a unit on heat and temperature. The syllabus during the remain-
der of the year included waves and sound as well as electricity and electromagnetism. We also
collected the records of all student activities, that is, all laboratory reports produced by the
students during the 1990191 school year and the teacher’s typewritten feedback for each labora-
tory report. We supplemented our primary data sources listed above by including the teacher’s
course of study (provincially mandated, teacher-produced course outline including content,
A CONSTRUCTIVIST PHYSICS LAB 203

objectives, resources, teaching-learning strategies, student evaluation, self-evaluation), reflec-


tive notes, the students’ essays on “Knowing and Learning Physics,” and interviews with the
students that focused on their views of knowing and learning physics as additional sources of
data.
The school-based researcher viewed and re-viewed the videotapes as soon as possible after
they were recorded by a laboratory and teaching assistant. During these first viewing sessions,
the teacher constructed initial, tentative assertions. These initial assertions directed our further
data collection. After the first three weeks of the study, both researchers met in several sessions
to view the tapes jointly. During these sessions, we adopted a technique employed by anthro-
pologists and sociologists who study various interactive behaviors (McArthur, Stasz, &
Zmuidzinas, 1990; Suchman, 1987). We repeatedly watched the videotapes, discussing and
writing down our reactions. In many cases, we began with an initial description of a phenome-
non that seemed to be interesting, such as “blind alley.” When viewing other tape materials
later, we tried to construct a better understanding by refining, modifying, or discarding an initial
description. Through this process, we constructed categories such as “framing and solving of
authentic problems ,” “blind alley,” or “concrete mediator” (see the Results and Discussion
section) that we tested in the remainder of the data sources. From these sessions emerged new
data-collection strategies with the intent of gathering additional information to support our
constructions. For example, after we identified the category of blind alley, we began noting
further instances of this category with groups other than those appearing on film. Thus, these
discussions helped us to collect further instances of categories.
The transcripts of the videotapes constituted the major source for reconstructing the stu-
dents’ understandings of the activities in which they were involved. Three to four students
worked together in each group to frame questions and to answer them experimentally. In order
to accomplish the complex tasks involved, they talked with their peers or with students from
other groups. Through this communication each student made available to others his under-
standing of the problem, the current status of the task, his ideas about the next move, how much
progress has been made, or how much remains to be done. Thus, our transcripts of these
interactions constituted natural protocols which provided us with the opportunity to reconstruct
the meaning of the students’ communication. Such protocols of collaborative work provided us
also with evidence for the individual reasoning of participants, for the processes of the partici-
pants’ meaning attribution to the context, and for a sense of how participants see their activities
as coherent and consistent (Brown, Rubinstein, & Burton, 1976; Burrell & Morgan, 1985).
We analyzed the written artifacts independently at first, then collaboratively constructed
assertions following a process similar to that for the tapes. In addition to face-to-face meetings,
we communicated amply throughout the study by using an electronic mail facility. The present
report emerged out of these and additional meetings held throughout the year and out of an
intensive work period during the summer following the school year. We constructed patterns of
student understanding of their task and task status at various points of progress. We have
provided several instances of these patterns to show similarities and variations across groups.

Unit of Analysis
The videotapes and laboratory reports provided us with a window on collaborative work. In
our effort to assess the learning from these sources it became apparent that the individual was
not the most useful unit of analysis: in many cases, the cognitive and interactional achievement
of individuals and/or the group could not be separated. For the same reasons, various units of
analysis have been suggested in the past for the analysis of social units. Among these were
204 ROTH

activity or event (Cole, 1985); socially assembled situations or cultural practices (Laboratory of
Comparative Human Cognition [LCHC], 1983); task within practices or work task (Scribner,
1984); whole task (Newman et al., 1989); and work sequences, networks, or techniques of
argument (Latour & Woolgar, 1979). Although we had a hunch that the structure of the
interactions-the unit of analysis for this study-may lie in the relation between three compo-
nents, the individual, the social context, and the physical context, we did not want to presume
any particular structure. We began our study with the assumption that we lacked a description of
the structure of the interaction between collaborating peers in the contexts provided. In the
process of our study it became clear that many accomplishments could not be attributed to
specific individuals, but arose out of the interaction of groups in a specific setting. Thus, for the
most part we report here the achievements of groups and focus on individuals only in a few
specific instances. Although we went into the study without assuming a particular structure for
the students’ interactions, we knew that our epistemological commitment to constructivism
would structure our analysis. For example, we assumed that the actions and communications we
had recorded were traces of the students’ sense-making activities. Furthermore, it was clear to
us that any description we could develop would arise out of the meanings we attributed to the
actions and interactions as well as the products thereof.

Results and Discussion


This study was designed to investigate whether and how students in a constructivist labora-
tory environment (a) frame questions for laboratory inquiry and design strategies for finding
answers, (b) implement their plans to collect data, and (c) analyze the data to answer their initial
question and construct new knowledge. Our results are in marked contrast to earlier studies in
science laboratories, which indicated that open inquiry was too confusing and did not work for
most students, that learning outcomes were too uncertain, and that students were little con-
cerned with meaningful learning but mostly pursued their own agendas of a social nature
(Costenson & Lawson, 1986; Gallagher & Tobin, 1987; Welch et al., 1981). We found a
remarkable ability and willingness to generate research questions, to design and develop appara-
tus for data collection, to deal with problems arising during implementation out of the context of
inquiry, and to pursue meaningful learning during the interpretation of data and graphs to arrive
at reasonable answers of their focus questions. And the students demonstrated the ability to
function in an environment in which they constructed their science in a social setting.
Although there are many possible ways for structuring this account, we chose to present our
findings in such a way that the reader could get a sense for the flow of the events in this
laboratory. Accordingly, the phases of planning, collecting data, and analyzing data provide the
overall structure to our results. This form of presenting our constructions also corresponds to the
full cycle of routine activities typical for scientific work (Traweek, 1988).

Planning
A major part of the students’ planning sessions was taken up by their attempts in framing
research questions. Framing questions is a critical part of any exploration because the research
problem takes shape at this stage. In order to frame and negotiate their research questions, the
students often resorted to narrative descriptions of interesting situations. The use of narrative
descriptions to deal with complex situations is in stark contrast to the logico-mathematical mode
of thought that encourages the immediate isolation and testing of causally related variables. In
A CONSTRUCTIVIST PHYSICS LAB 205

some cases, the students framed a research question and planned an experiment in which they
did not observe the expected effect. We termed such situations blind alleys.

Framing Research Questions.


Of all the research questions investigated, the students framed about two-thirds on their
own. The other one-third of the questions, predominantly at the beginning of a new unit, were
suggested by the teacher. At this point in a new topic, the students had little experience with
either the phenomena or the equipment to investigate these phenomena. After some experience,
the students’ own questions most often arose directly from their work. In other cases, students
wanted to find out about phenomena which they had encountered outside school. For example,
Fred’ had been interested in the wind noise generated by driving cars, and in the effect of wind
friction on the velocity of cars. Thus, two of his group’s experiments were designed to investi-
gate the effect of air friction on a moving cart. Another group, CJP*, independently designed
two experiments on hydrodynamics of falling objects because of their interest in car racing.
They tested the effects the shape of an object as well as the viscosity and density of a retarding
liquid would have on the acceleration and velocity of an object falling in this liquid.
Even when students were presented a question to investigate, the teacher framed it openly
so that students could construct their own meaning. And he gave few details with respect to
specific experimental designs. This led to situations where various groups constructed different
meanings, thus different problems based on the same question. The same question became a
different problem, which students dealt with in essentially distinct ways. For example, for the
question3 “What is the relationship between mass and acceleration in free fall?” students
essentially planned three different types of experiments that led to significantly different results:
A constant acceleration of about 10 m/s2, a constant acceleration but less than 10 m/s2, usually
around 2 m/s2, and a curvilinear relationship with an asymptote for large masses of about 10
m/s2 and an intercept at 0 m/s2 (function a(m) = 1/(1 + M/m)g). Thus, this question had
invited students to make their own decisions and to frame their own problem in a way that
Wheatley (1 991) advocated for a problem-centered constructivist classroom. As a follow up, the
teacher provided each group with the opportunity to present, discuss, and defend their inter-
pretations of the problem in whole-class settings.
The various student interests always led to a remarkable range of topics they investigated
(Table 1). Of the 16 concurrent experiments, one research question (4a) was framed in the same
way twice, with a variation in the substance by a third group (4b). Two other pairs of experi-
ments ( l a and l b and 8a and 8b) were similar. However, these similarities of research questions
were often coincidental rather than the result of copying, particularly when the questions had
been framed by groups from different sections of the course.
There were different routes for groups to reach consensus regarding the research question or
the instrumental and material specifics for each e ~ p e r i m e n tIn
. ~ some groups, large numbers of
topics and/or materials were discussed and rejected, sometimes taken up again only to be
rejected later on, or in some cases to be accepted as a viable focus for an experiment. CJP, for
example, considered 19 topics before settling on their research question (Table 2). In other
groups, such as MCMT, only two alternatives had been given significant attention, and the third
alternative, although brought up by one member, was never really discussed. However, in such
cases, we often observed discussions during which many of the experimental details were
explored. Thus, MCMT spent about 15 to 18 minutes on each of the proposals 1 and 2 ,
exploring the pros and cons of each in terms of the instrumental and material contingencies of
206 ROTH

Table 1
Research Topics Concurrently Developed by 16 Groups

1. a. Thermal expansion of water, antifreeze and alcohol (S 1 )”


b. Thermal expansion of water and vegetable oil (S2)
2. Thermal expansion of air, carbon dioxide, and oxygen (SI)
3. Changes in freezing point of solution of two substanccs with varying composition: A phase diagram
63)
4. a. Changes in boiling point with amount of salt added (S2, S2)
b. Changes in boiling point with amount of sugar added (Sl)
5 . Changes in the freezing point of water with amount ethyl alcohol added (S I )
6. Determination of latent heat of fusion of water (SI)
7. The heat released during chemical reactions as a function of amount of chemical (SI)
8. a. Specific heat capacity of four metals (S3)
b. Heat capacity of four liquids and four solids (S3)
9. Relationship between density and heat capacity (SI)
10. Specific heat of water as a function of amount of sugar added 6 2 )
11. Factors affecting specific heat (S 1)
12. Relationship between temperature and time when heat is transferred to a sample of water (S3)

a SI , S2, and S3 indicate the section in which the experiment was designed

the laboratory. In other groups, such as ARR, the students quickly decided on a general
phenomenon, such as thermal expansion, then settled on one research question and spent the
remainder of the planning period on discussing details of the instrumental and material require-
ments. In all groups, however, the research question for an experiment evolved out of the
negotiations between the members of each group and the instrumental contingencies, prior
knowledge, and interaction with the teacher. Although the seed of an idea may have been
verbalized by one student, the final research question and the experimental design usually arose
out of the negotiations; or an initial idea led to negotiations with consequent changes, such that
the focus question as well as the experiment always became the product of a group (such as
CJP’s phase diagram experiment in Table I). In some cases, the negotiations led a group to a
research question very similar or equivalent to that proposed by the teacher as an alternative. In
these cases, students still considered it their own experiment.
The students in this study demonstrated that they could frame research questions and plan
investigations to find answers to these questions. From the very beginning of the course, the
students were encouraged to include in their laboratory reports spin-off questions that could
serve as the focus for the next inquiry. Another source of questions that students framed was
phenomena they had observed outside of school. But as indicated, the questions a group
investigated arose out of the interaction between the group members, the teacher, and the
context. These influences almost assured a great variety of questions. A second source for the
variation in research topics was the students’ interest in distinguishing their work from that of
neighboring groups.
Starting with the premise that we cannot teach everything of importance, Wiggins (1989)
argued that we must foster students’ ability to ask and keep questioning. In his view, a modern
curriculum should (a) equip students with the ability to learn through careful questioning, (b)
enable students to turn these questions into systematic knowledge, and (c) engage students in
important questions so that they learn to take pleasure in constructing new knowledge. We
believe that the classrooms observed in this study provided such a context of questioning. The
students’ responses on essays, o n questionnaires, and during interviews seem to validate Wig-
A CONSTRUCTIVIST PHYSICS LAB 207

Table 2
Range of Topics Discussed until a Group Reached Consensus

CJP
I . Changes in heating curves with addition of salt
2. Effect of pressure on heating curve
3. Change in boiling point with pressure
4. Measurement of freezing points of substances such as ice-water mixtures
5. Freezing point variation with addition of antifreeze
6. Boiling point variation with addition of antifreeze
7. Changes in fixed points of water with addition of salt
8. Relationship between temperature and reaction rates
9. Reaction times (unspecified substances)
10. Changes in boiling point of liquids with addition of salt
1 1 . Thermal expansion of solids
12. Thermal expansion of liquids and gases
13. Changes in freezing point with addition of antifreeze
14. Changes in freezing point with addition of salt
15. Freezing point and boiling point of pure salt
16. Freezing point of paradichlorobenzene and naphthalene mixtures
17. Freezing point of water and something else, such as antifreeze
18. (inaudible suggestion requiring four temperature probes)
19. Freezing point changes as a function of amount of salt added
20. Freezing point and boiling point changes of miscible substances
a . Salt and water
b. Water and antifreeze
c . PDB and naphthalene
MCMT
1 . Relationship between velocity and work via conservation of energy relationship (kinetic energy and
potential energy = work)
2 . Inertia experiment
3. Air resistance
(Decision was made between the end of one class, planning, and the beginning of the next,
instrumental.)
ARR
1 . Specific heat of metals
2. Thermal expansion of solid rods
3. Thermal expansion of liquids
4. Thermal expansion of gases
5. Thermal expansion of liquids

gins’ third point: The students in this study took pleasure in tackling interesting research
questions and answering them.

Narrative Explanations.
Proposals for an experiment did not always start with a research question, but sometimes
began with the description of a phenomenon. As the students discussed the phenomenon and
quantities they could measure, the relationship between two variables arose and with it the
possibility of framing a question. David gave the following description out of which developed
an experiment in which the students measured the spring constant and the relationship between
the period of a spring and the masses suspended on it.
208 ROTH

What would it be to measure, to measure/An example of it would be if you take a spring


and another on the bottom if the spring is really stretched out, if it goes really quickly
(shows a stretched spring by holding his two hands about 1 meter apart, the lower one
oscillating quickly)/if it wasn’t stretched out very much ? (hands about 20 centimeters
apart, the lower one showing a lower frequency than the first time)5

Beginning with this description, David’s group (DJMG) identified amplitude, mass supported
by the spring, spring constant, and period as variables in the phenomenon. In the discussion
following David’s proposal, DJMG decided to determine the spring constants for various
springs and then to study the relationship between mass and period. Similarly, the following
episode introduced a 18-minute discussion about a possible experiment:

Morris: You know, have you been wondering why you were flying forward in the car
when. . .
Michael: No!
Morris: But, how about we do an experiment?
Cale: Oh, you know how, The0 will love that, because he loves to wing the things on
the air track, that’s what it does, stick one of these things (card?) on a piece of play dough,
stop it on one side, it will fling itself through the photo gate. (Cale spoke very quickly,
without any pause, almost overlapping with Morris. It looks as if he knows exactly what
Morris wanted to do.)

In this episode, Morris provided a sketch of a real-life situation. Cale responded immediately by
proposing a simulation. MCMT talked a lot about the variables they wanted to measure, such as
velocity and force, and how they would measure them. However, the students did not explicitly
relate two variables, nor did they frame a focus question. In the end, MCMT decided to go with
the second experiment, which they had discussed at length.
Bruner (1986) identified two modes of thought, a paradigmatic or logico-scientific and a
narrative mode. He characterized the paradigmatic mode by its focus on formal, mathematical
systems of description and explanation. This mode employs “categorization or conceptualiza-
tion and the operations by which categories are established, instantiated, idealized and related
one to the other to form a system” (p. 12). The narrative mode, on the other hand, deals with
human or humanlike intentions and actions. It is characterized by its use of particulars of
experience and by placing this experience in a spatial and temporal context. Both of our
exemplars show these qualities. David and Morris drew on their experience to instantiate an
interesting phenomenon by providing a narrative description, accompanied by vivid gestures.
They were able to communicate to their peers both the phenomenon and a preliminary analysis
of potential relationships. This led us to use Bruner’s concept of narrative when students
described their ideas about an experiment rather than framing questions in which variables were
explicitly stated.
The fact that students do not state the variables-as in the paradigmatic mode of thinking-
has been seen as a deficit that needed to be overcome through teaching or by familiarizing
students with the context of inquiry for longer periods of time (Friedler, Nachmias, & Linn,
1990; Roth & Roychoudhury, 1993b). However, the use of narratives in problematic situations
is not singular to students, but is a powerful means to deal with complex real-world problems.
Economists, when their predictions based on economic theory fail, take to “telling stories”
(Bruner, 1986). “These narratives, once acted out, ‘make’ events and ‘make’ history. They
contribute to the reality of the participants” (p. 42). The teachers in Granott’s (1991) study used
narrative descriptions in their social constructions of explanations for the complex behavior of
A CONSTRUCTIVIST PHYSICS LAB 209

robots built in the Lego-Logo Laboratory. Similarly, narratives are used by Ilongots to respond
to the unexpected (Rosaldo, 1989) and are one form of social analysis (Marcus & Fischer,
1986). Rather than being an inferior mode of thinking, narrative analysis may be an important
step in students’ dealing with new and unexpected phenomena. Narrative analysis provides
students with an opportunity to relate the construction of new knowledge to their prior experi-
ence. Thus, this form of analysis is compatible with constructivism and should not be discour-
aged. Brown et al. (1989) agree by pointing out that narratives are used to exchange, modify,
and appropriate ideas within a culture. In their view, narratives are an essential component of
social interaction and allow access to much of the socially distributed knowledge and to exten-
sive support from the social matrix. Once students are more familiar with the objects, events,
and conceptualizations of the new phenomena, they are better prepared to frame questions in a
paradigmatic mode, that is, to frame questions in terms of independent and dependent variables.
Such a trend was previously reported in a study with eighth-graders who engaged in an open
inquiry of biomes and in a study with teachers engaged in an open-inquiry environment of the
behaviors of weird creatures of Lego-Logo design (Granott, 1991; Roth & Roychoudhury,
1993b).

Blind Alleys.
A major factor in a group’s decision for or against an experiment was its potential for
observing an effect. “Let’s do one IexperimentJ where we know that it works” became a very
important adage during the planning phase, and some of the groups consulted with the teacher
when they were too uncertain. However, based on the subjective decision whether a group could
deal with a negative finding, we did not discourage blind alleys. For example, TFNT wanted
to investigate the effect of a magnetic field on the flow of water. Because of the polar nature of
the water molecule, they expected water to flow at varying rates through a burette depending on the
presence or absence of a magnetic field. Their findings were negative. However, we used the
occasion to familiarize the students with t tests to compare the means of their flow-rate measure-
ments with and without magnetic field. To extent the experience, the students observed the
behavior of an electrically charged stream of water flowing through a magnetic field where there
was a clear effect, similar to the one the students expected in their experiment. But even in this
case, students and teacher engaged in a collaborative inquiry to deal with emerging problems of
a new and never-tried experiment. Thus, we tried to make sure that the students realized they
had learned something in spite of the negative outcome. In another blind alley, DJMG tried to
measure the heat of fusion of water by immersing a small amount of water into a bath of alcohol
at low temperature. Although they tried a number of variations in the instrumentation, they did
not succeed. Rather than viewing this experiment as a failure, we encouraged the students (a) to
regard their original idea as reasonable; (b) to use their experiments as evidence that with their
apparatus and their procedure, the latent heat could not be measured; and (c) to submit a report
that was not to be penalized for the negative outcome.
Although blind alleys are part of the daily experience of scientists, they rarely become
publicized events. If they do, it is usually in the context of research that eventually led to a
significant finding, as illustrated by the widely known search for the structure of DNA (Watson,
1968). But as the case of cold fusion showed, a widely publicized blind alley before an ultimate
success can lead to the disgrace of the scientists involved. In terms of teaching science we have
to ask ourselves whether or not we should allow students to experience such failures as part of
their science experience. For those who advocate the coverage of material and planned experi-
ences, blind alleys constitute a waste of scarce teaching-learning time. On the other hand,
210 ROTH

educators have called in recent years for authentic learning environments (Brown et al., 1989;
Wiggins, 1989). In authentic learning environments, students engage in tasks that are identical
to or closely resemble ordinary practices of the fields they study. As we pointed out above, blind
alleys are part of the ordinary practices of sciences and thus should be part of the learning
experience. Also, the experience of blind alleys would help students in constructing an under-
standing of the nature of scientific inquiry that is closer to the description provided by scientists
and sociologists of science alike. Blind alleys allow students to experience that science and the
construction of knowledge are not straightforward enterprises but tentative, a fact recognized by
only a minority of students (Solomon, 1991).

Collecting Data
Once the students had decided on a focus question and research design, they implemented
their plans. During this implementation, many problems emerged that the students had to
resolve in order to achieve their overall goal, that of answering the focus question. The students’
accomplishments in framing and resolving these problems led us to a different view of problem
solving. As we analyzed the students’ problem solving, we noticed that the decisions students
made were contingent upon the social and material setting. This led us to construct the dimen-
sion of the situated nature of decisions.
The design of multiple experiments in the same classroom sometimes gave rise to confusing
moments during the first 15 minutes of a data-collection period. Many students were looking for
pieces of instrumentation and materials, hurrying all over the classroom. The teacher was at the
center of the room, assisting students in locating equipment or indicating which existing appara-
tus from other classes could be disassembled for their own purposes. Thereafter, the students by
and large worked independently, while the teacher circulated from group to group to encourage
students, and to suggest alternative instruments, materials, or measurements. When the labora-
tory assistant was available, much of this initial confusion was avoided as she helped in
organizing the distribution of the materials. Also, when students knew beforehand that they had
special requests, the materials or equipment were organized from other labs, constructed by the
school’s shop, or purchased if feasible.

A Direrent View of Problems.


There existed a substantial difference between problems when they were set by the textbook
or the teacher and when they were framed by the students themselves. When the problems were
set, either as word problems from the textbook or as questions to be experimentally answered,
the students’ first problem was that of constructing a meaningful representation of the required
task. Then, students could solve the problem as they had reconstructed it, which was not
necessarily how the teacher had envisioned. On the other hand, when the students framed their
own problems, solution processes and products were often entailed in the problems. As indi-
cated in the section on framing research questions, the negotiation of the events’ meaning and
’the research question often entailed each other.
We found such entailment of problem formulation and resolution in many instances as
students developed their apparatus and collected the data. For example, DMT was working on
problems involving springs, including Hook’s law, harmonic oscillation, and damped oscilla-
tion. The videotapes showed the students muttering to themselves or to their partners phrases
such as, “How we’re gonna hook that is the problem” or “We have a small problem, we can’t
hook that here?” As soon as they had framed the problem in this way, the students went about
A CONSTRUCTIVIST PHYSICS LAB 21 I

constructing a support mechanism for their spring. In another instance illustrated in the follow-
ing vignette, CJP prepared an experiment on the frictionless air track:

CJP test various carts, close the unused outlets of the air pump, and level the air track.
Pete searches for something and discovers a leak in the plate at the end of the track. He
communicates his finding to Carl. Shortly after, the two use Scotch tape to tape the leak.
They change to a wide plastic tape when it becomes apparent that the narrow Scotch tape
fails to seal the leak well enough.

By framing the problem as “not enough air” rather than as “faulty cart,” “bent track,” or “dirty
surface,” Pete could immediately envision a solution to the problem. Or rather, by finding a
solution, escaping air, Pete found the problem, friction between the air track and the cart. This
phenomenon of finding solutions that entail the identification of a problem is not unfamiliar to
corporate decision makers (March, 1991) or scientists (Knorr-Cetina, 1981a). Thus, the prob-
lems the students framed and resolved were problematic in Wheatley’s (1991) sense; that is, the
situations were meaningful as problems. Textbook problems, on the other hand, were often
meaningful only to the extent that students try to determine “how to get a good grade, to please
the teacher, [or] to avoid looking stupid” (p. 15).
The students indicated during interviews that they were quite aware of the difference
between the open-inquiry and traditional laboratory exercises they conducted in their chemistry
course. Most students did not like the cookbook approach of traditional laboratories because the
purpose of most steps remained hidden from them. Thus, they completed a chemistry laboratory
exercise without knowing why they took each step. Because these exercises also focused on the
verification of established laws and theories, students often altered their data to conform with
the expected values. In the open-inquiry physics laboratory, on the other hand, they knew each
step and why they were taking it: they had designed the experiment. The students indicated that
they saw the problems emerging out of the context of the inquiry as challenges to be resolved
with the locally available physical and social resources. In some cases, they even decided to
abandon the line of inquiry they had taken. This, however, was not a weakness but a reflection
of everyday problem solving. According to Lave (1988), everyday problem solving is so
efficient because people own their problems so that they are free to transform, solve, or resolve a
problem, or abandon them in favor of other options. Such cognitive flexibility to abandon
designs, plans, and solutions and to search for better alternatives is linked to one’s control over
the problems at hand:

[In real life] individuals experience themselves as in control of their activities, interacting
with the setting, generating probkms in relation with the setting and controlling the
problem-solving processes. . . In contrast, school and experiments create contexts in
which children and “subjects” experience themselves as objects, with no control over
problems or choice about problem-solving processes (Lave, 1988, pp. 69-70).

The Situated Nature of Decisions.


The choices students made affected the data they collected and the conclusions they could
draw from the experiment. In the following episode, CJP decided to measure the viscosity of
several liquids for their hydrodynamics experiment.

Jim: Should we measure the viscosity in milliliter/minute?


Carl: Just fill it [funnel] up to a certain (indicates height), the same [level] for each
212 ROTH

liquid. Then set it over the sink, or whatever (holds it over the graduated cylinder; shows
finger on bottom of funnel) Measure the time it takes to fall out.
WMR: Or we could just take burettes and just time a certain amount, how long it
takes.
Jim: But this one [funnel] will be faster.
WMR: But it [measurement] will be more exact.

Although the teacher’s reasoning seemed to be satisfactory to Jim at this point, he had not given
up on his idea of using the funnel. Using the funnel would also provide him with another
opportunity, namely, that of measuring the viscosity of all the liquids relative to air. Because
they lacked an absolute standard for the viscosities of liquids, Carl and Pete established the
necessity of determining the relative viscosities. Carl indicated that the air could not be used as
the substance against which to assess all the liquids. But Jim presented his idea about using the
funnel with great great excitement:

Fill the funnel with a liquid. Put a sheet of paper on top of surface. Then turn the funnel
upside down and let the liquid fall out (wide side of funnel). Air has to come in through
the narrow side. The time it takes the liquid to fall out is the time it takes air to fill the
funnel through the narrow side (this gives indication of relative viscosity of other sub-
stances to that of air).

Because they found a burette, CJP discarded the plan of using the funnel and decided to settle
for a measurement of viscosities relative to water. Thus, a decision at one point simultaneously
became a contingency for future states of the experiment. Given the dependence of the final
experiment on so many local selections and negotiations we could not even conjecture about the
outcomes if the group had used the funnel, and tried to measure the viscosities with respect to
air. This episode also illustrates how groups collaboratively constructed their experiments.
Through this discussion, both the need for establishing relative viscosities and the function of air
as a standard was incorporated into the shared knowledge. All three contributed to this collab-
orative construction, and the final experiment and knowledge could not be attributed to any
single student of the group. Students themselves recognized the effect of these local decisions on
the overall outcome of their experiment: “Throughout the procedure of a lab we are constantly
making decisions that will affect the outcome. Overall, a lab requires a great deal of planning,
precision, interpreting, and thinking, and each time we learn more about the process and
effectiveness of experimentation” (Allan).
In order to deal with constantly arising problems, the students adapted measurement instru-
ments, tools, and materials in various creative ways to make their experiment work. The
specific uses arose out of the context and through the interaction among peers. This view of
working in a laboratory invokes the image of a tinkerer, who makes use of whatever he finds
around him. In the process, his products will be highly affected by the local character of his
choices. Thus, “the mechanisms ruling the progress of research are more adequately described
as successful ‘tinkering’ rather than as hypothesis testing or cumulative verification” (Knorr,
1979, p. 350). Accordingly, the processes of understanding, interpreting, interacting, and com-
municating in the local and idiosyncratic context of specific laboratories are both conditions and
actual characteristics of ongoing research. The notion of the tinkerer is quite similar to that of
the bricoleur (Turkle & Papert, 1991). On the basis of their findings with Grade 4 students
programming instructional lessons, Turkle and Papert contrast the bricoleur and the person
following plans:
A CONSTRUCTIVIST PHYSICS LAB 213

Bricoleurs use a mastery of associations and interactions. For planners, mistakes are
missteps; bricoleurs use a navigation of midcourse directions. For planners, a program is
an instrument for premeditated control; bricoleurs have goals but set out to realize them in
the spirit of a collaborative venture with the machine. For planners, getting a program to
work is like “saying one’s piece”; for bricoleurs, it is more like a conversation than a
monologue (p. 169).

Similar observations and conclusions were reported by sociologists of scientific laboratories


(Garfinkel, Lynch, & Livingston, 1981; Latour & Woolgar, 1979; Zenzen & Restivo, 1982) and
by ethnographers reporting on Athabascan hunting, Ilongot visiting, and Micronesian naviga-
tion (Brody, 1982; Hutchins, 1983; Rosaldo, 1989).

Analyzing the Data


Once students had completed the instrumental part of their experiment, they proceeded to
plotting, transforming, and curve fitting the data they had collected. From the very beginning of
the course, the teacher had required that students use one or more of the various software
packages available for the graphical analysis of data. Because of a lack in the mathematical
preparation of the students, there often arose the need for introducing new mathematical con-
cepts. In order to fit the data students had collected with a curve, the teacher assisted individual
groups in learning about exponential, logarithmic, and trigonometric functions; or he assisted
students in using log-log, sine-sine, or lly-llx transformations to transform the data so that they
could be regressed by linear functions.
Our analysis of the students’ interpretation of the data they had collected and the graphs
they had produced reveaIed a great concern for understanding and meaning. The students
confirmed this construction in individual interviews during which they also highlighted the
crucial role of interpretation to scientific inquiry. Although data interpretation usually followed
the completion of the data-collection phase, we observed that our students engaged in cycles of
interpretation, redesigning setup and procedures, and collecting data. Thus, students were
engaged in processes of iterative interpretation of the phenomena they studied.

A Concern for Understanding and Meaning.


The observations, videotapes, and laboratory reports made it very clear that students were
concerned with understanding. This understanding arose out of the negotiations around the
graphs they had generated and the questions raised by the teacher. Such concern for meaning
was evidenced in the following transcript, in which students discussed the reasons for the
sudden drop in a temperature curve:
Michael: That’s when we added salt (pointing to a bend in the curve).
Morris: Because all the ice has to be melted before the temperature can
Michael: Remember we added the salt here (pointing to the bend in the curve).
Morris: But still, water has to stay stationary around freezing, should stay until all
the ice has melted, then as soon as all the ice/all the energy is going so that the ice can
melt and as soon as all the ice has melted, which is at this point (points to bend in freezing
curve).
Michael: We added the salt!
Morris: Its not melted here [upper part of freezing curve], so the ice is being frozen,
and all the energy it looses water is being cooled, so that the water converts to ice
(pointing to flat part).
214 ROTH

One can sense the urgency this problem had for the students. It was important to Michael and
Morris to establish the meaning of each part of the freezing curve. In such discussions and
negotiations students constructed new knowledge as a group; the knowledge was then available
to be appropriated by each individual. Our observations concerning these discussions and
negotiations to student learning were corroborated by the evidence gathered from students’
essays and interviews. With the exception of four individuals (of the 46), all students empha-
sized the importance of the group discussion. Here, according to the students, they learned to
incorporate different viewpoints into their own understanding, to elaborate their own under-
standing because they had to defend their own ideas, or to compound each other’s explanations
to produce a better report. Students also indicated that they learned most through the interpreta-
tions of the data they collected. Thus, in their view, “Physics is experimental, as I often learn
principles or formulas out of the interpretation of our data” “Physics relies on interpretation of
results. Interpretation is the actual learning process, and varies from person to person” and “The
lab provides us with the chance to interpret our own ideas which aids us attaining more
knowledge.”
The students’ concern for meaning when they interpreted the d a a and graphs, and for
constructing sensible knowledge claims, was apparent from the laboratory reports they submit-
ted. They showed concern not only for the meaning of the actual data, but also for understand-
ing the transformation to which they submitted these data. The following excerpt from one of
CAT’s reports illustrates these students’ concern for meaning:

In setting out to study density and wavelength, the graphs are crucial to the experiment.
Since the normal function f(h) must have an asymptote (since neither frequency nor
wavelength can be 0), we must graph frequency versus lih. In graphing the results of
straight lines, it is important to note that the slope of the lines is the velocity (because f =
vih). . . The final graph, and also the most important, is velocity versus density. . . The
higher the density, the slower the velocity, and therefore also the wavelength. However,
the slope is -0.157~. and therefore, though wavelength is affected. it is not drastically so.

After a transformation of the initial data followed by a regression, CAT found the velocities of
the waves in three materials. They then plotted these velocities versus the densities of the
liquids. The concerns for understanding each individual step paid off in that they realized that
the four graphs they produced were not of equal importance for answering their research
question (“The final graph, and also the most important, is velocity versus density”). In order to
answer their focus questions, CAT needed to evaluate the velocity-density graph, from which
they could find the relationship between wavelength and density via the wave equation. The last
two lines of the excerpt indicate that students did not simply accept relationships because of a
high correlation, r = 0.98, but they also showed a concern for the magnitude of the effects
which they observed.
In their claims, CAT linked these results to a possible applicationhelevance in real life. “In
dealing with the highly toxic PCBs, since they are so dense, it is not likely to produce waves,
and is more readily poured than alcohol, which could make waves and splash everywhere.”
Such linking to everyday experience, or to subject matter in other classes, increased the cross
references students constructed in their knowledge, and increased the meaningfulness of the
results of the experience in the laboratory. These findings were supported by the importance
students gave to the construction of links from the everyday life to the lab experience, and vice
versa. Thus, students indicated that “By carrying out the everyday life situations in the lab, we
figure out patterns and relationships which we bring to [the] understanding of our w o r l d or
A CONSTRUCTIVIST PHYSICS LAB 215

“[We] use information and ideas from everyday situations in class to help with experiments and
to reinforce the nature of the concepts that we are learning.”

Iterative Interpretutions.
The interpretations of graphs did not only begin after the data had been completely col-
lected. Both the collection and interpretation of data and the redesign of apparatus often went
hand in hand, one affecting the other, as the following vignette illustrates:

MCMT had collected a velocity-time graph for a rubber stopper falling through water. The
graph indicated that, after the stopper had reached terminal velocity (term v) asymp-
totically, term V suddenly jumped to a lower value. The students repeatedly collected data
and saw that the pattern was consistent, but they could not explain it. After changing the
timing apparatus, they achieved a similar result. Finally, Theo realized that the stopper
tipped over from tapered end first to the butt end. When they fixed the stopper so that it
could not turn over during its descend, term V did not change.

We frequently observed such integration of redesigning experiments, collecting data, interpret-


ing, formulating models, and so forth. Thus, the student behaviors observed in this laboratory
are markedly different from that in cookbook laboratories. When they do cookbook laboratories,
students are not concerned with meaning but with completing the exercise by following the step-
by-step procedures (Gallagher & Tobin, 1987). Our students indicated that the very nature of
their chemistry cookbook laboratories prevented them from understanding why they completed
a particular procedure. They were not interested in simply collecting data and thereby complet-
ing a task. But they were genuinely concerned with finding an answer to the question they had
framed themselves. Their interests were thus epistemic, concerned with learning and knowing,
rather than being pragmatic, concerned with the completion of tasks. In the open-inquiry
physics course, the students knew what they wanted to measure; they knew what data they had
to collect; and they often had hunches as to what data they should expect. Thus, when the data
and graphs did not confirm to their expectations, or when they showed irregularities, the
students changed part of the setup or the data collection. They studied the effect of these changes
on the data and graphs they collected to construct a better understanding of the phenomenon. In
the end, the data and graphs were constructed interactively between the students, the environ-
ment, and the goal students had specified for themselves. Such a dialectic between people, their
activities, and the environment is characteristic of everyday, situated cognition (Lave, 1988).
Problems that arise out of individuals’ purposeful activities are not structured as ends in them-
selves, but constitute dilemmas or problematics with which the problem solver is personally
engaged. Such a personal engagement is critical for constructivist learning environments.
The evidence furnished here supported one of our previous constructions, that the descrip-
tion of laboratory activity in terms of isolated science process skills may be inappropriate (Roth
& Roychoudhury, 1993a). The students did not identify variables separate from interpreting
graphs or redesigning an experiment. Rather, students were engaged in a holistic resolution of
problems. Our contention is supported by evidence in the literature of various fields. Human
pursuits, be they scientific, professional (such as architecture), or everyday activity (shopping or
dairy factory arithmetic), are much too complex to be decomposed into a small set of isolated
skills (Hutchins, 1983; Knorr-Cetina, 1981a; Latour & Woolgar, 1979; Lave, 1988; Schon,
1983; Scribner, 1984). This led Lave (1988) to claim that “if relations among activity, setting
and processes of dilemma-resolution are dialectically constituted, then it is not possible to
216 ROTH

separate means of problem-solving activity from its ends. Gap closing processes unite means
and ends, transforming both in the process into means-end and ends-mean, a distinction without
difference” (p. 175).

Conclusion
Three aspects of the planning stage of students’ investigations are presented here, framing
research questions, narrative descriptions, and blind alleys. In each of these aspects it became
quite clear that the students in an open-inquiry laboratory used reasoning modes similar to those
that appear during everyday practices of scientists and nonscientists alike.
The most important result of this part of the study of this class was that students learned to
frame appropriate research questions. Framing problems is an important skill in everyday
environments, where problems often are undefined or ill-defined, in contrast to textbook prob-
lems students encounter in schools, which are well defined and of extremely limited context.
This skill is so important that Schon (1983) considers it central for effective problem solving.
Thus, educators have called for making exploration through the pursuit of questions a key
ingredient of teaching (Collins et al., 1989; Wiggins, 1989). This implies that the essential
questions must derive from pupils. As students pursue these questions of their own interest, they
not only learn to gain pleasure from inquiry, they also gain ownership over problems and
solutions. For if questions are endogenous to the constitution of problems, they will not be
structured as an end in themselves, but will invite the emotional engagement of the questioner.
This ownership implies that “individuals experience themselves as in control of their activities,
interacting with the setting, generating problems in relation with the setting and controlling
problem-solving processes” (Lave, 1988, p. 69). In traditional school settings, on the other
hand, students experience themselves as objects with no control over problems and solutions.
During their attempts in formulating questions, students used narrative descriptions to
analyze a complex situation before they could identify variables to be correlated. Though
anathema to present school practices, both are important modes of thinking and analyzing
complex situations. Thus, they must be promoted, not inhibited. Narrative descriptions are
important vehicles for the construction of individual representations, for mediating taken-to-be
shared understanding, and for establishing a functioning classroom community. They provide
access and add to the collective wisdom of this community (Brown et al., 1989).
The participants in traditional problem-solving studies solve well-defined problems from
well-structured knowledge domains. Some best solutions to these problems are used as norms
against which the performance of experts and novices are compared. In this study, however,
well-structured problems did not exist. What was to become a problem always arose from the
interplay between the participants, the activity, and the context. These problems were unpredict-
able both in their form and their content. Our participants had to frame these problems on their
own before they could resolve them; that is, they had to impose their own meaning to structure
the phenomena. Because these problems arose in the context of an investigation the students had
designed on their own, they had the character of a problematic situation. From a constructivist
view, such problematic situations provide favorable conditions for learning, because the prob-
lem solver is facing conditions for which no known procedures are available (Wheatley, 1991).
Such problem solving in ill-structured domains (from a student’s perspective) may lead to blind
alleys, unexpected results that constitute insurmountable barriers to finding answers for research
questions. A supportive classroom culture needs to be provided so that students can learn to deal
with such situations without experiencing failure and considerable frustration.
The notion of a problem-rich environment begs the discussion of topic boundaries. Readers
A CONSTRUCTIVIST PHYSICS LAB 217

may have detected in our excerpts and vignettes that many investigations involved concepts
from varying topical areas not normally connected in textbook presentations. In traditional
science classrooms, teaching focuses on distinguishable topics. Thus, the students learn about
density in one unit, about motion in a second, and about waves in a third unit. In this open-
inquiry laboratory, such topic boundaries did not exist. Our students needed to understand
density, viscosity, and buoyancy when they studied the motion of objects falling through a
liquid. They investigated each of these topics in order to achieve their overall goal of under-
standing the motion of objects in liquids. In another investigation on wave propagation one
student group needed to understand the interaction of density, viscosity, temperature, and the
velocity of surface waves in different substances. Again, the students did not stay within the
topic boundaries usually set by textbooks. Our interviews revealed that students were quite
aware of the missing connections between topics in textbooks. They felt that their inquiries
provided a coherence to studying physics which textbooks did not provide.
In our search for problem-solving achievements comparable to those we observed, we
found a resemblance between sociologists’ descriptions of scientists at work and anthropolo-
gists’ descriptions of everyday cognition. The chief metaphors that seemed to apply to our work
were those of the tinkerer or bricoleur and that of the dialectic of individual, setting, and social
context. The work of the tinkerer or bricoleur is characterized by its indexical logic, its situa-
tionally contingent and circumstantial nature, the local nature of results of an inquiry, and its
implied opportunism (Knorr-Cetina, 1981a). The problems the tinkerer thus faces arise out of
the context and in relation to the individual and its social context. This leads to a conception of
problems and problem solving that differs from traditional problem-solving research. Rather
than being prefabricated with a definite structure, problems that arise out of engaged activity
develop a structure determined by the dialectic between its constituting components, the indi-
vidual, the setting, and the social context. As we pointed out earlier, in this conceptualization
the solution process cannot be divided into separate means and ends as information-processing
psychologists want to do: the formulation of problems and solutions (processes and products)
were inseparable and could occur in any order. Thus, solutions that emerge out of the aforemen-
tioned dialectic process integrate means and ends into holistic means-ends or ends-means
processes. Because in this process the goals are endogenous to the constitution of problems,
problems are owned by individuals, rather than given in a normative, decontextualized form by
external agents. The solution process becomes meaningful and purposeful and constitutes a truly
constructivist activity. As a result, we agree with those who advocate that students should be
learning from self-directed projects rather than from isolated problems (Collins et al., 1989;
Harel & Papert, 1991). In these projects, students can frame their own problems guided by
phenomena they find interesting. They can then develop their own solutions to these problems
within the constraints of the embedding context, giving them ownership over the generated
knowledge. The task for educators then consists (a) in providing for problem-rich learning
environments in which students can frame interesting tasks, and (b) in providing the necessary
support system so that students can succeed. In this process, it might be necessary to find tools
such as computers, which can be used to bridge the cultures of school, everyday, and scientific
thinking (Harel & Papert, 1991; Hawkins & Pea, 1987).
As we discussed the students’ work in each of the three phases of experimenting, we
compared their work with that of scientists, particularly pointing out similarities between the
two. According to Cobb, Wood, and Yackel (1991) such similarities may actually arise out of
the epistemological stance from which the study was conducted. That is, once investigators are
concerned with the way science is constructed and used by human beings, and once science is
viewed as a creative human activity, “analogies between the activity of the scientist or mathe-
218 ROTH

matician and the school student become apparent” (p. 24). However, we do not want to neglect
pointing out the essential differences between the two. These differences seem to lie in the
conceptual background rather than in a rationality specific to scientific work (Knorr-Cetina,
1981a).
First, between the students in this study and scientists, there are considerable differences in
the relative conceptual backgrounds they bring to their work. Second, there are differences in
the stock of embodied laboratory practices, the vulgar competence, and familiar efficiency
scientists exhibit on the work bench (Garfinkel et al., 1981; Knorr-Cetina & Amann, 1990).
Thus, students have had little opportunity-six months in one course versus years in research
laboratories-to develop experiential meanings in terms of laboratory practice which are so
important to success in making a preconceived solution work; they have had only a short time to
become familiar with the modes of discourse and thinking accepted in the field of physics; and
they have had little practice in dealing with the contingencies of a developing experiment. In
terms of educational practice one might ask whether students will be able to acquire such
experiential meanings in traditional cookbook laboratories. As most practicing teachers know,
cookbook laboratories prize the replication of procedures and verification of well-known laws
and theories so that students often adjust their data accordingly. On the other hand, “exact
replication is treated as something that one might seek to demonstrate very occasionally”
(Mulkay & Gilbert, 1986, p. 35). Rather, it is the systematic pursuit and interpretation of
instrumental variation through which the experimentally constructed objects of science are
stabilized and reified.
There is a third, crucial difference between scientists and the students. The students’
constructions were guided by a teacher who had already constructed much scientific knowledge
for himself, who is a member of a larger community of scientifically literate individuals, and
who can be thought of as a representative of the scientific community (Cobb et al., 1991;
Roschelle, 1990). Cobb et al. indicated that the interactions between students and teacher
constrain the construction of knowledge “in the same way as empirical and conceptual anoma-
lies constrain the scientist’s construction of theoretical knowledge” (p. 38). These constraints
could help students to develop forms of discourse that are closer to those of canonical science.
The notion of legitimate peripheral participation (Lave, 1992; Lave & Wenger, 1991) provided
us with another framework for understanding the differences between scientists, science teach-
ers, and students. Only trained scientists who are actively engaged in scientific research fully
participate in scientific discourse. Teachers, through their training, in-service, teaching prac-
tice, and daily encounters of science through the media engage in this discourse more at the
periphery. Students have an even more peripheral status in this discourse. However, by complet-
ing a trajectory from school to university to postdoctoral studies, students increasingly engage in
scientific discourse, which may lead to full participation in the community of practicing scien-
tists.

Educational Implications
We presented here a high school laboratory course that in many respects is different from
traditional science instruction. Although we could have tried to interpret students’ actions in
terms of past conceptualizations in science education such as science process skills, the frame-
work we developed and used here seemed to be more suitable to the classroom we investigated.
This framework, which arose out of current work in epistemology, cognitive anthropology, and
sociology of knowledge, also permitted us to refer to the work of scientists from a constructivist
perspective. The picture that emerged in our description is that of (a) a laboratory where plans
A CONSTRUCTIVIST PHYSICS LAB 219

are inherently underdetermined, permitting multiple courses of action; (b) a laboratory where
experiments that evolve as plans are implemented; (c) a view of problems as arising out of the
locally contingent description of actors; (d) a view of problem solving that is essentially
constructivist and suitable for a tinkerer; and (e) a view of a laboratory in which understanding
and meaning were central to the students’ activities. In this sense, our results are consistent with
several trends in contemporary educational discussion such as “situated learning,” “total learn-
ing environments,” “self-directed project activity,” “authentic practice,” and “collaborative
learning” (Collins et al., 1989; Hare1 & Papert, 1991; Schon, 1983; Wiggins, 1989).
The present results have important implications for teaching science. In our view, it is of
utmost importance that students be provided with a learning environment (a) in which they learn
to find and frame problems by imposing their own meaning on ill-defined and open-ended
conditions; (b) in which they develop the ability to negotiate constraints with others and with the
setting; (c) in which they pursue solutions by means best suited to their thinking styles; (d)
which provides challenging tasks with high motivational values; (e) which provides intellectual
support through collaborative problem solving; and (f) which encourages the construction of
knowledge and negotiation of meaning. In such a learning environment, evaluation procedures
will have to change to take account of the open nature of the inquiry situation. Despite great
effort, we cannot expect that students will always arrive at the claims of canonical science.
Rather, evaluation has to be process oriented, focusing on (a) the contributions of individuals to
the inquiry, (b) whether claims are viable in terms of the data students collected, (c) the
creativity of research questions, and (d) the skills of solving problems in the process of finding
answers.
Although this study has provided some answers with respect to planning, collecting data,
and interpreting data in a constructivist laboratory environment, many more questions remain to
be answered, such as, “How do highly structured laboratory environments change the cognitive
demands on students as compared to the present, unstructured environment?” “Which kinds of
cognitive support do highly structured laboratory groups provide for the collaborative construc-
tion of meaning and individual appropriation that are unavailable to unstructured groups?” and
“What are the representations students construct in such a laboratory environment compared to
those they construct in regular classrooms‘?’’

Notes
The author thanks Anita Roychoudhury for her help and support throughout the data collection and
interpretation phases.
We used pseudonyms throughout the manuscript.
* Groups are identified by a string composed of the first initials of each member: CJP (Carl, Jim, and
Pete); ARR (Allan, Ron, and Rex); TFNT (Tom, Fred, Nicki, and Taylor); MCMT (Michael, Cale,
Morris, and Theo); DJMG (David, Jonathon, Marcus, and George); DMT (Dan, Mick, and Ty); CAT
(Cory, Alan, and Timothy).
This example has been discussed in greater detail elsewhere (Roth, 1990; Roth, in press).
We provided elsewhere a detailed analysis of the processes by which the students arrived at a
consensus (Roth & Roychoudhury, 1993a).
We used the following conventions for transcription: (a) Square brackets to add words that would
facilitate the comprehension of the transcript: “It [light] consists of quanta,” (b) parentheses to indicate
nonverbal cues and action (begins to draw sine waves), (c) parentheses and a question mark (card?) to
indicate uncertain hearing and that the word may have been card, and (d) a slash to indicate a break in the
thought expressed.
220 ROTH

References

Anderson, J.R. (1985). Cognitive psychology and its implications. San Francisco, CA:
Freeman.
Brody, H. (1982). Maps and dreams. New York: Pantheon Books.
Brown, J.S., Collins, A., & Duguid, P. (1989). Situated cognition and the culture of
learning. Educational Researcher, 18(2), 32-42.
Brown, J.S., Rubinstein, R., & Burton, R. (1976). Reactive learning environment for
computer assisted electronics instruction (BBN Report NO. 33 14). Cambridge, MA: Bolt,
Beranek, and Newman, Inc.
Bruner, J.S. (1986). Actual minds, possible worlds. Cambridge, MA: Harvard University
Press.
Burrell, G., & Morgan, G. (1985). Social paradigms and organizational analysis. Hants:
Gower Publishing.
Carraher, T.D., & Schliemann, A. ( 1 982). Computation routines prescribed by schools:
Help or hindrance. Paper presented at the NATO conference on the acquisition of symbolic
skills, Keele, England.
Cazden, C.B. (1986). Classroom discourse. In M.C. Wittrock (Ed.), Handbook for re-
search on teaching (3rd ed., pp. 432-463). New York: Macmillan.
Cobb, P., Wood, T., & Yackel, E. (1991). Analogies from the philosophy and sociology for
understanding classroom life. Science Education, 75, 23-44.
Cole, M. (1985). The zone of proximal development. In J.V. Wertsch (Ed.), Culture,
communication, and cognition: Vygotskian perspectives (pp. 146- 161). Cambridge: Cambridge
University Press.
Collins, A., Brown, J.S., & Newman, S . (1989). Cognitive apprenticeship: Teaching the
craft of reading, writing, and mathematics. In L. Resnick (Ed.), Cognition and instruction:
Issues and agendas (pp. 453-494). Hillsdale, NJ: Lawrence Erlbaum Associates.
Costenson, L.J., & Lawson, A.E. (1986). Why isn’t inquiry used in more classrooms. The
American Biology Teacher, 48, 150- 158.
Eylon, B.S., & Linn, M.C. (1988). Learning and instruction: An examination of four
perspectives in science education. Review of Educational Research, 58, 25 1-301.
Friedler, Y., Nachmias, R., & Linn, M.C. (1990). Learning scientific reasoning skills
in microcomputer-based laboratories. Journal of Research in Science Teaching, 27, 173-
191.
Gallagher, J.J., & Tobin, K. (1987). Teacher management and student engagement in high
school science. Science Education, 71, 535-555.
Garfinkel, H. ( 1 967). Studies in ethnomethodology. Englewood Cliffs, NJ: Prentice-Hall.
Garfinkel, H., Lynch, M., & Livingston, E. (1981). The work of a discovering science
construed with materials from the optically discovered pulsar. Philosophy of the Social Sci-
ences, 11, 131-158.
Gergen, K.J. (1985). The social constructivist movement in modem psychology. American
Psychologist, 40, 266-275.
Goodman, N. (1978). Ways of worldmaking. Indianapolis, IN: Hackett Publishing Co.
Granott, N. (1991). Puzzled minds and weird creatures: Phases in the spontaneous process
of knowledge construction. In 1. Harel & S . Papert (Eds.), Constructionism: Research reports
and essays, 1985-1990 (pp. 295-309). Norwood, NJ: Ablex.
Harel, I . , & Papert, S. (1991). Software design as a learning environment. In I. Harel &
S . Papert (Eds.), Constructionism: Research reports and essays, 1985-1990 (pp. 41 -84).
Norwood. NJ: Ablex.
A CONSTRUCTIVIST PHYSICS LAB 22 1

Hawkins, J., & Pea, R. (1987). Tools for bridging the culture of everyday and scientific
thinking. Journal of Research in Science Teaching, 24, 291-307.
Heller, J.I., & Reif, F. (1984). Prescribing effective human problem-solving processes:
Problem description in physics. Cognition and Instruction, I , 177-216.
Hofstein, A., & Lunetta, V.N. (1982). The role of the laboratory in science teaching:
Neglected aspects of research. Review of Educational Research, 52, 201-2 17.
Hutchins, E. (1983). Understanding Micronesian navigation. In D. Gentner & A.L. Stev-
ens (Eds.), Mental models (pp. 191-225). Hillsdale, NJ: Lawrence Erlbaum Associates.
Knorr, K.D. (1979). Tinkering toward success: Prelude to a theory of scientific practice.
Theory and Society, 8, 347-376.
Knorr-Cetina, K.D. (198 la). The manufacture of knowledge: An essay on the constructivist
and contextual nature of science. Oxford: Pergamon Press.
Knorr-Cetina, K.D. (1981b). Social and scientific method or what do we make of
the distinction between the natural and the social sciences? Philosophy of Social Sciences, 11,
335-359.
Knorr-Cetina, K., & Amann, K. (1990). Image dissection in natural scientific inquiry.
Science, Technology, & Human Values, 15, 259-283.
Laboratory of Comparative Human Cognition ( 1983). Culture and cognitive development.
In W. Kessen (Ed.), Mussen’s handbook of child psychology: Vol. 1. History, theory, and
method (4th ed., pp. 295-356). New York: John Wiley & Sons.
Larkin, J.H., McDermott, J., Simon, D.P., & Simon, H.A. (1980). Expert and novice
performance in solving physics problems. Science, 208, 1335- 1342.
Latour, B. (1987). Science in action: How to follow scientists and engineers through
society. Milton Keynes: Open University Press.
Latour, B., & Woolgar, S . (1979). Laboratory life: The social construction of scienti9c
facts. Beverly Hills, CA: Sage Publications.
Lave, J. ( 1977). Tailor-made experiments and evaluating the intellectual consequences of
apprenticeship training. Quarterly Newsletter of the Institute for Comparative Human Develop-
ment, 1(2), 1-3.
Lave, J. (1988). Cognition in practice: Mind, mathematics and culture in everyday life.
Cambridge: Cambridge University Press.
Lave, J. (1992, April). Legitimate peripheral participation. Paper presented at the annual
meeting of the American Educational Research Association, San Francisco, CA.
Lave, J., Murtaugh, M., & de la Rocha, 0. (1984). The dialectic of arithmetic in grocery
shopping. In B. Rogoff & J. Lave (Eds.), Everyday cognition: Its development in social context
(pp. 67-94). Cambridge, MA: Harvard University Press.
Lave, J., & Wenger, E. (1991). Situated learning: Legitimate peripheral participation.
Cambridge: Cambridge University Press.
Lincoln, Y.S., & Guba, E. (1985). Naturalistic inquiry. Beverly Hills, CA:,Sage.
March, J.G. (1991). How decisions happen in organizations. Human-Computer Interac-
tion, 6 , 95-117.
Marcus, G.E., & Fischer, M.M.J. (1986). Anthropology as a cultural critique: An experi-
mental moment in the human sciences. Chicago, IL: University of Chicago Press.
McArthur, D., Stasz, C., & Zmuidzinas, M. (1990). Tutoring techniques in algebra.
Cognition and Instruction, 7 , 197-244.
Mulkay, M., & Gilbert, G.N. (1986). Replication and mere replication. Philosophy of the
Social Sciences, 16, 21-37.
Newman, D., Griffin, P., & Cole, M. (1989). The construction zone: Workingfor cognitive
change in school. Cambridge: Cambridge University Press.
222 ROTH

Novak, J.D., & Gowin, D.B. (1984). Learning how to learn. Cambridge: Cambridge
University Press.
Rogoff, B . ( 1990). Apprenticeship in thinking: Cognitive development in social context.
New York: Oxford University Press.
Rosaldo, R. (1989). Culture & truth: The remaking of social analysis. Boston, MA: Beacon
Press.
Roschelle, J. ( 1990, April). Designingfor conversations. Paper presented at AAAI Sympo-
sium on Knowledge-Based Environments for Learning and Teaching, Stanford, CA, March
1990 and at AERA Symposium on Dynamic Diagrams for Model-Based Science Learning,
Boston, MA.
Roth, W. -M. ( 1990, April). Collaboration and constructivism in the science classroom.
Paper presented at the annual convention of the American Education Research Association,
Boston.
Roth, W.-M. (in press). Construction sites: Science labs and classrooms. In K. Tobin (Ed.),
Constructivism and the teaching and learning of science and mathematics. Washington, DC:
American Association for the Advancement of Science.
Roth, W.-M., & Roychoudhury, A. (1993a). The concept map as a tool for the collabora-
tive construction of knowledge: A microanalysis of high school physics students. Journal of
Research in Science Teaching, 30, 503-534.
Roth, W.-M., & Roychoudhury, A. (1993b). The development of science process skills in
authentic contexts. Journal of Research in Science Teaching, 30, 127- 152.
Schliemann, A.D., & Acioly, N.M. (1989). Mathematical knowledge developed at work:
The contribution of practice versus the contribution of schooling. Cognition and Instruction, 6 ,
185-22 1 .
Schon, D.A. (1983). The rejective practitioner: How professionals think in action. New
York: Basic Books.
Scribner, S . (1984). Studying working intelligence. In B. Rogoff & J. Lave (Eds.), Every-
day cognition: Its development in social context (pp. 9-40). Cambridge, MA: Harvard Univer-
sity Press.
Shulman, L.S., & Tamir, P. (1973). Research on teaching in the natural sciences. In
R.M. W. Travers (Ed.), Second handbook of research on teaching (pp. 1098- 1148). Chicago,
1L: Rand McNally.
Solomon, J. (1991). Teaching about the nature of science in the British national curriculum.
Science Education, 75, 95-103.
Stake, R.E., & Easley, J.A. (1978). Case studies in science education (Vols. I and 2).
Urbana: Center for Instructional Research and Curriculum Evaluation and Committee on Cul-
ture and Cognition, University of Illinois at Urbana-Champagne.
Strauss, A.L. ( 1987). Qualitative analysis for social scientists. New York: Cambridge
University Press.
Suchman, L.A. (1987). Plans and situated actions: The problem of human-machine com-
munication. Cambridge: Cambridge University Press.
Tamir, P. (1989). Training teachers to teach effectively in the laboratory. Science Educa-
tion, 73, 59-69.
Tobin, K. (1990a). Teacher mind frames and science learning. In K . Tobin, J.B. Kahle, &
B .J. Fraser (Eds.). Windows into science classrooms: Problems associated with high level
cognitive learning in science (pp. 33-86). London: Falmer Press.
Tobin, K. (1990b). Research on science laboratory activities: In pursuit of better questions
and answers to improve learning. School Science and Mathematics, 90, 403-418.
A CONSTRUCTIVIST PHYSICS LAB 223

Tobin, K., Espinet, M., Byrd, S.E., & Adams, D. (1988). Alternative perspectives of
effective science teaching. Science Education, 72, 433-45 1 .
Tobin, K., & Gallagher, J.J. (1987). What happens in high school science classrooms'?
Journal of Curriculum Studies, 19, 549-560.
Tobin, K.G., & Ulerick, S.J. (1989, March). An interpretation of high school science
teaching based on metaphors and beliefs of social roles. Paper presented at the annual meeting
of the American Educational Research Association, San Francisco, CA.
Toulmin, S. (1982). The construal of reality: Criticism in modern and post-modern science.
Critical Inquiry, 9, 93- 1 1 1 .
Traweek, S. (1988). Beamtimes and lijetimes: The world of high energy physicists. Cam-
bridge, MA: MIT Press.
Turkle, S., & Papert, S. (1991). Epistemological pluralism and the revaluation of the
concrete. In I. Hare1 & S. Papert (Eds.), Constructionism: Research reports and essays, 1985-
1990 (pp. 161-191). Norwood, NJ: Ablex.
von Glasersfeld, E. (1987). Learning as a constructive activity. In C. Janvier (Ed.), Prob-
lems of representation in the teaching and learning of mathematics (pp. 3-17). Hillsdale, NJ:
Lawrence Erlbaum Associates.
Watson, J.D. (1968). Double helix. New York: Atheneum Publications.
Welch, W.W., Klopfer, L.E., Aikenhead, G.S., & Robinson, J.T. (1981). The role of
inquiry in science education: Analysis and recommendation. Science Education, 65, 33-55.
Wheatley, G.H. ( 199I). Constructivist perspectives on science and mathematics learning.
Science Education, 75, 9-2 1.
Wiggins, G. (1989). The futility of trying to teach everything of importance. Educational
Leadership, 46(9), 44-48, 57-59.
Zenzen, M., & Restivo, S. (1982). The mysterious morphology of immiscible liquids: A
study of scientific practice. Social Science Information, 21, 447-473.

Manuscript accepted March 18, 1993.

You might also like